Skip to main content

Full text of "Spaceflight revolution : NASA Langley Research Center from Sputnik to Apollo"

See other formats


SPACEFIIGHT 
REVOLUTION 


APACHE  JUNCTION  PUBLIC  LIBRARY 


3  9971  00211  4624 


NASA  Langley  Research  Center 
From  Sputnik  to  Apollo 


From  the  collection  of  the 

j  d 

•7       n 

•n   -r   m 

o  Jrrelinger 

v 

t      w  ^ 


San  Francisco,  California 
2007 


DO  .NOT 


JUN  1  4  1995 


v/ui  I  witl 


SPACEFLIGHT 
REVOLUTION 


L-66-6399 

The  "picture  of  the  century"  was  this  first  view  of  the  earth  from  space. 
Lunar  Orbiter  I  took  the  photo  on  23  August  1966  on  its  16th  orbit  just 
before  it  passed  behind  the  moon.  The  photo  also  provided  a  spectacular 
dimensional  view  of  the  lunar  surface. 


NASA  SP-4308 


SPACEFLIGHT 
REVOLUTION 

NASA  Langley  Research  Center 
From  Sputnik  to  Apollo 


James  R.  Hansen 


The.  NASA  History  Series 


1995 

National  Aeronautics  and  Space  Administration 

Washington,  DC 


NASA  maintains  an  internal  history  program  for  two  principal  reasons: 
(1)  Sponsorship  of  research  in  NASA-related  history  is  one  way  in  which 
NASA  responds  to  the  provision  of  the  National  Aeronautics  and  Space 
Act  of  1958  that  requires  NASA  to  "provide  for  the  widest  practicable  and 
appropriate  dissemination  of  information  concerning  its  activities  and  the 
results  thereof."  (2)  Thoughtful  study  of  NASA  history  can  help  agency 
managers  accomplish  the  missions  assigned  to  the  agency.  Understanding 
NASA's  past  aids  in  understanding  its  present  situation  and  illuminates 
possible  future  directions.  The  opinions  and  conclusions  set  forth  in  this 
book  are  those  of  the  author;  no  official  of  the  agency  necessarily  endorses 
those  opinions  or  conclusions. 


On  the  cover:  Langley's  innovative  Little  Joe  rocket  streaks  into  space  from  its  launchpad 
at  Wallops  Island,  Virginia,  on  4  October  1959,  two  years  to  the  day  after  the  historic  first 
orbit  of  the  Soviet  Sputnik  1. 


Library  of  Congress  Cataloging-in-Publication  Data 

Hansen,  James  R. 

Spaceflight  revolution  :  NASA  Langley  Research  Center  from 
Sputnik  to  Apollo  /  James  R.  Hansen. 

p.     cm.  --  (NASA  history  series)    (NASA  SP  ;  4308) 
Includes  bibliographical  references  and  index. 
1.  Langley  Research  Center — History.    2.  Astronautics — United 
States— History.     I.  Title.    II.  Series.    III.  Series:  NASA  SP  ;  4308. 
TL521.312.H36    1995 

629.4'0973— dc20  94-24592 

CIP 


For   sale   by   the    U.S.    Government    Printing    Office,    Superintendent    of 
Documents,  Mail  Stop:    SSOP,  Washington,  DC  20402-9328 


Contents 


Illustrations -.. » ^ . . ,   - ix 

Foreword       xv 

Acknowledgments xvii 

Prologue xxv 

1.  The  Metamorphosis 1 

The  Venerable  Order  of  the  NACA 3 

Glennan:  Welcome  to  NASA 11 

Air  versus  Space 15 

The  Public  Eye 22 

2.  The  First  NASA  Inspection 27 

Following  the  NACA  Way 33 

Project  Mercury 37 

Big  Joe,  Little  Joe      46 

3.  Carrying  Out  the  Task 51 

A  Home  at  Langley 55 

The  Tracking  Range 63 

Shouldering  the  Burden 69 

The  End  of  the  Glamour  Days 76 

4.  Change  and  Continuity 81 

The  Organization 85 

Thompson's  Obscurantism 91 

The  Sinking  of  Hydrodynamics — and  Aeronautics?      ....  93 

Growth  Within  Personnel  Ceilings      102 

The  Shift  Toward  the  Periphery      107 

Contracting  Out 109 

The  Brave  New  World  of  Projects 112 

Uncharted  Territory 117 

5.  The  "Mad  Scientists"  of  MPD 121 

The  ABCs  of  MPD 121 

The  Solar  Wind  Hits  Home      122 

The  MPD  Branch  .  126 


Spaceflight  Revolution 

Out  of  the  Tunnel •  ...  132 

Into  the  Cyanogen  Fire 138 

The  Barium  Cloud  Experiment 142 

The  Search  for  Boundless  Energy 147 

A  Hot  Field  Cools  Off 150 

6.  The  Odyssey  of  Project  Echo      153 

The  International  Geophysical  Year  and  the  V-2  Panel  .    .    .  156 

O'Sullivan's  Design 159 

Extraterrestrial  Relays 162 

Finessing  the  Proposal 164 

The  "Sub-Satellite" 166 

Something  the  Whole  World  Could  See 170 

Big  Ideas  Before  Congress 175 

Assigning  Responsibilities 177 

Shotput 179 

A  Burst  Balloon 185 

"Anything's  Possible!" 187 

Reflections 189 

The  Hegemony  of  Active  Voice 193 

7.  Learning  Through  Failure:  The  Early  Rush  of  the  Scout 

Rocket  Program 197 

"Itchy"  for  Orbit 197 

Little  Big  Man 200 

Little  Foul-Ups 205 

"3-2-1  Splash" 209 

Recertification 214 

An  Unsung  Hero 216 

Postscript 219 

8.  Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous 

Concept 221 

Brown's  Lunar  Exploration  Working  Group      222 

Michael's  Paper  on  a  "Parking  Orbit" 226 

The  Rendezvous  Committees 230 

Houbolt  Launches  His  First  Crusade      233 

The  Feelings  Against  LOR 237 

The  Early  Skepticism  of  the  STG 241 

Mounting  Frustration 245 

President  Kennedy's  Commitment      248 

Houbolt 's  First  Letter  to  Seamans      249 

A  Voice  in  the  Wilderness 257 

The  LOR  Decision 260 

Postscript 267 


VI 


Contents 

9.  Skipping  "The  Next  Logical  Step" 269 

"As  Inevitable  as  the  Rising  Sun" 271 

The  First  Space  Station  Task  Force 274 

From  the  Inflatable  Torus  to  the  Rotating  Hexagon    ....  277 

Betwixt  and  Between 286 

Manned  Orbital  Research  Laboratory 293 

Keeping  the  "R"  Alive 301 

Understanding  Why  and  Why  Not      305 

Lost  in  Space? 307 

10.  To  Behold  the  Moon:  The  Lunar  Orbiter  Project 311 

The  "Moonball"  Experiment 315 

Initiating  Lunar  Orbiter 319 

Project  Management      321 

The  Source  Evaluation  Board 326 

Nelson's  Team 332 

The  Boeing  Team 334 

The  "Concentrated"  versus  the  "Distributed"  Mission    ...  336 

"The  Picture  of  the  Century" 344 

Mission  More  Than  Accomplished 346 

Secrets  of  Success 350 

11.  In  the  Service  of  Apollo 355 

Langley's  "Undercover  Operation"  in  Houston 357 

The  Dynamics  of  Having  an  Impact 361 

Inside  the  Numbers 366 

The  Simulators 369 

Rogallo's  Flexible  Wing 380 

The  Apollo  Fire  Investigation  Board 387 

12.  The  Cortright  Synthesis 393 

The  Stranger 394 

The  Reorganization 401 

New  Directions 413 

Critique  from  the  Old  Guard 418 

Epilogue   ..,.., 427 

Abbreviations      441 

Notes    ...... 447 

Index 519 

The  Author 537 

The  NASA  History  Series 539 

vii 


Illustrations 


Earth  as  photographed  by  Lunar  Orbiter  /,  1966 ii 

Dwight  D.  Eisenhower,  1958 3 

Map  of  Tidewater  Virginia,  1930s 5 

Floyd  L.  Thompson 6 

Langley  Aircraft  Manufacturers'  Conference,  1934 6 

NACA  Main  Committee,  1929 8 

Henry  J.  E.  Reid,  Vannevar  Bush,  and  George  W.  Lewis 9 

George  W.  Lewis  and  Hugh  L.  Dryden 11 

T.  Keith  Glennan,  1958 13 

T.  Keith  Glennan  and  Henry  J.  E.  Reid,  1959  14 

NACA  test  pilot  Paul  King,  1925 16 

Variable-Density  Wind  Tunnel,  1922 18 

Bell  P-59  Peashooter  in  Full-Scale  Tunnel,  1944 18 

Swallow  arrow- wing  model  in  16- Foot  Transonic  Tunnel,  1959  ...  21 

X-15  model  in  7  x  10-Foot  High-Speed  Tunnel,  1958  21 

Aerial  photo  of  Langley,  1950 25 

Mercury  exhibit  at  NASA's  First  Anniversary  Inspection,  1959  ...  29 

Ira  H.  Abbott  and  Henry  Reid,  1959  29 

T.  Keith  Glennan  and  Floyd  L.  Thompson,  1959 29 

Walter  Bonney  and  T.  Keith  Glennan,  1959 31 

Full-size  mock-up  of  the  X-15,  1959 32 

John  Stack  and  Axel  Mattson  35 

Goddard's  exhibit  at  NASA's  First  Anniversary  Inspection, 

1959 36 

Robert  R.  Gilruth  ...... 38 

Diagram  of  Mercury  mission  concept  39 

The  Mercury  astronauts 40 

Molded  couches  for  Mercury  capsule 44 

Diagram  of  Mercury  capsule 44 

John  Glenn  inside  Mercury  capsule 45 

John  Glenn  and  Annie  Castor  Glenn,  1959 46 

Little  Joe  capsules  constructed  in  Langley  shops 48 

Little  Joe  on  the  launchpad  at  Wallops  Island  48 

Little  Joe  blasting  off  from  Wallops  Island,  1959  49 

Little  Joe  capsule  recovered  at  sea  49 

Model  of  Mercury  capsule  in  Full-Scale  Tunnel,  1959  61 

Model  of  Mercury  capsule  in  7  x  10-Foot  High-Speed  Tunnel, 

1959  .  61 


IX 


Space/light  Revolution 

Model  of  Redstone  booster  in  the  Unitary  Plan  Wind  Tunnel, 

1959 62 

Impact  studies  of  Mercury  capsule  in  the  Back  River,  1960 62 

George  Barry  Graves,  Jr 67 

Layout  of  Project  Mercury  tracking  site 68 

John  A.  "Shorty"  Powers,  1962 78 

Walter  M.  Schirra,  1962 78 

Robert  R.  Gilruth  and  the  mayor  of  Newport  News,  1962 79 

John  Glenn  and  his  wife,  Annie,  1962 79 

Floyd  L.  Thompson,  1963 83 

Floyd  L.  Thompson,  James  E.  Webb,  and  John  F.  Victory 84 

Langley  organization  chart,  1962 87 

Clinton  E.  Brown,  Eugene  C.  Draley,  and  Laurence  K. 

Loftin,  Jr 88 

Langley's  top  staff  members  greet  Raymond  Bisplinghoff 90 

Aerial  view  of  the  Full-Scale  Tunnel  and  Tank  No.  1,  1959 95 

X-20  Dyna-Soar  model  in  Tank  No.  2,  1961  96 

Aeronautics  and  Space  Work  as  Percentages  of  Langley's  Total  Effort, 

1957-1965,  table  97 

John  Stack,  1959 99 

Scale  model  of  the  General  Dynamics  F-lll A 100 

Model  of  SCAT  15F  in  Unitary  Plan  Wind  Tunnel 101 

Number  of  Paid  Employees  at  NASA  Langley,  1952-1966, 

graph 103 

Paid  Employees  at  NASA  Langley  as  Percentage  of  NASA  Total, 

1958-1968,  graph 103 

Kitty  O'Brien- Joyner,  1964 105 

Langley's  women  scientists,  1959 105 

Langley's  computer  complex,  1959  Ill 

Scale  model  of  WS-110A  in  7  x  10-Foot  High-Speed  Tunnel  ....  117 

Schematic  drawings  of  the  Van  Allen  radiation  belts 124 

John  V.  Becker,  1961 128 

The  Continuous-Flow  Hypersonic  Tunnel  128 

Macon  C.  Ellis,  1962  129 

Paul  W.  Huber  and  Marc  Feix 131 

Philip  Brockman  and  the  MPD-arc  plasma  accelerator,  1964  ....  134 

George  P.  Wood,  1962 136 

The  accelerator  section  of  the  20-megawatt  plasma  accelerator 

facility 136 

Charlie  Diggs  and  an  early  version  of  a  Hall-current  plasma 

accelerator 137 

Langley's  Hall-current  plasma  accelerator,  1965 137 

Robert  V.  Hess,  1962 139 

Langley's  cyanogen  burner 141 

Concept  for  a  Mars  landing  vehicle 151 


Illustrations 

Failed  deployment  of  Echo  test 155 

William  J.  O'Sullivan  and  his  family,  1961 157 

30-inch  Sub-Satellite  168 

Heat  test  of  30-inch  Sub-Satellite 168 

Jesse  Mitchell,  1958 171 

Folded  Beacon  satellite  174 

William  J.  O'Sullivan  and  the  Beacon  satellite 174 

William  J.  O'Sullivan,  1958  175 

Walter  Bressette  and  prototype  of  the  satelloon 175 

Echo  /container 181 

Inflation  of  Echo  I  in  Weeksville,  N.C 182 

Edwin  Kilgore  and  Norman  Crabill 183 

The  Echo  I  team  and  inflated  Echo  I 183 

Will  Taub  and  James  Miller  assembling  Shotput  launch  vehicle  .  .  .  184 

Shotput  ready  for  launch 184 

Explorer  24  192 

Pageos  satelloon 193 

Scout  on  launchpad,  Wallops  Island 201 

James  R.  Hall,  1961 203 

LTV  Scout  team,  1967 204 

Spectators  at  Wallops  Island  rocket  launch 206 

The  first  Scout  launch,  1  July  1960 207 

Scout  launch  control  building,  Wallops 208 

Scout  control  room,  Wallops 208 

Eugene  D.  Schult,  1963 211 

Scout  launch,  30  June  1961 212 

Scout  launch,  1  March  1962 213 

Launchpad  damaged  by  Scout,  20  July  1963 213 

Vought  Astronautics  technicians  assemble  Scout  214 

San  Marco  launch  operation 218 

San  Marco's  floating  platform  218 

Committees  Reviewing  Lunar  Landing  Modes,  table 225 

Clinton  E.  Brown,  William  H.  Michael,  Jr.,  and  Arthur  Vogeley, 

1989 ~. 227 

John  D.  Bird,  1962  .  .  ' 229 

Sketch  "To  the  Moon  with  C-l's  or  Bust" 229 

Houbolt's  Early  Crusades,  table  234 

Houbolt's  Later  Crusades,  table 240 

Early  version  of  a  lunar  excursion  module 243 

John  C.  Houbolt,  1962 244 

John  C.  Houbolt  explaining  lunar-orbit  rendezvous  scheme 247 

Viewgraph  comparing  the  propulsion  steps  of  the  three  lunar  mission 

modes 254 

Comparative  sizes  of  manned  mission  rockets 256 

Comparison  of  lander  sizes 256 


XI 


Space/light  Revolution 

George  M.  Low 259 

Wernher  von  Braun  at  Langley 264 

Life  magazine  cover  featuring  the  lunar  excursion  module 266 

Rejected  Life  cover  of  John  C.  Houbolt 266 

75-foot-diameter  rotating  hexagon 273 

Paul  R.  Hill  and  Robert  Osborne,  1962  275 

Rene  Berglund,  1962  278 

Early  space  station  configurations 278 

Inflation  of  the  full-scale  model  of  the  inflatable  torus 279 

Floyd  L.  Thompson,  James  Webb,  and  T.  Melvin  Butler  with  the 

24-foot  inflatable  torus  279 

24-foot  inflatable  torus  280 

10- foot-diameter  scale  model  of  torus 282 

Zero  gravity  mock-up  of  the  24-foot-diameter  torus 282 

Drawing  of  winning  rotating  hexagonal  configuration 284 

Model  of  rotating  hexagon  assembled  and  collapsed 284 

Douglas  MORL  baseline  configuration 296 

Cross  section  of  interior  of  Douglas  MORL 296 

MORL  illustration  from  Douglas  manual 298 

Test  of  space  station  portal  air  lock 299 

William  N.  Gardner  explains  model  of  the  MORL 300 

MORL-Saturn  IB  model  in  8-Foot  Transonic  Tunnel 302 

Otto  Trout,  1966 303 

Integrative  Life  Support  System  arrives  at  Langley 304 

Integrative  Life  Support  System  in  Building  1250 304 

William  N.  Gardner,  1966 308 

Lunar  Orbiter  above  the  lunar  surface 314 

Associate  Director  Charles  J.  Donlan 317 

Structural  dynamics  testing  for  lunar  landing 318 

Lunar  Orbiter  III  photo  of  Kepler  crater 318 

Israel  Taback  and  Clifford  H.  Nelson,  1964 323 

Eastman  Kodak  dual-imaging  camera  system 329 

Lee  R.  Scherer 330 

Signing  of  the  Lunar  Orbiter  contract,  1964  331 

Clifford  H.  Nelson  and  James  S.  Martin 333 

Floyd  L.  Thompson  and  George  Mueller,  1966 337 

The  Lunar  Orbiter  area  of  interest  337 

Typical  flight  sequence  of  Lunar  Orbiter 340 

Lunar  Orbiter  with  labeled  components 343 

Final  inspection  of  Lunar  Orbiter  I 343 

Lunar  Orbiter  I  liftoff 344 

Lunar  Orbiter  team  displays  first  photo  of  the  earth  from  deep 

space 345 

The  dark  side  of  the  moon 347 

The  lunar  surface,  Copernicus  crater  348 


xn 


Illustrations 

The  lunar  surface,  Tycho  crater 349 

The  "Whole  Earth"  as  photographed  by  Lunar  Orbiter  V 352 

Floyd  L.  Thompson  and  James  E.  Webb,  1961 358 

Axel  Mattson,  Robert  R.  Gilruth,  Charles  Donlan, 

and  Donald  Hewes  361 

Impact  tests  of  the  Apollo  capsule  364 

Project  Fire  wind-tunnel  test 367 

Number  of  Langley  Research  Projects  Directly  Related  to  Apollo 

Program,  1962-1968,  table 368 

The  Langley  Rendezvous  and  Docking  Simulator 372 

Time-lapse  sequence  of  a  docking  on  the  Rendezvous  Simulator  .  .  .  372 

A  pilot  eyeballing  a  rendezvous  on  the  simulator 372 

Donald  Hewes  and  William  Hewitt  Phillips 374 

Langley  Lunar  Landing  Research  Facility  375 

Early  LEM  used  with  the  Lunar  Landing  Research  Facility 376 

LEM  control  cab,  the  Lunar  Landing  Research  Facility  376 

LEM  "in  flight"  using  the  Lunar  Landing  Research  Facility 376 

Time-lapse  sequence  of  an  LEM  landing  using  the  simulator  ....  377 

Modeled  floor  of  the  Lunar  Landing  Research  Facility 377 

Walter  Cronkite  using  the  Reduced  Gravity  Walking  Simulator  .  .  .  378 

Lunar  Orbit  and  Letdown  Approach  Simulator 380 

Francis  and  Gertrude  Rogallo,  1963 382 

Test  flight  of  the  Parasev  382 

Parasev  in  Langley's  Full-Scale  Tunnel 386 

Floyd  L.  Thompson  and  Thomas  O.  Paine,  1968 388 

Gus  Grissom  in  the  Rendezvous  and  Docking  Simulator,  1963  .  .  .  390 
Roger  Chaffee  using  the  Reduced  Gravity  Walking  Simulator, 

1965 390 

Neil  Armstrong  and  the  staff  of  the  Lunar  Landing  Research  Facility, 

1967 392 

Cortright  appointment  announced  in  the  Langley  Researcher  ....  396 

Langley  old  guard  welcomes  Cortright,  1968 398 

Edgar  M.  Cortright,  1970 399 

Cortright  speaks  in  the  Morale  Activities  building 399 

Organization  of  Langley,  1970,  chart 405 

Viking  Lander  model  in  Langley  wind  tunnel,  1970 406 

View  of  Mars  from  Viking  Orbiter  1,  1976 407 

The  Viking  Lander  2  on  the  Martian  surface,  1976  407 

John  E.  Duberg,  George  M.  Low,  and  Edgar  M.  Cortright, 

1970 409 

Organization  changes  announced  in  the  Langley  Researcher,  1970  .  .  410 

Model  of  the  Boeing  737  in  the  Anechoic  Antenna  Test  Facility  ...  416 

Richard  H.  Petersen  in  the  National  Transonic  Facility,  1984  ....  417 

The  National  Transonic  Facility 417 


Xlll 


Foreword 


James  R.  Hansen  has  impeccable  credentials  as  a  thorough,  perceptive 
investigator  and  writer  of  technological  history.  His  accomplishments  in  the 
field  are  outstanding,  as  exemplified  by  his  book  Engineer  in  Charge,  which 
was  published  in  1987.  This  book  presents  a  careful  analysis  of  the  history 
of  the  Langley  Memorial  Aeronautical  Laboratory  of  the  National  Advisory 
Committee  for  Aeronautics  (NACA)  from  its  formation  in  1917  to  the  demise 
of  the  NACA  in  October  1958  when  this  prestigious  organization  became 
the  centerpiece  of  the  new  National  Aeronautics  and  Space  Administration 
(NASA).  Whereas  the  NACA  was  concerned  primarily  with  aeronautical 
research  conducted  by  government  employees  in  its  own  laboratories,  NASA 
would  have  a  much  broader  charter  that  included  not  only  aeronautical  and 
space  research  but  also  the  development  and  operation  of  various  types  of 
space  vehicles,  including  manned  vehicles.  Within  this  new  organization, 
the  Langley  Aeronautical  Laboratory  became  the  Langley  Research  Center 
of  NASA. 

As  a  part  of  NASA,  Langley  underwent  many  profound  changes  in 
program  content,  organization  and  management,  and  areas  of  personnel 
expertise.  Although  aeronautical  research  continued  in  the  NASA  era, 
research  in  support  of  such  projects  as  Echo,  Scout,  Mercury,  Apollo,  and 
the  Space  Shuttle  occupied  a  larger  percentage  of  the  Langley  research  effort 
as  the  years  passed.  In  addition,  Langley  forged  into  new  fields  by  assuming 
management  responsibility  for  such  large  space  projects  as  Lunar  Orbiter 
and  Viking.  This  responsibility  involved  major  contract  activities  and 
support  of  in-house  research.  New  research  facilities,  such  as  large  vacuum 
tanks  and  high-speed  and  high-temperature  air  jets  capable  of  simulating 
atmospheric  entry  from  space,  were  developed  and  constructed. 

Although  many  new  personnel  were  eventually  hired,  large  numbers  of 
the  existing  Langley  complement  easily  made  the  transition  to  space-related 
research  and  thus  showed  that  a  proficient  research  professional  could  shift 
without  too  much  difficulty  into  new  fields  of  technical  endeavor.  For  ex- 
ample, in  orbital  mechanics  and  space  rendezvous,  individuals  who  had 
previously  worked  in  such  diverse  disciplines  as  theoretical  aerodynamics, 
high-speed  propellers,  and  aeroelasticity  quickly  became  expert  and  as- 
sumed roles  of  national  leadership.  A  well-known  case  is  found  in  the  ac- 
tivities of  Dr.  John  C.  Houbolt,  an  expert  in  aeroelasticity  and  dynamic 
loads,  who  became  a  leading  proponent — according  to  Hansen,  perhaps 
the  key  proponent — of  Lunar  Orbit  Rendezvous  as  the  preferred  means  of 


xv 


Space/light  Revolution 

accomplishing  the  Apollo  lunar  landing  mission.  This  technique,  of  course, 
turned  out  to  be  incredibly  successful. 

A  very  unsettling  aspect  of  the  transition  of  Langley  in  the  1958-1975 
period  was  the  replacement  of  the  director,  longtime  Langley  engineer  Floyd 
L.  Thompson,  with  Edgar  M.  Cortright.  Cortright  came  from  NASA 
headquarters  and  had  had  prior  research  experience  at  the  NACA  Lewis 
Flight  Propulsion  Laboratory  (later  designated  as  the  NASA  Lewis  Research 
Center).  In  the  Cortright  regime,  along  with  many  significant  changes 
in  center  organization  and  management,  there  came  a  closer,  and  many 
thought  an  undesirable,  control  of  Langley  programs  by  a  centralized  NASA 
management. 

James  Hansen's  new  book,  Space/light  Revolution,  covers  the  turbulent 
seventeen-year  period  from  1958-1975  in  great  and  interesting  detail.  With 
his  usual  thoroughness,  Hansen  has  based  this  book  on  careful  analysis 
of  hundreds  of  written  records,  both  published  and  unpublished,  as  well 
as  on  numerous  personal  interviews  with  many  of  the  key  individuals 
involved  in  the  great  transition  at  Langley.  One  Langley  activity  that 
was  intentionally  omitted  from  this  study  is  aeronautical  research  which, 
as  the  author  mentions,  will  hopefully  be  covered  in  a  separate  book. 
Space/light  Revolution  is  a  very  complete  and  well-researched  exposition  and 
interpretation  of  a  period  of  great  change  at  the  Langley  Research  Center. 
The  main  events  and  trends  are  clearly  and  succinctly  presented.  Although 
many  who  worked  for  Langley  during  the  period  covered  may  not  agree 
entirely  with  some  of  Hansen's  interpretations  and  conclusions,  sufficient 
information  is  given  in  the  text,  references,  and  notes  to  permit  the  reader 
to  evaluate  the  work.  In  any  event,  anyone  who  ever  worked  for  Langley  or 
NAC A/NASA  or  who  has  any  interest  in  the  history  of  technology  will  find 
the  book  fascinating  and  thought  provoking.  In  addition,  anyone  interested 
in  the  present  and  the  future  of  NASA  and  the  American  space  program  will 
want  to  pay  close  attention  to  the  insights  found  in  his  epilogue.  Readers 
will  see  that  Jim  Hansen  has  again  demonstrated  his  great  abilities  as  a 
historian,  and  he  deserves  a  well-earned  "Thank  you"  for  creating  what  will 
no  doubt  prove  to  be  an  enduring  classic. 


November  1994  Laurence  K.  Loftin,  Jr. 

Director  for  Aeronautics  (Retired) 
NASA  Langley  Research  Center 


xvi 


Acknowledgments 


In  writing  this  book,  I  am  indebted  not  only  to  the  many  talented  and 
caring  people  who  have  helped  my  project  in  one  way  or  another  in  the  past 
seven  years  but  also  to  a  seminal  event  of  my  adolescence  that  has  fed  my 
adult  interest  and  colored  my  historical  perspective  on  what  I  now  see  to 
have  been  "the  spaceflight  revolution"  of  the  late  1950s  and  1960s. 

People  all  over  the  world  have  their  personal  stories  to  tell  about  what 
they  were  doing  and  thinking  when  they  first  spotted  a  mysterious  object 
in  the  night  sky.  For  many,  these  stories  involve  Sputnik  because  it  was 
the  first  man-made  object  to  be  observed.  But  for  those,  like  myself,  who 
were  too  young  to  be  stargazing  in  1957,  the  stories  often  involve  the  Echo 
balloon,  NASA's  first  communications  satellite.  Stories  about  both  objects 
may  indeed  relate  to  Sputnik  because  it  was  our  hysterical  reaction  to  the 
Soviet  satellite  that  tempered  our  feelings  about  objects  in  space  for  some 
time  to  come. 

For  me,  the  memory  of  my  first  satellite  sighting  is  still  vivid.  One 
sultry  evening  in  mid- August  1960  while  I  was  serving  as  the  batboy  for  my 
brother's  Little  League  team  in  Fort  Wayne,  Indiana,  something  unusual  and 
a  little  unnerving  took  place.  About  halfway  through  the  game,  I  noticed 
that  fans  in  the  bleachers  were  no  longer  watching  the  game,  but  instead 
were  standing,  looking  at  the  sky,  and  pointing  at  something.  When  our 
team  was  in  the  field  and  my  batboy  duties  were  temporarily  over,  I  found 
my  mother  in  the  crowd  and  asked  her  what  the  fuss  was  all  about.  She 
said  she  had  heard  someone  in  the  crowd  call  it  "Echo."  She  reassured  me 
that  it  was  nothing  to  be  afraid  of,  as  it  "belonged  to  us." 

But  who  exactly  was  "us,"  I  wondered?  To  an  eight-year-old  in  1960,  "us" 
meant  human  beings  or  "earthlings" ;  "them"  meant  "aliens."  I  was  glad  to 
hear  from  my  mother  that  the  bright  little  light  that  I  now,  too,  spotted 
moving  so  slowly  yet  perceptibly  in  the  heavens  did  not  mean  "they"  were 
coming  to  get  me,  but  I  was  still  concerned.  Even  at  eight,  I  was  informed 
enough  about  what  was  going  on  in  the  world  to  know  that  "us"  and  "them" 
also  meant  something  else  almost  as  sinister  as  earthlings  versus  aliens.  "Us" 
meant  "Americans"  and  "them"  meant  "Russians,"  and  somehow  I  knew 
that  it  was  better  for  us  to  have  put  something  up  into  the  sky  for  the  world 
to  look  at  than  it  was  for  them  to  have  done  it.  Whether  I  knew  that  they 
in  fact  had  already  done  it  some  three  years  earlier,  I  really  cannot  say.  I  do 
remember  being  so  entranced  by  the  man-made  star  that  I  had  to  be  told 
more  than  once  by  the  coach  of  our  Little  League  team  to  "get  my  head  in 
the  game"  and  go  out  and  pick  up  the  baseball  bats. 


xvn 


Spaceflight  Revolution 

The  next  night,  as  soon  as  it  started  getting  dark,  my  entire  family 
headed  to  the  backyard  to  look  for  Echo,  only  to  find  that  parents  all  over  our 
neighborhood  were  leading  their  children  to  hunt  for  the  artificial  star.  This 
time  my  feelings  about  the  bright  dot  of  light  moving  so  clearly  across  the  sky 
were  more  positive.  We  were  moving  out  into  space.  Like  the  morning  paper 
had  said,  Echo  was  "the  visible  symbol  of  American  creativity  for  all  the 
world  to  see."  In  the  next  several  weeks,  a  number  of  library  books  about 
space  would  come  home  from  school  with  me.  For  me,  too,  a  spaceflight 
revolution  had  begun. 

As  I  grew  up,  so  did  the  American  space  program.  As  a  second-grader, 
near  the  end  of  the  school  year  that  followed  the  summer  of  Echo  1,  I  sat 
on  the  wooden  floor  of  a  gymnasium  with  all  the  other  kids  in  my  school 
and  watched  shadowy  black-and-white  television  pictures  of  the  suborbital 
flight  of  Mercury  astronaut  Alan  Shepard.  Gus  Grissom's  suborbital  flight 
came  next;  I  watched  it  at  home  while  on  vacation  that  July.  Then 
came  John  Glenn's  historic  orbital  flight  in  February  1962  and  a  return  to 
TV-watching  from  telescopic  distance  on  the  school  gym  floor. 

After  that,  my  memory  of  NASA's  space  missions  is  cloudy  and  does 
not  sharpen  again  until  December  1968,  when  with  the  crew  of  Apollo  8, 
my  family  and  I  spent  Christmas  Eve  circling  the  lunar  sphere,  seeing  awe- 
inspiring  pictures  of  the  moon's  surface,  and  listening  to  the  astronauts 
conclude  their  TV  broadcast  with  "Merry  Christmas  and  God  bless  all  of 
you — all  of  you  on  the  good  earth."  I  also  clearly  remember  July  1969, 
when  the  Apollo  11  lunar  module  Eagle  landed  on  the  Sea  of  Tranquility 
and  Neil  Armstrong  took  that  first  "small  step  but  one  giant  leap"  onto 
another  heavenly  body. 

These  wondrous  events  of  the  space  age  made  a  big  impression  on  me, 
as  they  did  in  one  way  or  another  on  nearly  every  human  being  alive  at 
the  time.  But  no  space  event  ever  surpassed  that  first  sighting  of  the  Echo 
balloon,  glittering  like  a  diamond  over  the  baseball  field. 

For  a  while,  mostly  on  warm  summer  evenings,  I  continued  to  look  for 
Echo  and  for  other  objects  moving  mysteriously  through  the  sky.  But 
gradually,  I  lost  almost  all  interest  in  space.  A  child  of  the  Age  of  Aquarius 
and  the  Vietnam  War,  I  wondered,  like  so  many  others  did  at  that  time  why, 
if  we  could  put  a  man  on  the  moon,  we  couldn't  do  so  many  other  things. 
Only  much  later  would  I  begin  to  look  up  again,  seeking  Echo,  perhaps 
trying  to  find  lost  innocence  and  youth.  Little  did  I  know  in  1960  that 
30  years  later  I  would  reexperience  the  orbits  of  Echo  and  write  a  detailed 
history  of  the  satelloon's  genesis,  as  I  have  in  chapter  6  of  this  book. 

Whatever  the  object  of  fixation,  be  it  Sputnik  or  Echo,  stories  like  mine 
represent  an  illuminating  cultural  expression  of  the  young  space  age.  It  was 
with  our  stirring  personal  experiences  of  these  moving  little  lights  in  the 
night  sky  that  the  spaceflight  revolution  began.  As  one  young  Canadian  girl 
wrote  to  NASA  in  1968  in  a  poem  entitled  "To  a  Falling  Star,"  on  the  eve 


xvm 


Acknowledgments 

of  Echo  fs  falling  back  to  its  destruction  into  the  atmosphere,  "Thanks  for 
making  me  look  up." 


Many  times,  in  thanking  all  the  people  who  have  helped  in  the  research 
and  writing  of  a  book,  an  author  waits  until  the  end  of  the  acknowledgments 
to  thank  his  own  family  for  their  love  and  support.  But  in  this  case,  I  want 
to  thank  my  family  first.  My  wife,  Peggy,  and  my  two  children,  Nathaniel 
and  Jennifer,  have  been  last  too  many  times  in  the  seven  years  it  took  me 
to  research  and  write  this  book  to  be  last  once  again.  I  was  away  from 
them  and  at  NASA  Langley  in  Virginia  for  most  of  every  summer  from  1987 
to  1993  writing  this  book.  This  means  we  all  sacrificed  and  missed  each 
other  a  lot.  Summertime  experiences  my  wife  enjoyed  with  the  children  at 
our  home  in  Alabama,  she  enjoyed  alone.  I  only  heard  about  them  in  our 
many  long-distance  telephone  calls.  When  my  children  are  grown-up  and 
gone  from  home,  I  am  sure  I  will  regret  what  I  missed  with  them  even  more 
intensely. 

I  would  also  like  to  thank  Charles  and  Robert  Stanton.  I  spent  my 
summers  from  1987  to  1993  in  their  respective  homes  in  Hampton,  Virginia, 
and  I  enjoyed  those  times  (especially  the  golf  games)  tremendously.  I  am 
sure  that  Charlie  and  Bob  heard  much  more  about  NASA  history  than  they 
ever  cared  to,  but  they  never  let  on.  My  friendships  with  Sharon  Buchanon 
and  Rick  Thompson  while  at  Bob  Stanton's  also  kept  me  from  being  too 
lonely,  as  they  were  a  regular  part  of  my  Hampton  "family."  Dr.  Fereidoun 
"Feri"  Farassat,  a  remarkable  person  and  accomplished  acoustical  scientist 
at  NASA  Langley,  was  also  a  valued  companion.  I  have  learned  a  great 
deal  from  him  about  science  and  technology,  but  what  I  most  cherish  is  his 
friendship. 

Steve  Corneliussen,  a  talented  writer  from  Poquoson,  Virginia,  who 
edited  my  book  Engineer  in  Charge  and  who  then  became  one  of  my  closest 
friends,  has  contributed  immensely  to  my  perspective  on  aerospace  history 
and  life  in  general.  Over  the  years  Steve  has  given  me  constant,  generous 
encouragement  and  good  advice.  I  regret  that  he  was  so  busy  with  his  work 
at  the  Continuous  Electron  Beam  Accelerator  Facility  (CEBAF)  in  Newport 
News,  Virginia,  that  he  could  not  serve  once  again  as  my  book  editor. 

But  how  lucky  I  was  to  have  Kathy  Rawson,  of  Williamsburg,  Virginia, 
edit  this  manuscript.  Kathy  has  done  many  wonderful  things  for  this  book, 
turning  an  overly  long  and  in  some  essential  ways  ailing  manuscript  into  a 
much  healthier  one.  Her  consummate  professionalism  and  her  friendly  words 
of  encouragement  inspired  me  to  keep  working  for  our  book's  improvement. 
In  particular,  Kathy  prodded  me  in  her  gentle  way  to  rewrite  what  was 
originally  a  weak  epilogue. 

As  Kathy  has  told  me,  many  other  people  associated  with  the  Research 
Publishing  and  Printing  Branch  at  NASA  Langley  came  together  as  a  team 

xix 


Spaceflight  Revolution 

to  see  this  book  to  its  completion.  In  particular,  I  wish  to  thank  Lynn 
Heimerl,  who  supervised  the  entire  project,  and  Mary  McCaskill,  the  branch 
head,  whose  strong  support  for  NASA  Langley's  major  investment  in  the 
production  of  this  book  is  sincerely  appreciated.  Others  involved  at  RPPB 
that  I  would  like  to  thank  individually  include  Nancy  Sheheen,  who  oversaw 
the  editing  and  typesetting  process;  Linda  Carlton,  who  formatted  and 
typed  the  majority  of  the  book;  Peggy  Overbey,  who  took  over  the  typing 
for  the  homestretch;  and  Sybil  Watson  and  Mary  Edwards,  who  diligently 
proofread  every  page. 

I  also  want  to  thank  the  staff  of  the  Floyd  L.  Thompson  Technical 
Library  at  NASA  Langley  for  their  strong  support  of  my  project,  notably 
H.  Garland  Gouger,  Jr.,  Jane  Hess  (retired),  Sue  Miller,  Sue  Seward 
(retired),  Susan  A.  Motley,  and  George  Roncaglia.  Also,  the  Photographies 
Section  at  Langley,  under  Alton  T.  Moore,  performed  yeoman's  service  for 
this  book  by  providing  excellent  prints  of  its  many  photographs.  I  am 
particularly  indebted  to  Frederick  D.  Jones  not  only  for  doing  much  of  the 
photo  lab  work  but  also  for  giving  me  access  to  a  number  of  pictures  from 
the  early  days  of  Project  Mercury,  many  of  which  he  took  on  his  own  time 
with  his  own  camera. 

Without  the  generous  support  and  personal  interest  of  Richard  T. 
Layman  of  the  Facilities  Program  Development  Office,  who  has  been  in 
charge  of  the  history  program  at  NASA  Langley  since  the  late  1970s,  this 
sequel  to  Engineer  in  Charge  surely  would  not  have  been  written.  Dick  has 
been  constantly  available  to  help  me  access  historical  materials  and  to  solve 
problems  associated  with  my  work  in  the  Langley  Historical  Archives.  Dick 
himself  started  work  at  Langley  in  the  early  1960s,  and  his  insights  into  the 
center's  history  proved  very  helpful. 

A.  Gary  Price  and  J.  Campbell  Martin  of  Langley's  Office  of  External 
Affairs  have  also  provided  tremendous  support  over  many  years  for  my  work 
as  the  Langley  historian,  as  have  Richard  H.  Petersen  and  Paul  F.  Holloway, 
the  Langley  center  directors  during  the  years  I  prepared  this  book.  I  came 
to  know  "Pete"  Petersen  particularly  well  and  wish  to  express  special  thanks 
to  him  for  his  genuine  interest  in  what  history  books  such  as  mine  can  offer 
to  NASA  management  and  the  public  at  large. 

And  then  there  are  the  "NACA  Nuts,"  the  dozens  of  men  and  women 
whom  I  first  got  to  know  while  researching  Engineer  in  Charge  and  came 
to  know  even  better  while  investigating  their  metamorphosis  into  "NASA 
Wizards."  I  wish  I  could  mention  all  of  them  by  name  but  must  focus  on 
the  few  whom  I  have  come  to  know  the  best:  John  V.  Becker,  William 
Boyer,  Clinton  E.  Brown,  Norman  Crabill,  Charles  J.  Donlan,  John  E. 
Duberg,  Macon  C.  "Mike"  Ellis,  Robert  R.  Gilruth,  Richard  Heldenfels, 
Jane  Hess,  Robert  Hess,  John  C.  Houbolt,  Vera  Huckel,  Kitty  O'Brien- 
Joyner,  Abraham  Leiss,  Axel  T.  Mattson,  William  A.  Michael,  Mark  R. 
Nichols,  W.  Hewitt  Phillips,  Edward  C.  Polhamus,  John  P.  "Jack"  Reeder, 
Joseph  A.  Shortal  (deceased),  William  Sleeman,  Israel  Taback,  Helen  Willey, 


xx 


Acknowledgments 

Herbert  A.  "Hack"  Wilson  (deceased),  Richard  T.  Whitcomb,  and  Charles  H. 
Zimmerman.  To  those  with  whom  I  talked  about  Langley's  history  but  have 
failed  to  name,  please  accept  my  apologies  and  sincere  thanks.  Getting  to 
know  all  of  you  was  the  best  thing  about  writing  this  book. 

I  need  to  single  out  Edgar  M.  Cortright,  another  Langley  director  (1969- 
1975)  and  a  major  player  in  the  history  examined  at  the  end  of  this  book, 
and  thank  him  for  the  long  and  comprehensive  interviews.  Dr.  Cortright 
withheld  very  little  from  my  tape  recorder,  and  for  that  I  sincerely  thank 
him.  I  hope  he  feels  that  I  have  treated  his  time  and  his  achievements  at 
Langley  fairly. 

Laurence  K.  Loftin,  Jr.,  the  author  of  this  book's  foreword,  also  deserves 
a  special  acknowledgment.  Over  the  course  of  my  14  years  as  Langley's 
official  (and  unofficial)  historian,  Larry  has  spent  hundreds  of  hours  with  me, 
talking  about  the  history  of  airplanes,  NACA  research,  and  the  transition 
from  the  NACA  to  NASA.  Much  of  my  perspective  about  all  these  things 
has  been  shaped  in  my  conversations  with  Larry.  I  owe  him  a  huge  debt  of 
gratitude,  not  only  because  he  has  saved  me  from  some  major  technical  and 
historical  blunders  but  also  because  he  and  his  wife,  Agnes,  came  to  treat 
me  over  the  years  almost  like  a  son.  Much  of  my  appreciation  for  what  it 
means  to  be  an  engineer  comes  from  the  time  I  spent  with  Larry. 

I  cannot  fail  to  mention  the  help  and  encouragement  given  to  me  freely 
by  my  colleagues  in  the  Department  of  History  at  Auburn  University,  a 
department  for  which  I  have  been  serving  as  chairman  since  my  election 
in  1993.  A  faculty  workshop  in  1991  took  a  very  critical  look  at  an  early 
draft  of  my  first  chapter,  thus  resulting  in  a  major  revision.  My  colleague, 
William  F.  Trimble,  who  is  one  of  this  nation's  preeminent  historians  of 
naval  aviation  (and  who  stays  abreast  of  the  history  of  space  exploration), 
offered  a  valuable  critique  of  chapter  8  on  the  genesis  of  the  lunar-orbit 
rendezvous  concept.  Major  Roy  F.  Houchin  (USAF),  one  of  my  doctoral 
students  at  Auburn,  read  a  few  of  the  chapters  and  offered  some  critical 
insights.  Others  in  my  department  whom  I  have  bothered  regularly  with 
reports  on  my  work  include  Guy  Beckwith,  Lindy  Biggs,  Anthony  Carey, 
J.  Wayne  Flynt,  Larry  Gerber,  W.  David  Lewis,  and  Steve  McFarland.  I 
thank  them  for  being  splendid  colleagues  and  good  listeners. 

Two  people  at  Auburn  University  that  I  wish  to  thank  for  finding 
the  means  and  the  tolerance  to  support  me  in  the  carrying  out  of  my 
research  projects  are  Gordon  Bond,  Dean  of  the  College  of  Liberal  Arts,  and 
Paul  F.  Parks,  University  Provost.  Before  becoming  my  Dean,  Gordon  Bond 
was  my  department  head  in  history,  a  job  whose  difficulties  I  appreciate 
now  more  than  ever,  since  taking  on  departmental  administration  myself. 
Also,  without  the  assistance  of  an  unbelievably  hardworking  and  talented 
administrative  assistant,  Jane  Dunkelberger,  I  am  afraid  the  job  of  the 
department  chairman  might  have  eaten  me  alive.  Jane  did  an  especially 
good  job  keeping  people  away  from  me  in  the  hectic  weeks  when  I  just  had 
to  work  on  this  book  to  meet  its  deadlines. 


xxi 


Space/light  Revolution 

Other  scholars  outside  of  Auburn  University  also  offered  critical  eval- 
uations of  all  or  part  of  my  manuscript.  In  particular,  I  wish  to  thank 
Virginia  P.  Dawson  of  Case  Western  Reserve  University  and  Michael  Corn, 
former  chief  historian  of  the  Air  Force  Systems  Command  and  current  histo- 
rian of  the  U.S.  Environmental  Protection  Agency,  for  providing  very  careful 
and  constructive  reviews  of  the  entire  manuscript.  Also,  Richard  K.  Smith, 
one  of  the  venerable  sages  in  the  study  of  American  aviation  history,  gave 
the  first  three  chapters  a  stern  critical  reading. 

Finally,  I  have  been  fortunate  beyond  any  reasonable  expectations  to  have 
had  the  enthusiastic  support  of  Roger  Launius,  chief  historian  for  NASA. 
Roger  allowed  this  book  project  a  high  degree  of  independence.  Apparently, 
he  trusted  that  I  could  produce,  and  he  had  faith  that  the  people  at  NASA 
Langley  had  the  ability  and  judgment  to  take  my  book  from  start  to  finish 
without  too  much  management  from  Washington.  I  hope  the  result  is  a 
book  that  he  will  be  proud  to  say  was  published  in  the  prestigious  NASA 
History  Series. 

Finally,  I  thank  you,  the  reader,  for  picking  up  such  a  big  book  and  giving 
it  more  than  a  passing  glance.  For  you,  I  have  given  it  my  best. 


December  1994  James  R.  Hansen 

Auburn,  Alabama 


xxu 


In  science  as  in  life,  it  is  well  known  that  a  chain  of 
events  can  have  a  point  of  crisis  that  could  magnify 
small  changes. 

— James  Gleick, 

Chaos:  The  Birth  of  a  New  Science 


Times  go  by  turns,   and  chances  change  by  course, 
From  foul  to  fair,  from  better  hap  to  worse. 

—Robert  Southwell 
"Times  Go  By  Turns" 


xxm 


Prologue 


Historians  should  start  from  the  premise  that  what  happened  did  not 
have  to  happen.  They  can  then  do  a  better  job  of  explaining  why  it  did. 

Too  often  we  think  about  history  as  something  that  had  to  happen 
just  the  way  that  it  did.  We  think  about  the  past  as  inevitable  and 
predetermined.  For  example,  we  think  about  the  American  Civil  War  as 
an  irreconcilable  conflict  that  had  to  occur  given  the  depth  of  the  regional 
differences  between  the  North  and  the  South  or  as  a  war  that  the  North, 
given  its  greater  population  and  industrial  might,  was  bound  to  win — when, 
perhaps,  neither  necessarily  had  to  be  the  case.  The  war  might  have  been 
avoided,  or  the  Southern  states  might  have  won  their  independence,  if  certain 
things  about  the  flow  of  history  had  been  different,  perhaps  only  slightly 
different. 

In  1991  a  controversy  developed  concerning  the  death  of  the  twelfth 
president  of  the  United  States,  Zachary  Taylor,  who  died  in  1850  from  a 
mysterious  intestinal  ailment,  conceivably  a  type  of  cholera.  Given  the 
symptoms  of  his  illness,  some  believed  that  Taylor  might  in  fact  have 
died  from  arsenic  poisoning;  maybe  a  Southerner,  angry  at  Taylor  for  his 
opposition  to  the  expansion  of  slavery,  found  a  way  to  murder  him.  Based 
on  this  theory,  in  1991  a  coroner  and  a  forensic  anthropologist  obtained  legal 
approval  to  exhume  Taylor's  body  from  his  tomb  in  Louisville,  Kentucky, 
and  conducted  an  autopsy  to  try  to  find  traces  of  arsenic  in  bits  of  hair, 
fingernail,  bone,  and  tissue.  As  it  turned  out,  they  found  nothing  to 
substantiate  the  theory  that  Taylor  was  murdered. 

While  this  investigation  was  going  on,  columnist  George  Will  wrote  a 
thoughtful  essay  about  the  whole  affair,  in  which  he  suggested  that  the 
country  might  have  followed  a  different  path  if  Zachary  Taylor  had  lived: 
the  Civil  War  might  have  been  avoided.1  Even  more  likely,  had  he  lived, 
Taylor  might  have  provoked  the  secessionist  movement  and  brought  on  the 
bloodshed  10  years  sooner.  The  South  would  have  faced  a  North  deprived  of 
a  decade's  worth  of  growth  in  industrialism  and  immigration  and  would  not 
have  confronted  a  new  political  party,  which  found  a  nation-saving  leader  in 
a  former  Illinois  congressman  named  Lincoln.  This  Civil  War  of  the  1850s 
the  South  might  have  won. 

A  more  fanciful  variation  on  this  what-if  theme,  again  involving  the 
Civil  War,  can  be  found  in  Ward  Moore's  classic  novella  of  1955,  Bring 
the  Jubilee.2  One  of  the  great  stories  of  time  travel,  this  fascinating  little 
book  is  based  on  the  idea  that  the  South  won  the  Civil  War  because  of  a 
single  turn  of  events  at  the  Battle  of  Gettysburg.  Moore's  story  is  rooted  in 


xxv 


Space/light  Revolution 

a  historical  event  in  which  a  Confederate  patrol  fails  to  arrive  at  a  certain 
place  at  a  given  time,  a  failure  that  enabled  the  Northern  forces  to  occupy 
a  strategic  place  on  the  battlefield  atop  Little  Roundtop.  In  Moore's  book, 
however,  the  Confederate  patrol  does  secure  this  strategic  position,  and 
the  South  goes  on  to  win  the  war.  Moore  draws  a  stunning  counterfactual 
portrait  of  post- Civil  War  America.  The  reader  encounters  a  prosperous 
and  progressive  South,  which  has  all  the  great  universities,  and  a  backward 
and  poverty-stricken  North. 

I  have  taken  the  time  to  mention  Ward  Moore's  fantasy  and  the  specula- 
tion surrounding  Zachary  Taylor's  death  simply  to  introduce  the  underlying 
theme  of  the  epic  story  of  space  exploration  that  follows:  the  past  was  no 
more  inevitable  than  is  our  future.  Contrary  to  what  we  might  have  been 
taught  in  school,  or  to  what  we  might  in  fact  still  be  teaching,  history  is  not 
a  straight  highway.  To  study  history  is  not  simply  to  take  a  pencil  and  play 
dot-to-dot.  Rather,  it  is  to  thread  a  maze,  to  follow  a  course  of  what  are 
potentially  limitless  directions,  including  "all  sorts  of  twists  and  turns  and 
fresh  choices  of  route  confronting  each  new  generation."  As  George  Will 
pointed  out  in  his  column  on  Zachary  Taylor,  history — whether  it  is  the 
history  of  the  American  Civil  War  or  the  history  of  our  own  individual 
lives — is  "a  rich  weave  of  many  threads."  Any  one  of  these  threads,  if  pulled 
out,  could  cause  a  radical  unraveling,  "setting  the  past  in  motion  as  a  foam- 
ing sea  of  exhilarating  contingencies."  In  other  words,  history  could  have 
been  different:  "Choices  and  chance  cannot  be  scrubbed  from  the  human 
story.  The  river  of  history  could  have  cut  a  different  canyon." 3  That  is  the 
theme  I  wish  to  explore  in  relation  to  the  history  of  one  of  the  premier  in- 
stitutions in  the  American  space  program,  NASA  Langley  Research  Center 
in  Hampton,  Virginia. 

In  the  keynote  address  of  a  conference  on  the  history  of  space  exploration 
held  at  Yale  University  in  1981,  New  York  Times  reporter  and  prominent 
American  space  journalist  John  Noble  Wilford  asked  a  provocative  what- 
if  question:  what  if  the  United  States  had  launched  the  first  satellite  in 
1957  instead  of  the  Soviets?  The  United  States  could  have  done  it.  We 
had  German  scientists  and  engineers  who  had  more  technical  expertise  than 
those  "recruited"  by  the  Soviets.  As  Wilford  explains,  "Wernher  von  Braun 
had  the  rocket  [a  modified  Redstone  designated  the  Jupiter  C]  and  could 
have  done  it  about  a  year  before  Sputnik,  but  was  under  orders  from  the 
Eisenhower  administration  not  to — the  first  American  satellite  was  supposed 
to  be  a  civilian  operation,  and  von  Braun  was  working  for  the  army  at 
the  time."4  To  guarantee  that  the  president's  orders  were  followed,  army 
inspectors  kept  a  careful  watch  on  the  prelaunch  activities  of  von  Braun  and 
his  men  at  Cape  Canaveral;  they  suspected  that  the  Alabama-based  rocket 
team  might  just  "accidentally"  launch  a  satellite  using  what  was  supposed 
to  be  a  dummy  upper  stage  of  the  Jupiter  C  to  boost  a  nose  cone  into  orbit.5 

In  terms  of  technical  capability  alone,  the  United  States  could  have 
beaten  the  Russians  into  space  with  a  satellite.  Explaining  why  our  country 


xxvi 


Prologue 

did  not  and  why  the  Eisenhower  administration  did  not  have  the  ambition 
to  do  so  is  difficult  without  reconstructing  some  complex  histories.  As 
Walter  A.  McDougall  argues  in  his  Pulitzer  Prize- winning  book  of  1985,  The 
Heavens  and  the  Earth:  A  Political  History  of  the  Space  Age,  the  explanation 
hinges  on  Eisenhower's  philosophy  of  government,  especially  his  fear  of  the 
growing  influence  of  what  he  would  come  to  call  "the  military-industrial 
complex."  More  specifically,  it  involves  his  administration's  recognition  of 
the  need  for  satellite  reconnaissance  of  the  closed  and  secretive  Communist 
world,  but  at  the  same  time,  the  administration's  concern  that  a  hot  (and 
expensive)  new  battle  in  the  cold  war  would  erupt  if  an  American  satellite 
with  military  associations  flew  over  the  airspace  of  the  Soviet  Union.  To 
avoid  such  an  eruption,  Eisenhower's  political  strategists  suggested  that  it 
would  be  best  to  let  the  Soviets  set  the  legal  precedent  by  orbiting  the  first 
satellite;  then,  when  an  American  satellite  followed,  the  Soviets  would  not 
have  solid  grounds  for  protesting  any  American  overflight.6 

With  these  issues  and  others  in  mind,  President  Eisenhower  made  his 
fateful  decision  to  support  the  more  peaceful-appearing  but  technically 
inferior  Vanguard  satellite  project  rather  than  the  project  involving  the 
Army  Ballistic  Missile  Agency's  (ABM A)  Jupiter  rocket.  Jupiter,  of  course, 
would  ultimately  boost  the  first  U.S.  satellite,  Explorer,  into  space  on  31 
January  1958,  nearly  two  months  after  the  Vanguard-carrying  Viking  rocket 
exploded  in  flames  on  the  launchpad  at  Cape  Canaveral  (the  press  dubbed 
it  "Flopnik,"  "Kaputnik,"  and  "Stayputnik" )  and  nearly  three  months  after 
the  Russians  successfully  orbited  their  canine-carrying  Sputnik  2.7 

If  Eisenhower  could  have  known  how  traumatic  and  revolutionary  the 
launching  of  the  first  satellite  would  prove  to  be  and  what  a  challenge  it 
would  pose  to  his  presidency  and  his  political  party,  he  might  have  decided 
differently.  The  von  Braun  team  might  have  been  turned  loose  sooner,  and 
the  beep-beep-beeping  that  radio  operators  heard  around  the  world  in  early 
October  1957  might  have  come  from  a  small  American  satellite  rather  than 
a  Russian  one. 

What  if  the  Americans  had  launched  a  satellite  first?  According  to 
Wilford,  "An  American  first  would  not  have  startled  the  world  as  much  as 
Sputnik  did,  for  American  technological  leadership  was  taken  for  granted. 
The  impact  of  Sputnik,  when  it  followed,  would  have  been  much  less, 
another  case  of  the  Russians  catching  up,  as  with  the  atomic  and  hydrogen 
bombs."8  And  if  that  had  been  the  case,  if  Americans  had  not  found 
Sputnik  so  challenging,  what  kind  of  space  program  would  U.S.  leaders  have 
formulated?  Surely,  that  program  would  have  differed  from  the  ideologically 
motivated  and  in  key  respects  shortsighted  one  that  was  mobilized  in  such 
a  hurry  to  win  the  space  race.  If  Sputnik  had  not  provoked  a  major 
international  crisis,  much  about  the  history  of  the  world  in  the  last  four 
decades  of  the  twentieth  century  would  have  been  significantly  different. 

Consider  America  without  a  Sputnik  crisis.  Without  the  snowball- 
ing political  repercussions  that  were  so  damaging  to  the  Republicans, 

xxvii 


Space/light  Revolution 

Richard  M.  Nixon,  Eisenhower's  vice-president,  possibly  would  have  de- 
feated Democratic  Senator  John  F.  Kennedy  of  Massachusetts  in  the 
whisker-close  1960  presidential  election.  A  reversal  in  that  election  alone, 
which  turned  on  a  few  thousand  questionable  votes  in  Illinois,  would  have 
produced  such  an  unraveling  of  contemporary  American  history  that  only  a 
Ward  Moore  could  do  it  justice.9 

The  character  of  the  country's  inaugural  ventures  into  space  would  have 
been  vastly  different.  Without  the  media  riot,  without  the  panic  incited 
by  cold  war  misapprehensions  about  the  Soviet  satellite,  without  the  feeling 
that  the  Russians  had  gotten  a  jump  on  us,  and  without  the  resulting  clamor 
for  our  government  to  do  something  dramatic  right  now  to  close  the  gap, 
the  National  Advisory  Committee  for  Aeronautics  (NAG A),  which  dated 
to  World  War  I  and  was  the  forerunner  of  the  National  Aeronautics  and 
Space  Administration  (NASA),  would  have  surely  lived  on.10  Most  likely 
this  agency  would  have  proceeded  calmly  with  plans  to  expand  its  space- 
related  research,  and  NASA  would  not  have  been  established,  at  least  not 
when  it  was.  The  United  States  would  still  have  entered  into  space,  but  the 
country  would  not  have  rushed  into  it. 

Instead  of  plunging  into  the  ocean  in  a  ballistic  capsule,  the  first 
American  astronauts  might  have  flown  back  from  space  on  the  wings  of 
a  hypersonic  glider  similar  to  those  NACA  researchers  had  been  working  on 
since  the  mid-1950s.  If  the  United  States  had  not  lacked  a  booster  rocket 
powerful  enough  to  lift  so  heavy  a  weight  out  of  the  atmosphere,  the  first 
spaceflight  might  have  happened  like  that  anyway,  even  with  the  Sputnik 
crisis.  The  original  seven  astronauts  (the  ones  with  "the  right  stuff")  or  more 
likely,  specially  trained  NACA  or  military  test  pilots  would  have  traveled 
to  space  and  back  in  a  laudable  space  plane  akin  to  a  small  space  shuttle. 
Given  the  time  needed  to  develop  the  requisite  booster  and  considering  the 
extensive  development  and  careful  flight  testing  that  such  a  radically  new, 
winged  reentry  vehicle  inevitably  would  have  undergone,  the  hypersonic 
glider  probably  would  not  have  been  launched  into  space  until  the  late  1960s, 
but  it  surely  would  have  proved  much  more  capable  and  versatile  than  the 
Mercury  capsules.11 

Moreover,  instead  of  sending  men  to  the  moon  by  the  end  of  the  decade 
as  President  Kennedy  had  wanted,  an  NACA-led  program  under  President 
Nixon  likely  would  have  focused  on  the  construction  of  a  small,  staffed 
space  station  that  could  have  been  serviced  by  the  shuttle-like  vehicle. 
Such  was  the  target  project  for  space  exploration  at  the  NACA  research 
laboratories  before  Sputnik,  and  it  remained  so  until  President  Kennedy's 
lunar  commitment  in  May  1961. 12 

Whatever  we  think  about  the  might-have-beens  and  paths-not-taken,  the 
undeniable  fact  is  that  Sputnik  changed  the  course  of  history.  Sputnik  was 
one  of  those  revolutionary,  megahistorical  events  that  interrupted  the  flow  of 
things,  altered  the  would-have-beens,  and  made  a  lot  of  very  unlikely  events 
happen.  No  one  has  expressed  the  irony  of  the  randomness  and  illogic  in  the 


XXVlll 


Prologue 

historical  process  better  than  the  longshoreman-philosopher  and  quasi-cult 
figure  of  the  1950s  and  1960s,  Eric  Hoffer.  "What  were  the  terrible  1960s 
and  where  did  they  come  from?"  asked  Hoffer  after  the  end  of  the  decade. 
"To  begin  with,  the  1960s  did  not  start  in  1960.  They  started  in  1957.  .  .  . 
The  Russians  placed  a  medicine-ball-sized  satellite  in  orbit.  .  .  .  We  reacted 
hysterically."1  If  we  had  not,  or  if  we  had  put  that  "ball"  in  orbit  first, 
everything  would  have  been  different.  For  the  past  was  no  more  inevitable 
than  is  our  future. 

After  Sputnik,  the  American  space  program  would  contend  with  other 
critical  turning  points  and  other  what-ifs:  What  if  President  Kennedy  had 
not  committed  the  country  to  the  manned  lunar  landing — or  at  least  not  to 
accomplishing  it  so  quickly?  What  if  NASA  had  not  chosen  lunar-orbit 
rendezvous  as  the  mission  mode  for  Apollo  and  had  instead  gone  with 
direct  ascent  or  earth-orbit  rendezvous,  as  most  engineers  at  NASA  Marshall 
Space  Center  had  wanted?  What  if  the  national  supersonic  transport  (SST) 
program  had  not  been  cancelled  by  Congress  in  1971?  (The  U.S.  Senate 
killed  the  program  by  only  one  vote.)  Would  the  United  States  be  flying 
a  competitor  to  the  Concorde?  Would  the  resulting  airplane  have  been  a 
disastrous  failure,  thus  putting  Boeing  and  most  of  its  customers  out  of 
business?  What  if  the  Nixon  administration  in  1972  had  not  decided  to  go 
ahead  with  a  scaled-back  version  of  the  space  shuttle  but  instead  had  wanted 
to  develop  a  space  station?  What  if  President  Reagan  had  not  endorsed  the 
space  station  in  1984?  What  if  the  temperature  at  Cape  Canaveral  on  the 
morning  of  28  January  1986  had  been  only  a  few  degrees  warmer?  These 
are  just  some  of  the  what-if  questions  we  might  ask  about  NASA  and  the 
American  space  program.14 

The  study  of  history,  at  least  the  history  of  NASA,  reveals  something 
about  the  past  that  should  not  be  surprising,  but  is:  historical  development 
is  neither  linear  nor  logical.  In  practice,  talking  about  the  next  logical  step, 
something  that  NASA  planners  have  been  talking  about  nonstop  ever  since 
NASA  came  to  life,  does  not  ensure  that  step  will  be  the  next  one  taken. 
After  launching  a  man  into  space  via  Project  Mercury,  NASA  said  that  the 
next  logical  step  was  to  establish  a  permanent  manned  presence  in  low  earth 
orbit,  but  instead  the  country  landed  men  on  the  moon.  After  going  to  the 
moon  via  Project  Apollo,  the  next  logical  step  was  to  build  an  earth-orbiting 
space  station  along  with  a  space  shuttle  to  service  it,  but  instead  the  Nixon 
administration  decided  that  the  country  could  not  afford  both  and  could 
manage  temporarily  with  just  the  shuttle,  even  though  the  space  station 
had  always  been  the  shuttle's  main  reason  for  existing.  After  the  shuttle, 
surely  the  next  logical  step  was  to  build  a  space  station,  but  once  again  the 
country  has  found  reasons  to  postpone  building  one. 

Clearly,  logic  does  not  determine  our  history.  Historical  logic,  if  we  even 
want  to  use  that  phrase,  is  not  the  logic  of  scientists  and  mathematicians; 
it  is  the  logic  of  Through  the  Looking- Glass.  In  that  all-too-real  fantasy 
land,  Tweedledee  explains  logic  to  Alice:  "Contrariwise,  if  it  was  so,  it 


xxix 


Spaceflight  Revolution 

might  be;  and  if  it  were  so,  it  would  be;  but  as  it  isn't,  it  ain't.  That's 
logic."15  Tweedledee's  logic  is  the  only  kind  the  American  space  program 
has  ever  known,  or  probably  ever  will. 

In  this  book,  I  explore  the  impact  of  that  logic  on  the  research  and  de- 
velopment activities  conducted  at  Langley  Research  Center  in  the  12  years 
after  Sputnik.  As  the  book's  title  suggests,  this  impact  was  revolutionary.  I 
gave  much  thought  to  the  word  revolutionary  before  using  it.  In  the  history 
of  science,  since  the  publication  of  Thomas  S.  Kuhn's  seminal  study  The 
Structure  of  Scientific  Revolutions  in  1962,  no  historian,  in  fact  no  scholar, 
has  been  safe  in  the  use  of  the  term  revolution  without  reference  to  the  essen- 
tial Kuhnian  concepts  and  terminology:  "paradigm,"  "anomaly,"  "normal 
science,"  "Gestalt  switch,"  "paradigm  shift,"  and  the  "incommensurability 
of  paradigms,"  to  name  just  a  few.16  All  these  terms,  along  with  the  word 
revolution  itself,  which  Kuhn  defines  as  "those  noncumulative  developmen- 
tal episodes  in  which  an  older  paradigm  is  replaced  in  whole  or  in  part  by 
an  incompatible  one,"  have  thus  been  loaded  down  with  meaning,  nuance, 
argument,  controversy,  and  their  own  long  academic  histories.17 

But  the  reader  can  relax.  Nowhere  else  in  the  text  or  notes  of  this 
book  will  I  make  direct  reference  to  Thomas  Kuhn  or  his  sociological 
anatomy  of  revolution.  I  do  not  omit  Kuhn  because  of  any  disdain  for  his 
insights;  I  just  do  not  feel  that  any  explicit  application  of  Kuhn's  analysis 
of  scientific  revolutions  will  do  much  to  inform  my  chosen  topic  relevant 
to  NASA  Langley  history.  Whether  Kuhn's  notions  have  worked  implicitly 
to  influence  my  understanding  of  the  spaceflight  revolution  at  the  research 
center,  I  leave  to  the  reader  to  judge.18 

Most  scholars  are  familiar  with  Kuhn  and  his  concept  of  revolution;  far 
fewer  are  familiar  with  the  particular  concept  of  the  spaceflight  revolution  for 
which  Kuhnian  sociologist  William  Sims  Bainbridge  is  responsible.  Despite 
my  using  Bainbridge's  terminology  and  even  sympathizing  with  parts  of  his 
concept,  I  wish  to  distance  myself  and  this  book  on  NASA  Langley  from  it, 
even  farther  than  I  have  from  Kuhn. 

In  1976  Bainbridge,  a  professor  in  the  sociology  department  at  the 
University  of  Washington,  published  a  fascinating  if  eccentric  analysis  of 
the  enthusiasms  of  the  space  age,  The  Spaceflight  Revolution:  A  Sociological 
Analysis.*  According  to  its  thesis,  the  space  age  came  to  life  "despite  the 
world's  indifference  and  without  compelling  economic,  military,  or  scientific 


Even  Bainbridge  worried  that  the  word  revolution  might  be  too  strong.  In  the  introduc- 
tion to  his  book  he  defends  its  use,  saying  that  "the  scale  and  the  manner  of  the  achieve- 
ment" in  space  "demand  powerful  language."  According  to  his  estimates,  "approximately 
$100,000,000,000  has  been  spent  on  space  technology;  the  exact  figure  is  debatable,  but  the 
order  of  magnitude  is  not."  Moreover,  Bainbridge  continued,  "I  use  the  word  revolution  as 
a  scientifically  descriptive  term  [as  Kuhn  did],  not  a  metaphor.  The  development  of  space- 
flight  could  be  a  revolution  in  two  ways:  its  consequences  and  its  causes."  ( The  Spaceflight 
Revolution:  A  Sociological  Analysis  [New  York:  John  Wiley  &  Sons,  1976],  p.  1.) 

XXX 


Prologue 

reasons  for  its  accomplishment."  It  was  not  the  "public  will,"  declared 
Bainbridge,  but  "private  fanaticism"  that  drove  us  to  the  moon.  "When 
Neil  Armstrong  called  his  'small  step'  down  on  to  the  lunar  surface  a  'great 
leap  for  mankind',  he  spoke  as  the  partisan  member  of  a  revolutionary  social 
movement,  eager  to  convert  the  unbelieving  to  his  faith."19 

Bainbridge's  book  essentially  advances  a  conspiracy  theory.  The  majority 
of  people  did  not  want  spaceflight;  only  a  few  did.  And  those  few  romantic 
idealists,  that  extremely  small  but  dedicated  and  well-organized  network 
of  men  (very  few  women  were  at  first  involved,  according  to  Bainbridge), 
coaxed,  tricked,  lobbied,  and  coerced  the  greatest  technological  nations 
into  building  mammoth  programs  to  launch  them  into  space.  Bainbridge 
then  analyzes  the  historical  and  social  character  of  the  conspirators:  the 
pioneers  and  visionaries  of  spaceflight  (the  Russian  Konstantin  Tsiolkovskii, 
the  German  Hermann  Oberth,  and  the  American  Robert  Goddard,  among 
others);  the  enthusiastic  members  of  the  early  space  and  rocket  clubs  (such 
as  the  German  Society  for  Space  Travel,  the  British  Interplanetary  Society, 
and  the  American  Interplanetary  Society);  Wernher  von  Braun's  rocket  team 
in  league  with  the  Nazis  at  Peenemiinde;  the  agenda  of  the  Committee  for 
the  Future,  that  "mystical,  almost  religious  organization,"  which  came  to 
life  in  the  United  States  in  1970,  less  than  one  year  after  the  first  manned 
lunar  landing;  and  finally,  the  science-fiction  subculture,  which  he  calls  the 
"breeding  ground  of  deviant  movements,"  and  the  Star  Trek  and  Search  for 
Extraterrestrial  Intelligence  (SETI)  groupies  of  the  present  day.20 

The  book  is  a  brilliant  and  troubling  tour  de  force  from  a  sociologist  of 
some  estimable  abilities.  I  assign  it  perennially  to  my  graduate  students 
in  aerospace  history  and  not  just  to  get  a  rise  from  them,  which  it  always 
does — particularly  from  the  students  specializing  in  military  air  power  who 
usually  think  that  Bainbridge  is  simply  silly  or  crazy.  Bainbridge's  version  of 
the  spaceflight  revolution  is  worth  investigating,  if  only  because  it  explores 
the  question  of  why  something  that  did  not  have  to  happen,  happened.  In 
the  introduction  to  his  book,  Bainbridge  writes,  as  I  have  written  in  this 
prologue,  that  the  spaceflight  revolution  "was  a  revolution  that  need  not 
have  happened."21 

In  my  version  of  the  spaceflight  revolution,  however,  the  revolutionaries 
are  not  conspirators  from  rocket  enthusiast  organizations  and  science-fiction 
clubs,  nor  are  they  romantic  idealists  aspiring  to  some  quasi-religious, 
superhuman,  or  millenarian  experience  in  outer  space.  And  they  are  hardly 
members  of  a  deviant  social  movement.  Rather,  my  revolutionaries  are 
government  engineers  and  bureaucrats,  who  are  members  of  an  established 
research  organization  dating  back  to  1915,  the  venerable  NACA.  These 
revolutionaries,  because  of  the  hysteria  over  the  launch  of  Sputnik  1  in 
October  1957,  metamorphosed  along  with  their  organization  into  creatures 
of  the  space  age. 

My  spaceflight  revolution  is  an  unlikely  story — perhaps  as  unlikely  as 
Bainbridge's.  But  this  one  happened. 

xxxi 


The  Metamorphosis 


It  was  the  worm,  if  you  will,  going  into  the  cocoon 
and  coming  out  a  butterfly. 

-Walter  Bonney,  NACA/NASA 
public  relations  officer 


The  first  week  of  October  1958  was  a  busy  time  for  the  newspapers 
of  Tidewater  Virginia.  Top  stories  included  the  explosive  failure  of  an 
Atlas  missile  at  Cape  Canaveral,  an  atomic  blast  in  Nevada  that  sent  news 
and  test  personnel  scurrying  for  cover  from  radiation  fallout,  the  question 
of  Red  China's  membership  in  the  United  Nations,  and  a  United  Auto 
Workers  strike  against  the  Ford  Motor  Company.  Receiving  the  biggest 
headlines  in  the  local  papers,  however,  were  stories  concerning  the  path  of 
Hurricane  Helene  up  the  Atlantic  coast  and  the  furor  over  the  court-ordered 
integration  of  public  schools,  which  was  taking  place  as  far  away  as  Little 
Rock,  Arkansas,  and  as  nearby  as  Richmond  and  Norfolk.  Not  even  making 
the  front  page  of  the  Newport  News  Daily  Press  on  the  cool,  overcast  morning 
of  Wednesday,  1  October  1958,  was  the  news  that  the  National  Advisory 
Committee  for  Aeronautics  (NACA)  had  died  the  night  before  at  midnight, 
only  to  be  reborn  at  12:01  a.m.  as  the  National  Aeronautics  and  Space 
Administration.  Just  a  few  hours  earlier,  on  Tuesday,  7000  people  had  left 
work  as  NACA  employees,  but  when  they  reported  to  their  same  jobs  in  the 
same  buildings  the  next  morning,  they  became  members  of  NASA.* 

A  few  NACA  veterans  might  have  felt  a  twinge  of  doubt  as  they  drove 
past  the  new  NASA  sign  at  the  gates  of  Langley  Research  Center,  but  most 
NACA  personnel  were  not  at  all  nervous  or  wary  about  the  changeover. 
Plans  for  an  easy  transition  had  been  in  the  works  for  at  least  eight  months, 


Although  foreigners  tended  to  pronounce  it  as  a  two-syllable  word,  "Nacka,"  within  the  United 
States  the  organization  was  always  known  by  its  four  individual  letters,  "the  N-A-C-A."  Veterans  of  the 
NACA  assumed  that  the  same  would  be  true  for  NASA.  Into  the  1990s,  NACA  veterans  could  usually 
be  identified  by  the  way  they  treated  the  NASA  acronym  as  individual  letters. 


Spaceflight  Revolution 

since  President  Dwight  D.  Eisenhower's  panel  of  scientific  advisers  had 
recommended  that  a  new  civilian  space  agency  be  organized  around  the 
NACA.1  Almost  everything  about  working  at  Langley  Field,  or  at  any 
of  the  other  former  NACA  facilities  around  the  country,  was  supposed  to 
remain  the  same.  Employees  had  been  reassured  for  several  weeks  by  NACA 
headquarters  and  by  Langley  management  that  they  were  to  come  to  work 
as  always  and  do  the  same  things  they  had  been  doing.  Their  jobs  already 
had  much  to  do  with  the  nation's  quickly  accelerating  efforts  to  catch  up 
with  the  Soviet  Union  and  launch  America  into  space.  As  NASA  personnel, 
they  were  simply  to  keep  up  the  good  work. 

After  watching  from  a  distance  the  hysteria  provoked  by  the  Soviet 
satellites  and  the  political  jousting  and  bureaucratic  haggling  that  followed, 
Langley  employees  were  relieved  to  see  President  Eisenhower  resist  the 
pressures  applied  by  the  military,  particularly  the  air  force,  to  militarize  the 
infant  American  space  program.2  Ike,  the  former  five-star  army  general  and 
leader  of  the  invasion  of  Nazi-occupied  Europe  in  1944,  had  risen  above  these 
pressures  and  put  civilians  in  charge,  entrusting  the  NACA  with  the  space 
program.  A  small  overhead  agency  that  was  both  focused  and  accustomed 
to  squeezing  a  dollar,  the  NACA  appealed  to  a  genuine  balanced-budget 
man  like  Eisenhower. 

The  creation  of  the  NACA  had  been  quite  different  from  that  of  NASA. 
Although  a  group  of  prominent  Smithsonian  and  Washington  aviation 
enthusiasts  had  conceived  the  idea  of  an  organization  devoted  to  the  support 
of  aeronautical  development  as  early  as  1910,  the  actual  founding  of  this 
new  federal  agency  proved  difficult,  especially  since  aviation  had  not  yet 
demonstrated  its  efficacy  in  World  War  I  combat.  In  fact,  establishment  of 
the  NACA  might  not  have  been  approved  if  a  friendly  group  of  congressmen, 
fearing  that  President  Woodrow  Wilson's  policy  of  neutrality  was  preventing 
the  United  States  from  properly  preparing  for  its  inevitable  role  in  the  war, 
had  not  devised  a  successful  last-minute  maneuver.  In  a  classic  example  of 
American  political  sleight-of-hand,  they  attached  the  NACA  enabling  act  as 
a  rider  to  a  naval  appropriations  bill  that  was  sure  to  pass,  and  the  NACA 
came  into  being  on  3  March  1915.3 

For  an  important  new  government  body  to  be  established  in  such  a 
manner  was  really  quite  extraordinary.  But  certainly  no  one  in  1915  or  for 
several  years  thereafter,  perhaps  not  even  many  early  NACA  employees, 
considered  the  NACA  very  important.  Now,  43  years  later,  President 
Eisenhower  was  making  it  the  heart  of  the  new  American  space  program 
for  which  everyone  was  clamoring.  Because  of  the  heated  public  debate 
over  national  space  policy,  NASA  could  not  have  been  founded  in  the 
relatively  invisible  way  that  the  NACA  had  been  established.  Unlike  the 
old  agency,  NASA  was  going  to  be  exposed  to  direct  congressional,  media, 
and,  consequently,  public  scrutiny  from  the  start. 

Probably  no  NACA  employees  arriving  at  work  on  .NASA's  first  day 
anticipated  the  impact  that  this  new  life  in  a  goldfish  bowl  eventually  would 


The  Metamorphosis 


Although  his  administration  gave 
birth  to  NASA,  President  Dwight 
D.  Eisenhower  did  not  believe  that 
the  United  States  should  rush  into 
a  "crash"  federal  program  to  beat 
the  Soviets  into  space.  Instead, 
he  hoped  for  a  more  judicious 
and  less  hysterical  approach  to 
space  exploration,  one  that  would 
not  require  massive  infusions  of 
public  funds  but  would  still  en- 
able the  United  States  to  remain 
a  leader,  if  not  the  leader,  in 
space.  His  Democratic  successors, 
John  F.  Kennedy  and  Lyndon  B. 
Johnson,  would  commit  the  coun- 
try to  an  all-out  race. 


L-58-68a 


have  on  their  work  and  workplace.  Change  is  difficult  to  perceive  and 
evaluate  while  it  is  happening,  let  alone  when  it  occurs  in  the  middle  of 
a  week.  Charles  J.  Donlan,  veteran  Langley  researcher  and  soon-to-be- 
named  associate  director  of  NASA's  Space  Task  Group,  later  reminisced 
about  the  innocence  of  his  thoughts  on  the  day  the  NACA  became  NASA: 
"It  was  like  passing  from  December  31  to  January  1  without  going  to  a 
party.  You  didn't  know  the  difference  except  that  it  was  the  New  Year 
and  you  had  to  start  signing  your  checks  for  one  year  later. "^  Indeed,  a 
new  era  had  begun,  and  although  this  was  not  apparent  on  the  uneventful 
morning  of  1  October  1958,  Langley  Research  Center  was  now  exposed  to 
the  complex  forces  and  extreme  circumstances  that  were  rapidly  reshaping 
U.S.  aeronautical  research  and  blasting  the  center  pell-mell  into  space. 

The  Venerable  Order  of  the  NACA 


The  basic  duty  of  the  NACA,   as  expressed  in  its  charter,   was   "to 
supervise  and  direct  the  scientific  study  of  the  problems  of  flight,  with 


Space/light  Revolution 

a  view  to  their  practical  solution,  and  to  determine  the  problems  which 
should  be  experimentally  attacked,  and  to  discuss  their  solution  and  their 
application  to  practical  questions."  But  the  original  charter  of  1915  did 
not  assure  the  funds  for  the  large,  diversified,  and  increasingly  expensive 
research  establishment  that  the  NACA  eventually  became.  It  stated  only 
that  "in  the  event  of  a  laboratory  or  laboratories,  either  in  whole  or  in  part, 
being  placed  under  the  direction  of  the  committee,  the  committee  may  direct 
and  conduct  research  and  experiment  in  aeronautics."5 

That  mandate  was  general  enough  to  allow  widely  differing  interpreta- 
tions, and  not  everyone  responsible  for  the  NACA  in  its  formative  years 
agreed  on  what  the  mandate  meant  or,  rather,  what  it  should  mean.  Some 
felt  that  the  NACA  should  remain  small  and  continue  to  serve,  as  it  had 
throughout  World  War  I,  merely  as  an  advisory  body  devoted  to  scientific 
research.  Others  argued  that  the  NACA  should  grow  larger  and  combine 
basic  research  with  engineering  and  technology  development.  This  second 
group  wanted  the  NACA  to  attack  the  most  pressing  problems  obstruct- 
ing the  immediate  progress  of  American  aviation;  the  group  did  not  want 
the  agency  to  spend  all  of  its  time  on  ivory-tower  theoretical  problems  that 
would  not  result  in  many  quick,  practical  payoffs.  To  be  so  effective,  the 
NACA  needed  to  have  its  own  laboratory  facilities  and  conduct  its  own 
programs  of  research. 

The  NACA  moved  slowly  but  surely  along  the  second  course,  and 
building  a  laboratory  became  its  first  order  of  business.  Construction 
of  the  Langley  Memorial  Aeronautical  Laboratory,  the  NACA's  original 
field  station,  began  approximately  100  miles  southeast  of  Washington, 
on  an  isolated  peninsula  of  Tidewater  Virginia  in  1917.  Named  after 
Dr.  Samuel  P.  Langley  (1835-1906),  an  eminent  American  scientist  whose 
pioneering  experiments  with  powered  flight  at  the  turn  of  the  century  had 
been  a  mixture  of  success  and  failure,  Langley  served  as  the  NACA's  only 
research  center  for  the  next  20  years.6  Some  flight  research  was  conducted 
there  in  late  1919  and  early  1920,  but  the  laboratory  did  not  really  begin 
routine  operations  until  after  the  completion  of  its  first  wind  tunnel  in  the 
summer  of  1920. 

By  the  mid- 1920s,  engineers,  not  scientists,  were  put  in  charge  at  Langley. 
The  head  of  the  laboratory  would  in  fact  be  called  the  "engineer  in  charge." 
The  choice  of  engineers  over  scientists  reinforced  the  NACA's  decision  to 
become  an  agency  concerned  with  the  practical,  not  the  purely  theoretical. 
Engineers  would  always  support  the  NACA's  charter.  On  Langley  engineer 
Floyd  L.  Thompson's  desk  sat  a  framed  quotation  of  the  essence  of  the 
charter:  "The  scientific  study  of  the  problems  of  flight  with  a  view  to  their 
practical  solution."  The  quote  stayed  on  Thompson's  desk  until  he  retired 
from  NASA  as  the  director  of  Langley  Research  Center  in  1968. 

In  the  years  following  its  founding,  the  NACA  expanded  far  beyond  the 
advisory  role  defined  in  its  charter.  The  NACA  served  as  a  national  clear- 
inghouse for  scientific  and  technical  information  by  establishing  uniform 


The  Metamorphosis 


L-36,942 


Langley  map  of  the  Tidewater  Virginia  area  from  the  late  1930s. 


Spaceflight  Revolution 


L-62-6373 

Floyd  L.  "Tommy"  Thompson  was  Langley  Research  Center's  associate  direc- 
tor in  1958,  its  "number  two"  man  under  longtime  and  soon-to-retire  Director 
Henry  J.  E.  Reid.  The  number  two  man  in  those  days  also  acted  as  the  chief 
of  research. 


L-9850 


Orville  Wright,  Charles  Lindbergh,  and  Howard  Hughes  were  among  the  attendees 
at  Langley 's  1934  Aircraft  Manufacturers'  Conference.  Conference  guests  assembled 
underneath  a  Boeing  P-26A  Peashooter  in  the  Full-Scale  Tunnel  for  this  photo. 


The  Metamorphosis 

aeronautical  terminology;  publishing  reports;  and  collecting,  compiling,  and 
disseminating  basic  information  in  the  various  fields  pertinent  to  aeronau- 
tics. It  also  contracted  out  research  projects  to  universities.  From  1926  on, 
it  held  annual  meetings  known  as  the  NACA  Aircraft  Manufacturers'  Con- 
ferences, which  brought  in  experts  from  around  the  United  States  to  talk 
about  aviation  technology  and  what  the  NACA  should  be  doing  to  stimulate 
further  progress.7  It  built  up  staffs  to  conduct  research  in  aerodynamics, 
hydrodynamics,  structures,  and  propulsion.  Solutions  to  problems  in  these 
areas  led  to  the  design  and  operation  of  safer,  faster,  higher  flying,  and 
generally  more  versatile  and  dependable  aircraft.  With  these  aircraft,  the 
United  States  became  a  world  power  in  commercial  and  military  aviation, 
and  Allied  victory  in  World  War  II  was  assured. 

To  help  meet  the  demand  for  advanced  airplane  work  during  World  War 
II,  the  NACA  created  four  new  national  facilities  and  seeded  them  with 
staff  from  Langley.  They  were  the  Aircraft  Engine  Research  Laboratory, 
built  in  Cleveland,  Ohio,  in  1941  (later  renamed  the  Lewis  Flight  Propul- 
sion Laboratory  and  later  still  the  Lewis  Research  Center);  the  Ames  Aero- 
nautical Laboratory,  created  at  Moffett  Field,  California,  also  in  1941  (later 
renamed  Ames  Research  Center);  the  Pilotless  Aircraft  Research  Station, 
built  on  barren  Wallops  Island  on  Virginia's  Eastern  Shore  in  1944  (later 
renamed  Wallops  Station);  and  the  High-Speed  Flight  Station,  established 
at  Muroc  Field  (subsequently,  Edwards  Air  Force  Base  [AFB]),  California, 
in  1946  (later  renamed  Dryden  Flight  Research  Center).  At  the  last  facility 
in  the  high  California  desert,  a  special  unit  of  engineers  from  Langley  su- 
pervised the  flight  trials  of  the  first  supersonic  airplanes,  the  Bell  X-l  and 
the  Douglas  D-558.  Considering  the  many  technological  firsts  and  other 
achievements  arising  from  this  array  of  unique  facilities,  it  is  clear  why  many 
experts  believe  the  NACA  did  at  least  as  much  for  aeronautical  progress  as 
any  organization  in  the  world.8 

Indeed,  the  NACA's  track  record  was  not  bad  for  a  committee,  or 
rather,  for  a  pyramid  of  committees — the  NACA  consisted  of  more  than 
one.  Foremost  was  the  NACA's  Main  Committee,  an  unpaid  body  that  met 
twice  a  year  in  Washington  to  identify  and  discuss  the  key  research  prob- 
lems that  the  agency  should  tackle.  Until  World  Wax  II,  it  comprised  12 
members  and  from  then  on  15.  Members  represented  the  War  and  Navy 
departments  (normally  two  from  each),  the  Smithsonian  Institution,  the 
U.S.  Weather  Bureau,  and  the  National  Bureau  of  Standards,  as  well  as  se- 
lect universities,  industries,  and  airlines.  The  list  of  120  men  who  served  on 
the  NACA  Main  Committee  ("The  NACA")  from  1915  to  1958  is  a  "Who's 
Who"  of  American  aeronautics:  Dr.  Joseph  S.  Ames,  Gen.  Henry  "Hap" 
Arnold,  Dr.  Vannevar  Bush,  Harry  F.  Guggenheim,  Dr.  William  F.  Durand, 
Dr.  Jerome  C.  Hunsaker,  Charles  A.  Lindbergh,  Adm.  William  A.  Moffett, 
Capt.  Edward  V.  "Eddie"  Rickenbacker,  Gen.  Carl  "Tooey"  Spaatz,  Gen. 
Hoyt  Vandenberg,  and  Orville  Wright,  to  name  a  few.  The  president  of 
the  United  States  appointed  all  members,  and  in  turn  the  Main  Committee 


Space/light  Revolution 


L-90-3727 

An  18  April  1929  meeting  of  the  NACA  Main  Committee.  Around  this  ta- 
ble sat  some  of  the  most  outstanding  authorities  on  the  science,  technology,  and 
military  uses  of  flight.  Left  to  right:  John  F.  Victory,  NACA  secretary;  Dr. 
William  F.  Durand,  professor  and  head  of  the  Department  of  Mechanical  Engi- 
neering, Stanford  University;  Dr.  Orville  Wright;  Dr.  George  K.  Burgess,  di- 
rector, Bureau  of  Standards;  Brig.  Gen.  William  E.  Gilmore,  U.S.  Army;  Maj. 
Gen.  James  E.  Fechet,  Chief  of  Air  Service,  USA;  NACA  Chairman  Dr.  Joseph 
S.  Ames,  professor  of  physics  and  president  of  Johns  Hopkins  University;  NACA 
Vice- Chairman  Dr.  David  W.  Taylor,  former  Chief  Naval  Constructor,  U.S. 
Navy;  Capt.  Emory  S.  Land,  Navy  Bureau  of  Aeronautics;  Rear  Adm.  William 
A.  Moffett,  Chief,  Bureau  of  Aeronautics;  Dr.  Samuel  W.  Stratton,  former  di- 
rector, Bureau  of  Standards;  Dr.  George  W.  Lewis,  NACA  director  of  research; 
Dr.  Charles  F.  Marvin,  Chief,  Weather  Bureau. 


reported  directly  to  him  via  an  annual  written  report.  The  eighth  and  last 
chairman  of  the  Main  Committee  was  Dr.  James  H.  "Jimmy"  Doolittle,  the 
former  racing  pilot,  air  war  hero,  retired  air  force  general,  and  Ph.D.  in 
physics  from  Massachusetts  Institute  of  Technology  (MIT).  On  30  Septem- 
ber 1958,  the  day  before  NASA  took  over,  he  sent  the  NACA's  44th  and 
last  annual  report  to  President  Eisenhower. 

The  NACA  was  quite  independent.  Although  the  president  appointed  its 
members,  he  did  so  on  advice  from  the  standing  NACA  Main  Committee, 
advice  that  Eisenhower  and  his  predecessors  almost  always  took.  This 
helped  to  take  politics  out  of  the  selection  process.  Furthermore,  the  Main 
Committee  chose  its  own  chairman  and  director  of  research  and,  in  the  words 
of  longtime  NACA  member  (1922-1923,  1938-1958)  and  former  chairman 

8 


The  Metamorphosis 


L-16,243 

One  of  the  outstanding  men  to  chair  the  NACA  was  Vannevar  Bush  (center),  the 
computer  pioneer  and  head  of  the  Office  of  Scientific  Research  and  Development 
during  World  War  II.  Bush  chaired  the  NACA  from  1939  to  1941.  On  either  side 
of  Bush  stand  George  W.  Lewis,  the  NACA's  longtime  director  of  research  (right) 
and  Henry  Reid,  Langley's  engineer-in- charge  (left). 


(1941-1956)  Jerome  Hunsaker,  "ran  its  show,  within  its  budget,  made  its 
own  statements  to  Congress  for  what  it  wanted  to  do  and  could  do  and 
was  doing,  and  got  [its]  budgets  without  any  interference  from  the  executive 
branch  of  government."5 

Organizationally,  the  old  NACA  committee  system  did  not  stop  with 
the  Main  Committee.10  Its  members  elected  a  smaller  Executive  Commit- 
tee of  seven  who  served  terms  of  one  year  and  acted  as  the  NACA's  actual 
governing  body.  This  Executive  Committee  also  appointed  several  technical 
committees  that  provided  expertise  to  the  parent  committees  on  such  major 
subjects  as  aerodynamics,  power  plants  for  aircraft,  and  aircraft  construc- 
tion. In  turn,  these  committees  (actually  subcommittees)  created  sub(sub)- 
committees  of  their  own  to  study  and  give  advice  in  more  specialized  areas, 
such  as  aircraft  fuels,  aircraft  instruments,  and  aircraft  operating  problems. 
The  NACA  also  had  special  committees,  usually  ad  hoc,  that  dealt  with  ex- 
traordinary problems  such  as  the  need,  in  1938,  to  build  new  facilities  to 

9 


Spaceflight  Revolution 

meet  the  threat  of  another  world  war.  Twenty  years  later,  in  the  middle  of 
another  international  crisis,  the  NAG  A  had  a  special  committee  working  to 
explore  the  ramifications  of  Sputnik  and  to  help  formulate  a  space  policy 
for  the  NACA. 

The  committee  system  did  not  work  perfectly,  but  in  its  unique  way  it 
did  work.  Prominent  people  in  the  American  aviation  enterprise  became 
familiar  with  NACA  capabilities  and  NACA  results;  concurrently,  the 
NACA  benefited  from  the  insight  of  many  talented  and  experienced  men 
(no  women  ever  served  on  any  of  the  NACA  committees).  Further,  the 
connections  and  the  prestige  of  committee  members  helped  the  NACA 
to  win  friends  and  secure  appropriations  from  Congress.  Over  the  years, 
outsiders  such  as  the  Brookings  Institution,  self-styled  experts  in  government 
organization,  and  several  officers  in  the  Bureau  of  the  Budget  had  viewed 
the  committee  system  of  advise  and  consent  as  a  messy  way  to  structure  and 
manage  a  federal  agency.  But  NACA  insiders  did  not.  Nothing  about  the 
committee  system  meddled  seriously  in  any  unwelcome  fashion  with  work 
in  the  laboratories.  The  actual  management  of  the  research  operation  was 
left  to  the  civil  servants  who  worked  full-time  for  the  NACA.  Within  the 
laboratory  itself,  management  was  left  to  the  engineer-in-charge.* 

At  the  Washington  level,  the  management  of  research  was  left  to  the 
NACA's  director  of  research.  Only  two  men  held  this  post  during  the 
NACA's  43-year  history.  Dr.  George  W.  Lewis  (honorary  doctorate  from 
Swarthmore,  his  alma  mater)  held  the  post  from  the  time  it  was  established 
in  1919  until  his  retirement  in  1947.  Dr.  Hugh  Dryden  (one  of  the  youngest 
Ph.D.'s  ever  to  come  out  of  Johns  Hopkins  University,  in  1919,  at  age  21) 
served  from  1947  to  1958.  These  two  men,  of  very  different  backgrounds, 
demeanors,  and  talents,  guided  the  NACA  through  the  rapid  technological 
evolution  and  sudden  revolutions  that  in  less  than  half  a  century  had  taken 
aeronautics  on  a  turbulent  whirlwind  from  the  era  of  wooden  biplanes, 
ponderous  airships,  and  subsonic  flight  into  the  age  of  jets,  supersonics, 
and  rockets  at  the  edge  of  spaceflight.11 

Most  critics  agreed  that  the  NACA  had  served  the  general  cause  of 
American  aeronautics  well  for  more  than  40  years.  But  now  in  the  wake 
of  Sputnik,  they  felt  the  time  had  come  for  a  major  reorganization  and  the 
injection  of  new  blood.  By  early  1958,  a  growing  number  of  American  leaders 
joined  in  that  opinion  and  were  ready  to  tell  the  NACA  thanks,  slap  it  on 
the  back,  and  bring  its  experiment  in  government  organization  to  an  end. 
A  bold  new  initiative  was  required  if  the  United  States  was  to  catch  up  to 
the  Soviet  Union.  Space  enthusiast  Senator  Lyndon  B.  Johnson,  chairman 
of  the  Senate's  new  Special  Committee  on  Space  and  Astronautics,  felt  this 


In  1948  civil  service  requirements  had  forced  the  NACA  to  change  the  old  title  to  director.  No 
one  liked  the  change,  certainly  not  Langley's  top  man,  Henry  Reid,  who  had  been  engineer-in-charge 
for  22  years,  since  1926.  The  old  title  had  made  it  clear  that  an  engineer,  not  a  scientist,  headed  the 
organization. 

10 


The  Metamorphosis 


L-90-3751 

The  NA  CA  's  directors  of  Research,  George  W.  Lewis  (left)  and  his  successor,  Dr. 
Hugh  L.  Dry  den  (right). 

way,  as  did  others.  They  claimed  that  the  old  NACA  was  too  timid  and  too 
conservative  about  exploring  the  potential  of  space.  Such  critics,  as  well  as 
some  "young  Turks"  inside  the  NACA,  felt  that  if  the  organization  was  to  be 
reincarnated  as  NASA,  then  it  should  be  revamped  with  new  personnel  and 
additional  facilities  and  charged  up  by  new  leaders.12  Out  of  this  general 
sentiment  for  major  change  came  the  National  Aeronautics  and  Space  Act 
of  1958.  The  Space  Act  gave  NASA  an  advisory  board,  but  insiders  knew 
it  could  not  be  the  same  as  it  was  under  the  NACA.  On  NASA's  first  day, 
1  October  1958,  the  NACA  committee  system  was  essentially  discarded. 


Glennan:  Welcome  to  NASA 

At  the  head  of  NASA  was  Dr.  T.  Keith  Glennan.  When  Eisenhower 
announced  Glennan  as  his  choice  for  the  NASA  administrator  on  9  August 
1958,  people  at  Langley  and  at  other  NACA  centers  asked,  who  was 
Glennan?  They  learned  that  he  was  the  president  of  Case  Institute  of 
Technology  in  Cleveland.  Then  he  must  be  a  member  of  the  NACA  Main 
Committee?  No,  he  was  a  former  Hollywood  movie  mogul  and  a  minor  one 
at  that,  not  in  the  class  of  a  Samuel  Goldwyn  or  Louis  B.  Mayer. 

These  answers,  which  circulated  via  the  NACA  grapevine  late  in  the 
summer  of  1958,  appalled  some  NACA  employees,  did  not  make  much  sense 
to  most,  and  made  none  of  them  very  happy.  In  its  15  August  edition, 
the  Langley  Air  Scoop,  the  in-house  newspaper,  ran  a  picture  of  53-year-old 


11 


Space/light  Revolution 

Glennan  along  with  a  complete  biographical  sketch  provided  by  Case  In- 
stitute of  Technology.  Reading  this  article,  Langley  employees  found  that 
Glennan  indeed  had  been  a  manager  for  Paramount  and  Samuel  Goldwyn 
studios  during  World  War  II,  but  that  his  overall  career  was  marked  by 
"achievements  in  business,  education,  and  the  administration  of  scientific 
research."13  In  recent  years  he  had  served  on  the  Atomic  Energy  Commis- 
sion and  on  the  board  of  the  National  Science  Foundation,  and  he  was  sup- 
posed to  have  excellent  connections  in  Washington.  Considering  the  highly 
charged  and  politicized  atmosphere  now  surrounding  everything  that  had  to 
do  with  rockets  and  space,  something  finally  made  sense  about  Glennan's  se- 
lection. At  ceremonies  held  in  the  White  House  on  Tuesday,  19  August,  Dr. 
Glennan  raised  his  right  hand,  put  his  left  on  a  Bible,  and  pledged  the  oath 
as  NASA  administrator.  On  the  same  Bible,  close  enough  to  touch  the 
ends  of  his  fingers,  was  the  left  hand  of  faithful  Methodist  lay  minister  Dr. 
Hugh  Dryden,  the  NACA's  director  of  research.  Although  many  in  Congress 
wanted  Dryden  out  of  the  picture  because  they  thought  that  his  quiet,  al- 
most mousy  personality  and  conservative  approach  to  launching  America 
into  space  might  tarnish  the  images  of  youthfulness,  dynamism,  and  bold- 
ness they  wanted  for  NASA,  Glennan  had  insisted  on  making  him  his  deputy 
administrator,  and  Dryden  had  accepted.14  Glennan  thought  that  this  selec- 
tion would  help  provide  continuity  and  make  the  metamorphosis  into  NASA, 
as  well  as  his  own  administration,  easier  for  NACA  people  to  accept.  Other 
NACA  headquarters  officers  came  to  NASA  with  Dryden,  including  John  F. 
Victory,  the  Main  Committee's  fastidious  executive  secretary  and  first  em- 
ployee. (Victory  had  been  working  for  the  NACA  since  1915.)  Some  viewed 
President  Eisenhower's  appointment  of  Jimmy  Doolittle,  the  last  NACA 
chairman,  to  his  nine-member  National  Aeronautics  and  Space  Council  as 
another  gesture  toward  the  NACA  old  guard.  For  Eisenhower,  however, 
the  appointment  of  Doolittle  was  more  than  a  gesture.  Ike  knew  Doolittle, 
his  former  World  War  II  air  force  commander  in  North  Africa  and  Europe; 
trusted  his  judgment;  and  wanted  his  moderate,  reasonable,  and  experienced 
voice  on  the  newly  formed  space  council. 

On  the  morning  of  1  October  1958,  not  a  single  member  of  the  Langley 
senior  staff  was  likely  to  have  remembered  ever  meeting  Glennan.  The  new 
NASA  administrator  had  not  yet  visited  Langley  or  any  other  NACA  facility, 
at  least  not  as  the  NASA  administrator.  However,  the  former  Hollywood 
executive  had  appeared  at  Langley  via  motion  picture.  On  22  September, 
the  NACA  public  affairs  officer  in  Washington,  Walter  Bonney,  sent  copies 
of  a  short  10-minute  film,  "Glennan  Message  to  NACA  Employees,"  for 
immediate  showing  at  all  NACA  centers.15 

At  Langley,  employees  gathered  in  the  East  Area  a  few  days  later  to  watch 
the  film  in  the  air  force  base's  air-conditioned  theater,  next  to  the  old  19- 
Foot  Pressure  Tunnel,  which  dated  to  1939.  From  its  beginning,  something 
about  the  film  made  many  people  in  the  audience  uneasy.  Perhaps  they  were 
disturbed  by  the  Orwellian  undertone  of  the  presentation,  a  confident  and 

12 


The  Metamorphosis 


L-58-14a 
Glennan  introduces  himself  to  the  Langley  staff  via  motion  picture. 


soothing  "Big  Brother"  message  coming  to  the  people  electronically  from 
the  center  of  government.  This  message  did  not  come  from  the  NACA's 
staid  old  headquarters  at  1512  H  Street  NW  in  Washington  (referred  to 
as  the  "Washington  office"),  but  rather  from  Glennan's  new  deluxe  office 
within  the  recently  acquired  suite  of  NASA  administrative  offices  in  the 
Dolly  Madison  House  at  nearby  1520  H  Street  NW.  Word  had  circulated 
that  Glennan  had  had  his  office  suite  decorated  just  like  the  one  he  had 
enjoyed  as  president  of  Case  Institute  of  Technology. 

The  movie  opened  with  the  NASA  administrator  leaning  on  the  front  of 
his  desk.  "I  very  much  want  to  talk  with  you  about  our  future,"  Glennan 
began.  But  before  he  described  "the  mighty  big  job"  that  lay  ahead  for 
NASA,  he  took  time  to  praise  the  NACA.  He  explained  that  during  his  11 
years  at  Case  Institute  of  Technology  in  Cleveland,  he  had  worked  with  many 
people  at  NACA  Lewis.  He  was  both  "familiar  with  [the]  NACA's  traditions 
and  accomplishments"  and  "impressed  by  the  high  state  of  morale  and  by 
the  vigor"  with  which  the  NACA  conducted  its  research.16  Glennan  failed 
to  mention,  however,  what  he  would  soon  record  in  his  personal  diary,  his 
opinion  that  the  NACA  staff  was  "composed  of  reasonably  able  people," 
lacking  experience  in  the  "management  of  large  affairs."17  According  to 

13 


Spaceflight  Revolution 


L-59-22 

Glennan's  first  live  appearance  at  Langley  in  early  January  1959.  Here  he  is  being 
welcomed  by  Henry  Reid,  the  center  director. 

one  member  of  the  Langley  senior  staff,  Glennan  "had  so  little  knowledge 
of  the  organization"  at  the  outset  that  he  did  not  think  its  staff  "had 
any  competence."  Upon  seeing  the  huge  vacuum  spheres  belonging  to  the 
Gas  Dynamics  Laboratory  at  Langley,  Glennan  allegedly  remarked,  "NASA 
doesn't  have  any  capability  to  handle  that  kind  of  high  pressure  stuff.  You're 
going  to  have  to  get  some  help  from  outside  to  do  that,  you  know."18 

Despite  his  true  feelings,  Glennan  stressed  in  his  message  that  "NASA 
must  be  like  [the]  NACA  in  the  qualities  of  strength  and  character  that 
make  an  organization  great,"  but  he  also  emphasized  the  arrival  of  "a  new 
day"  at  Langley.  To  describe  that  new  day,  the  NACA's  changeover  to 
NASA,  Glennan  quoted  from  what  he  called  the  "legalistic  language"  of  the 
Space  Act:  "the  NACA  shall  cease  to  exist"  and  "all  functions,  powers, 
duties,  and  obligations  and  all  real  and  personal  property,  personnel  (other 
than  members  of  the  Committee),  funds,  and  records"  of  the  NACA  were 
to  be  transferred  to  NASA.  But,  he  explained,  he  preferred  to  think  of  it 
differently:  "I  would  like  to  say,  and  I  believe  that  I  am  being  very  realistic 
and  very  accurate  when  I  do,  that  what  will  happen  September  30  is  a  sign 


14 


The  Metamorphosis 

of  metamorphosis.  [It  is]  an  indication  of  the  changes  that  will  occur  as  we 
develop  our  capacity  to  handle  the  bigger  job  that  is  ahead."19 

The  bigger  job  was  outlined  in  the  Space  Act,  which  he  encouraged  all 
NACA  employees  to  take  the  time  to  read,  at  least  its  first  few  pages. 
The  job  included  the  "expansion  of  human  knowledge  about  space  . . . 
development  and  operation  of  vehicles  capable  of  carrying  instruments  and 
man  through  space  . . .  long-range  studies  of  the  benefits  of  using  aeronautical 
and  space  activities  for  peaceful  and  scientific  purposes  . . .  preservation  of 
the  role  of  the  United  States  as  a  leader  in  aeronautical  and  space  science 
and  technology."  Glennan  also  outlined  the  metamorphosis.  The  NACA's 
vital  function,  research  into  the  problems  of  atmospheric  flight,  would  now 
become  "only  one  part  of  NASA's  activities."  To  accomplish  the  goals 
set  out  in  the  Space  Act,  NASA  would  have  to  add  "new  and  extremely 
able  people"  to  its  staff;  administer  "substantial  programs  of  research  and 
development  and  procurement  with  others  on  a  contract  basis";  spend 
"large  amounts  of  money  outside  the  agency  by  contracts  with  scientific 
and  educational  institutions  and  with  industry" ;  use  military  facilities  "such 
as  the  launching  pads  at  Cape  Canaveral";  and  operate  satellite-tracking 
stations  around  the  world.  All  this  and  more  had  to  be  done  and  quickly 
in  preparation  for  a  manned  flight  into  space  and  exploration  into  the  Solar 
System.20 

Finally,  Glennan  tried  to  end  his  message  on  a  high  note  by  quoting 
from  a  speech  that  Lyndon  Johnson  made  in  August  during  the  Senate 
confirmation  hearings  of  the  top  two  NASA  officials: 

There  are  no  blueprints  or  roadmaps  which  clearly  mark  out  the  course.  The  limits 
of  our  job  are  no  less  than  the  limits  of  the  universe.  And  those  are  limits  which 
can  be  stated  but  are  virtually  impossible  to  describe.  In  a  sense,  the  course  of  the 
new  Agency  can  be  compared  to  the  voyage  of  Columbus  to  the  New  World.  The 
only  difference  is  that  Columbus  with  his  charts  drawn  entirely  from  imagination  had 
a  better  idea  of  his  destination  than  we  can  possibly  have  when  we  step  into  outer 
space. 

Most  NACA  employees  filing  out  of  the  base  theater  felt  positive  and  excited 
about  what  they  had  heard,  but  a  few  cynics  might  have  wondered  out  loud 
about  that  last  reference  to  Columbus:  "Wasn't  he  headed  for  China?  And 
didn't  he  believe  to  his  dying  day  that  he  had  landed  in  Asia?"  Hopefully, 
NASA  had  a  better  idea  of  its  destination  and  would  know  where  it  was 
when  it  got  there. 


Air  versus  Space 

NACA  explorers,  unlike  Columbus,  had  a  good  idea  of  where  they  were 
going.  They  were  going  into  the  air  faster,  farther,  higher,  and  more 
efficiently  in  a  modern  engineering  marvel  that  their  systematic  research  into 

15 


Spaceflight  Revolution 


In  this  1925  photo,  NACA  pi- 
lot Paul  King,  donned  in  a  fur- 
lined  leather  flight  suit  with  oxy- 
gen facepiece,  is  ready  to  test  a 
Vought  VE-7. 


L-1118 


aeronautics  over  the  last  43  years  had  helped  to  make  possible.  Aeronautics 
and  the  NACA  had  grown  up  together;  the  business  of  the  NACA  for  its 
entire  existence  had  been  to  see  that  American  aeronautics  continued  to 
progress.  For  NACA  veterans  who  took  Glennan's  advice  and  read  the 
Space  Act  of  1958,  the  time  when  the  airways  had  been  ruled  by  frail 
wooden  biplanes  covered  with  fabric,  braced  by  wires,  powered  by  heavy 
water-cooled  engines,  and  driven  by  hand-carved  wooden  propellers  did 
not  seem  so  long  ago.  When  20-year-old  Floyd  Thompson  served  as  a 
mechanic  in  Pensacola  with  the  U.S.  Navy's  first  torpedo  squadron  in  1918, 
the  navy's  fastest  aircraft,  an  R6L  biplane  amphibian,  had  a  top  speed  of 
110  knots  and  a  fuel  system  with  a  windmill  on  the  outside  to  pump  fuel 
up  to  an  overhead  gravity  tank.  When  flight  research  operations  began  at 
NACA  Langley  a  year  later,  NACA  researchers  hardly  knew  the  principles 
of  aeronautical  engineering.  Airplane  design  was  still  a  largely  intuitive  and 
empirical  practice,  thus  requiring  bold  speculation  and  risk  taking.  In  1920 
the  Langley  staff  copied  the  design  of  an  existing  wind  tunnel  at  the  British 

16 


The  Metamorphosis 

National  Physical  Laboratory  to  fashion  their  Wind  Tunnel  No.  1  because 
no  one  at  the  NACA  knew  how  to  design  a  wind  tunnel.22 

In  the  decades  that  followed,  the  NACA  designed  more  wind  tunnels 
than  staff  members  could  count  (many  of  them  unique  facilities)  and 
authored  more  reports  on  aeronautical  technology  than  any  other  single 
institution  in  the  world.23  With  the  aerodynamic  information  that  these 
tunnels  and  technical  reports  provided,  American  universities  educated 
most  of  the  country's  aeronautical  engineers,  and  U.S.  industry  became 
the  world  leader  in  the  manufacture  of  aircraft.  By  NASA's  first  day, 
the  NACA  had  helped  to  advance  aeronautics  far  beyond  the  primitive 
state  of  flight  at  the  end  of  World  War  I.  Commercial  jet  airliners  were 
beginning  to  fly  passengers  comfortably  around  the  world  in  pressurized 
cabins.  Sleek  military  jets  streaked  across  the  skies  at  speeds  in  excess  of 
Mach  1,  greater  than  the  speed  of  sound.  In  fact,  two  McDonnell  F-101A 
supersonic  jet  fighters  were  being  made  ready  in  the  hangar  for  further  flight 
testing.  (The  F-101A  was  nicknamed  "Voodoo"  but  known  to  enthusiasts 
as  the  "One-O- Wonder.")  Langley  acoustics  specialists  Domenic  Maglieri, 
Harvey  Hubbard,  and  Donald  Lansing  were  taking  ground  measurements 
of  the  shock- wave  noise  produced  by  one  of  the  F- 101  As  in  level  flight  at 
speeds  up  to  Mach  1.4  and  altitudes  up  to  45,000  feet.  A  team  of  engineers 
and  technicians  supervised  by  Langley  Assistant  Director  Hartley  "Buster" 
Soule,  the  NACA  Research  Airplane  Project  (RAP)  leader,  was  evaluating 
several  control  systems  for  the  North  American  XB-70  Valkyrie,  a  gigantic 
high-altitude,  delta-winged  bomber  of  some  550,000  pounds  to  be  built  of 
titanium  and  stainless  steel  and  capable  of  flying  to  Mach  3.24 

As  the  federal  agency  responsible  for  the  progress  of  the  nation's  aviation 
technology,  the  NACA  had  enough  to  do  without  getting  involved  in  what 
the  public  considered  "Buck  Rogers  stuff."*  During  the  first  four  decades 
of  Langley's  operation,  the  idea  of  working  to  promote  the  immediate 
achievement  of  spaceflight  had  been  too  ridiculous. for  consideration.  Into 
the  1940s,  NACA  researchers  were  not  certain  that  rockets  and  missiles  were 
a  part  of  aeronautics.  Langley  veteran  Christopher  C.  Kraft,  Jr.'  (the  "C" 
stood  for  "Columbus"),  who  later  became  famous  as  "The  Voice  of  Project 
Mercury"  and  the  director  of  NASA's  manned  spaceflight  operations  at 
Mission  Control  in  Houston,  remembers  that  before  the  late  1950s  "space" 
was  a  dirty  word:  "[It]  wasn't  even  allowed  in  the  NACA  library.  The 
prevailing  NACA  attitude  was  that  if  it  was  anything  that  had  to  do  with 
space  that  didn't  have  anything  to  do  with  airplanes,  [then]  why  were  we 


Younger  readers  may  need  to  know  that  Buck  Rogers  was  a  science-fantasy  comic  strip  created  by 
Dick  Calkins  around  1930;  the  comic  strip  remained  popular  until  it  was  terminated  in  the  1960s.  In  the 
1950s,  it  also  became  a  popular  television  "space  opera."  As  such,  "Buck  Rogers"  significantly  influenced 
American  popular  culture's  attitudes  about  rocketry  and  space  travel.  (In  the  late  1970s,  another  TV 
show,  "Buck  Rogers  in  the  21st  Century,"  went  on  the  air;  however,  the  updated  character  did  not  bring 
on  a  similar  craze.) 

17 


Space/light  Revolution 


L-3310 

In  this  photo  taken  on  15  March  1922,  NACA  researchers  conduct  tests  on  airfoils 
in  the  Variable- Density  Tunnel,  a  revolutionary  new  test  chamber  that  permitted,  for 
the  first  time  anywhere  in  the  world,  aerodynamic  testing  at  approximately  full-scale 
conditions. 


L-37,931 

Drag-cleanup  testing  of  America's  first  jet  airplane,  the  Bell  P-59,  is  conducted  in 
the  Full- Scale  Tunnel,  May  1944- 

18 


The  Metamorphosis 

working  on  it?"25  One  Langley  veteran,  Ira  H.  Abbott,  recalled  that  the 
NACA  stood  "as  much  chance  of  injecting  itself  into  space  activities  in  any 
real  way  as  an  icicle  had  in  a  rocket  combustion  chamber."  In  the  early 
1950s,  Abbott  had  mentioned  the  possibility  of  manned  spaceflight  to  a 
House  subcommittee,  and  one  of  the  congressmen  scornfully  accused  him  of 
talking  "science  fiction."26 

Nevertheless,  by  the  early  1950s,  the  NACA  had  become  seriously  in- 
volved in  the  study  of  rockets,  missiles,  and  the  potential  of  spacefiight; 
all  of  these  topics  related  to  aeronautics.  Anything  that  concerned  the 
science  and  technology  of  flight,  whether  it  be  in  the  atmosphere  or  be- 
yond, eventually  became  an  interest  of  the  NACA.  In  the  months  fol- 
lowing Sputnik,  NACA  leaders  tried  to  capitalize  on  the  agency's  re- 
search into  spaceflight  to  justify  a  central  role  in  whatever  space  program 
came  into  existence.  Acting  prudently  on  behalf  of  their  institution,  the 
NACA  Langley  management  and  most  staff  members  did  everything  possi- 
ble to  convince  everyone  concerned,  including  the  new  NASA  administrator, 
T.  Keith  Glennan,  that  the  old  NACA  laboratory  could  do  and  already  was 
doing  a  great  deal  more  than  playing  with  airplanes. 

For  example,  in  January  1958,  only  four  months  after  the  launching 
of  Sputnik  1,  a  special  Langley  committee,  surveying  current  and  pending 
projects,  documented  the  NACA's  transition  to  space  research.  Chairing  the 
committee  was  Langley  Assistant  Director  Robert  R.  Gilruth,  the  future 
head  of  Project  Mercury,  America's  first  manned  space  program.  Also 
serving  on  this  committee  were  Eugene  Draley,  head  of  the  laboratory's  Full- 
Scale  Research  Division  (and  soon  to  succeed  Robert  R.  Gilruth  as  assistant 
director  for  the  Dynamic  Loads,  Pilotless  Aircraft  Research,  and  Structures 
Research  Divisions);  John  V.  Becker,  chief  of  Langley 's  Compressibility 
Research  Division;  and  Charles  J.  Donlan,  technical  assistant  to  Associate 
Director  Thompson.  The  in-house  review  covered  the  activities  of  all  11 
Langley  research  divisions  during  fiscal  years  1955  and  1957,  as  well  as 
projected  activities  for  fiscal  year  1959.  Two  tables  of  numbers  accompanied 
the  committee's  final  report  to  Director  Reid,  and  the  more  important  of 
the  two  indicated  that  the  "research  effort"  in  the  fields  of  hypersonics  and 
spaceflight  should  increase  from  about  11  percent  in  1955  to  54  percent 
in  1959;  however,  it  was  unclear  what  these  percentages  actually  meant 
in  terms  of  money  and  personnel  hours.  In  fact,  Langley  management 
derived  these  percentages  from  hours  spent  on  projects  in  the  three  research 
directorates. 

According  to  the  review,  the  two  most  important  fields  of  application 
were  satellites  and  spacecraft,  and  ballistic  missiles.  Efforts  in  these  areas 
were  to  rise  from  less  than  1  percent  to  16  percent  and  from  3  percent  to  14 
percent,  respectively.  In  the  words  of  the  committee  members,  "all  research 
divisions  are  adjusting  and  reorienting  manpower,  curtailing  work  in  areas 
of  lesser  importance  [and]  continually  studying  and  developing  the  special 
facilities  needed  to  attack  these  problems,"  and  each  division  had  been  doing 

19 


Space/light  Revolution 

so  for  some  time.  "The  ability  to  reorient  the  Laboratory's  efforts  to  the 
extent  shown  in  the  brief  time  period  considered,"  the  report  concluded, 
"is  due  to  a  considerable  extent  to  active  planning  for  a  number  of  these 
[space-related]  fields  during  recent  years."27 

Langley  senior  management  knew  that  these  figures  were  authentic.  The 
transition  to  space  was  happening  at  Langley,  and  it  had  been  happening 
there  even  before  Sputnik.  Senior  management  also  knew  that  more  than 
a  little  finagling  was  done  to  get  the  space  numbers  up  as  high  as  possible, 
because  they  were  doing  the  finagling.  What  was  applicable  to  "space"  and 
what  was  applicable  to  "aeronautics"  depended  on  how  they  defined  the 
research  programs  and  divided  the  disciplines;  to  differentiate  was  splitting 
hairs.  The  Gilruth  committee  discovered,  in  January  1958,  that  much  of 
the  work  at  the  laboratory,  initially  instigated  to  support  what  the  NACA 
had  always  called  the  "aeronautics  program,"  could  in  fact  be  conveniently 
reclassified  as  space  research.  In  addition,  Langley  was  working  on  many 
projects  that  honestly  involved  both  aeronautics  and  space  (truly  "aero- 
space" research),  yet  could  be  classified  as  one  or  the  other  depending  on 
what  the  center  desired  to  emphasize.28  In  the  post-Sputnik  era  of  national 
debate  over  the  makeup  of  a  new  space  agency,  now  was  unquestionably 
the  time  to  emphasize  space,  an  emphasis  on  which  Langley's  future  would 
depend. 

However,  almost  no  one  at  Langley  on  the  first  day  of  NASA  would  have 
thought  that  the  time  had  come  to  abandon  the  quest  for  improved  aero- 
nautical performance.  Many  great  technological  advances  remained  to  be 
achieved  in  aeronautics:  greater  speeds,  bigger  airplanes,  and  superior  flight 
efficiencies.  Already  in  flight  were  radically  new  aircraft  like  Lockheed's 
supersonic  F-104  Starfighter,  the  still-secret  U-2  strategic  reconnaissance 
"spy  plane,"  and  Convair's  B-58  delta-winged  bomber,  which  was  capable 
of  Mach  2.  On  the  horizon  were  important  developments,  such  as  new  heli- 
copter applications,  tilt  wing,  and  other  innovative  vertical  and  short  take- 
off and  landing  (V/STOL)  capabilities.  Additionally,  new  high-performance 
wings  with  unusual  degrees  of  backward  and  even  forward  sweep  were  being 
designed  at  Langley  and  elsewhere.  One  of  the  wings  of  the  future  would 
probably  have  some  form  of  variable  sweep,  like  those  Langley's  foremost 
expert  on  high-speed  aerodynamics,  John  Stack,  had  seen  on  a  model  of  the 
arrow-winged  Swallow  aircraft  in  England.  This  wing  would  no  doubt  be 
part  of  a  commercial  supersonic  transport  (SST)  that  before  too  long  would 
be  taking  airline  passengers  from  New  York  to  London  or  Paris  in  a  few 
hours.29  Even  more  dear  to  the  heart  of  some  aerospace  enthusiasts  was 
the  first  of  the  next  generation  of  research  airplanes,  North  American  Avi- 
ation's rocket-powered  X-15,  designed  for  the  exploration  of  the  hypersonic 
speed  regime  up  to  Mach  6,  as  well  as  the  hypersonic  boost-glider  program, 
known  as  Project  Dyna-Soar,  sponsored  jointly  by  the  U.S.  Air  Force  and 
NASA.  In  one  of  these  "envelopes,"  many  NACA/NASA  engineers  felt,  an 
American  might  first  fly  into  space.30 

20 


The  Metamorphosis 


L-59-3817 

A  model  of  the  Swallow  arrow-wing  aircraft  is  tested  in  the  16-Foot  Transonic  Tunnel 
in  June  1959.  The  British  hoped  that  a  research  airplane  derived  from  the  Swallow 
configuration  would  be  the  progenitor  of  a  commercial  SST. 


L-58-2951 

In  1958  two  Langley  researchers  install  a  one-tenth  scale  model  of  the  X-15  rocket 
plane  in  the  Langley  7  x  10-Foot  High-Speed  Tunnel  to  study  its  spin  characteristics. 


21 


Space/light  Revolution 

Clearly,  now  was  no  time  to  take  a  hiatus  from  aeronautics.  Although 
many  congressional  leaders  and  probably  even  the  American  people  as  a 
whole  forgot  the  second  word  in  the  National  Aeronautics  and  Space  Act, 
calling  it  "the  Space  Act,"  most  of  the  research  staff  at  Langley  took  a 
different  view.  As  preliminary  drafts  of  the  Space  Act  made  their  way  to 
the  NACA  laboratory  for  review  in  the  spring  and  early  summer  of  1958, 
aeronautically  oriented  staff  members  like  RAP  leader  Hartley  Soule  and 
supersonics  pioneer  John  Stack  read  them  and  said  to  one  another,  "Well, 
we're  not  doing  that.  Let  those  guys  [up  in  Washington]  go  ahead  and  write 
it  up,  [but]  we'll  just  [keep  doing]  what's  necessary  and  get  on  with  the 
program."  Unlike  the  ardent  space  buffs,  these  men  read  the  Space  Act  to 
mean  that  they  "were  supposed  to  pick  up  the  space  program"  in  addition 
to  aeronautics  not  that  they  "were  supposed  to  get  out  of  aeronautics."31 

A  few  days  after  passage  of  the  Space  Act,  U.S.  Army  representatives 
visited  Langley  to  find  out  who  was  going  to  take  care  of  their  aircraft 
engine  problems  now  that  the  NACA  was  about  to  be  dissolved  in  favor  of  a 
space  agency.  The  surprised  Langley  people  answered,  "Well,  we  are!  We're 
here  and  we  know  what  we  are  doing,  and  under  NASA,  we  will  just  keep 
doing  it."32  That  literal  view  of  the  Space  Act  calmed  the  military  visitors 
and  reassured  their  hosts.  If  Langley  people  had  known  that  the  national 
commitment  to  space  was  going  to,  "backburner"  their  traditionally  strong 
aeronautical  programs  for  years  to  come,  they  might  not  have  responded  so 
glibly  to  questions  about  the  changeover. 

In  the  following  years,  the  aeronautics  effort  at  Langley  decreased 
significantly;  at  its  lowest  level,  it  shrank  to  about  25  percent  of  the  center's 
total  labor  hours.  Nonetheless,  aeronautics  was  never  allowed  to  die  at 
Langley.  Even  during  the  rushed  days  of  the  Apollo  lunar  landing  program 
in  the  1960s,  fruitful  aeronautical  programs  quietly  proceeded  behind  the 
scenes.  Langley  managed  to  retain  a  dedicated  cadre  of  aeronautical  people 
even  when  NASA  recruited  talent  primarily  in  support  of  the  space  program. 
But  for  John  Stack,  Hartley  Soule,  and  likewise  air-minded  NACA  veterans, 
aeronautical  research  would  often  seem  nearly  forgotten  at  Langley. 


The  Public  Eye 

Most  of  those  working  in  aviation  knew  about  the  NACA  through 
exposure  to  NACA  reports  and  articles  concerning  NACA  research  in 
aeronautical  engineering  magazines  and  other  trade  journals.  But  none  of 
the  NACA's  operations  had  high  public  profiles,  not  even  at  the  local  level. 
Until  1958  most  Americans  knew  nothing  about  the  NACA.  Before  World 
War  II,  some  congressmen  did  not  know  it  existed.  Even  the  people  near 
Langley  Field  ignored  the  place.  As  Langley  engineer  and  Hampton  native 
Caldwell  (pronounced  Cad-well)  Johnson  remembers,  "It  [the  NACA]  wasn't 
like  NASA.  The  press  didn't  care  about  it — to  them  it  was  a  dull  bunch 

22 


The  Metamorphosis 

of  gray  buildings  with  gray  people  who  worked  with  slide  rules  and  wrote 
long  equations  on  the  board."  Brain-busters  like  that  were  better-off  left 
alone.  Ironically,  throughout  its  entire  history,  the  only  time  the  NACA 
was  a  high-profile  agency  was  after  Eisenhower  had  selected  it  as  the  nucleus 
for  NASA. 

At  times  the  NACA's  obscurity  put  the  agency  at  a  disadvantage.  The 
NACA  could  not  rely  on  the  strength  of  favorable  public  opinion  in  its 
campaigns  for  appropriations;  such  battles  had  to  be  fought  and  won  quietly 
in  private  conferences  in  hallways  or  smoke-filled  rooms  with  admirals, 
generals,  and  congressmen.  These  "gold-braided  personages"  made  the  case 
for  the  NACA  to  Congress,  when  it  was  necessary  for  a  case  to  be  made. 

Handling  much  of  this  delicate  politicking  from  1919  until  his  retirement 
in  1947  was  the  NACA's  shrewd,  cigar-smoking  director  of  research,  "Doc" 
Lewis  (1892-1948).  Although  the  gregarious  Lewis  and  his  successor,  the 
quieter  and  scientifically  sharper  Dr.  Hugh  Dry  den,  usually  acquired  the 
necessary  backing  for  NACA  projects,  they  experienced  many  close  calls. 
The  closest  one  came  in  December  1932  when  President  Herbert  Hoover, 
looking  to  reduce  expenditures  and  increase  efficiency  in  government,  had 
ordered  the  NACA  abolished  and  most  of  its  resources  handed  over  to  the 
Bureau  of  Standards.  However,  House  Democrats,  anticipating  the  first 
term  of  Franklin  D.  Roosevelt,  overrode  the  lame-duck  executive  order,  and 
the  NACA  survived.34 

On  balance,  however,  the  advantages  of  the  NACA's  invisibility  out- 
weighed the  few  disadvantages.  It  certainly  benefited  the  researchers;  most 
of  them  thought  NACA  Langley  was  a  wonderful  place  to  work  and  "just 
a  splendid  organization."35  Although  administrative  policies  and  bureau- 
cratic guidelines  involving  anything  related  to  the  laboratory's  communi- 
cation with  the  outside  world  (such  as  mail,  telephone  calls,  and  technical 
reports)  were  rather  prescriptive,  considerable  leniency  existed  in  the  per- 
formance of  in-house  research.  Individuals  could  follow  their  own  ideas  quite 
far  without  formal  approval  from  superiors.  Any  scheme  that  survived  peer 
discussion  and  won  the  approval  of  the  research  section  was  likely  to  be 
implemented.  If  funding  was  not  formally  available  to  build  a  given  wind- 
tunnel  model,  flight  instrument,  minor  test  facility  component,  or  the  like, 
employees  were  usually  able  to  "bootleg"  what  they  needed  from  resources 
appropriated  to  approved  projects.  As  long  as  the  initiative  offered  some- 
thing promising,  did  not  cost  too  much,  and  did  not  have  the  potential  to 
get  the  NACA  into  real  trouble,  NACA  managers  rarely  complained  or  put 
tight  reins  on  the  researchers.  Within  the  laboratory,  few  barriers  limited 
innovation  and  the  free  dissemination  of  knowledge;  the  young  engineers 
could  discuss  their  work  comfortably  with  everyone  from  the  technicians  in 
the  shops  to  the  division  chief.36 

Such  freedoms  existed  because  neither  the  NACA's  own  management, 
other  government  bureaucrats,  nor  newspaper  or  magazine  journalists  (or 
the  American  people  as  a  whole)  spent  much  time  looking  over  the  shoulders 

23 


Space/light  Revolution 

of  NACA  researchers.  The  NACA  shared  what  it  did  with  major  clients; 
the  how  was  kept  more  or  less  within  the  NACA  itself.  Moreover,  almost 
none  of  NACA  Langley's  research  work  involved  contracts  with  outsiders; 
everything  was  accomplished  in-house.  As  Caldwell  Johnson  has  noted 
about  the  NACA,  "It  had  the  best  wind  tunnels,  the  best  model-builders, 
the  best  technicians,  the  most  rigorous  standards."  Nothing  gave  Langley 
people  more  pride  than  being  a  part  of  such  an  autonomous  organization.37 

If  Langley  engineers  had  cultivated  any  public  image  before  NASA,  it 
had  been  that  of  the  "NACA  Nuts."  All  the  local  hardware  salesmen  and 
auto  dealers  recognized  them  a  mile  away,  and  if  it  had  not  been  for  the 
federal  paychecks  that  the  NACA  folks  brought  to  the  local  economy,  the 
natives  would  have  dreaded  to  see  them  coming.  Not  only  were  most  NACA 
Nuts  overeducated  Yankees,  they  were  brilliant  technical  types  who  wanted 
to  know  the  revolutions  per  minute  (rpm)  of  their  vacuum  sweepers  and 
ordered  lumber  cut  to  the  sixteenth  of  an  inch.  Funny  stories  about  their 
eccentricities  abounded,  leading  everyone  from  Yorktown  to  Newport  News 
to  think  that  anyone  from  the  NACA  had  to  be  either  a  weirdo  or  a  screwball. 

The  truth  was  that  most  locals  in  those  days  had  not  the  faintest  idea 
what  the  NACA  people  did.  Few  residents  even  distinguished  the  NACA 
from  the  army  (and  later  the  air  force)  at  Langley  Field.  Langley  was 
all  about  flying  and  noisy  airplanes  that  woke  residents  before  their  alarm 
clocks  went  off.  But  the  people  at  the  NACA  were  not  concerned  about 
the  confusion.  Being  grouped  with  the  soldiers  in  uniform  was  often  useful 
camouflage.  This  camouflage  was  especially  helpful  during  World  War  II 
when  hard  feelings  were  expressed  by  local  families  who  saw  their  boys 
going  off  to  war  while  NACA  men  were  able  to  stay  put  because  of  a  special 
deal  made  between  the  NACA  and  the  Selective  Service  System.38 

In  1958  the  natives  still  poked  fun  at  the  NACA  Nuts,  but  they  did  so  in  a 
more  friendly  way.  Previously,  a  friction  similar  to  that  felt  typically  between 
university  "town  and  gown"  had  determined  much  about  the  Hampton- 
Langley  relationship.  The  softening  of  hard  feelings  between  locals  and  the 
NACA  was  due  in  large  part  to  the  marriage  of  many  Langley  engineers 
to  area  women  and  their  subsequent  assimilation  into  local  society.  For 
instance,  the  wife  of  Langley's  number  two  man  in  1958,  Associate  Director 
Floyd  Thompson,  was  Jean  Geggie,  a  native  Hamptonian  whose  father 
carved  wooden  figureheads  for  ships  at  the  nearby  Newport  News  shipyard. 

By  the  1950s,  NACA  employees  had  become  pillars  of  the  community. 
Thompson  himself  had  been  a  member  of  the  Hampton  Rotary  Club  for 
several  years  and  had  served  on  the  board  of  directors  of  the  local  "Dixie 
General"  hospital.  (In  the  late  1960s,  partly  through  Thompson's  efforts, 
the  hospital  board  voted  to  drop  the  racially  inflammatory  name  "Dixie" 
and  renamed  the  hospital  Hampton  General.)  Furthermore,  in  the  turbulent 
and  scary  weeks  following  the  first  Soviet  space  launches,  the  scientists  and 
engineers  "over  at  Langley  Field"  became  reassuring  figures.  Here,  right  in 
their  midst,  many  locals  felt,  were  experts  who  could  explain  the  meaning  of 

24 


The  Metamorphosis 


L-78,005 

This  1950  aerial  photo  of  Langley  shows  the  original  East  Area  along  the  Back  River 
(bottom)  and  the  West  Area,  constructed  during  World  War  II  (top). 


the  foreign  objects  orbiting  ominously  overhead.  Interviewed  for  stories  by 
the  local  newspapers,  NACA  personnel  discussed  the  progress  of  American 
space  efforts  and  helped  calm  local  hysteria.  Hamptonians  developed  greater 
appreciation  for  the  technical  talents  of  Langley  personnel,  and  the  once 
tepid  feelings  about  the  NACA  warmed. 

With  the  transition  to  NASA,  the  public  spotlight  would  inevitably  shine 
on  Langley.  Personnel  would  soon  figure  out  that  the  NACA  attitude  toward 
public  relations  had  to  change.  In  the  old  days,  most  NACA  staff  members 
could  have  cared  less  about  public  opinion.  They  only  cared  about  the 
opinion  of  generals,  congressmen,  and  other  powerful  people  who  could 
influence  the  budget  and  appropriation  processes.  With  NASA,  however, 
things  had  to  be  different.  Beginning  the  day  after  the  launch  of  Sputnik  1, 
researchers  had  to  make  their  case  before  a  much  more  concerned  public. 
Without  hesitating,  they  got  right  to  it. 


25 


The  First  NASA  Inspection 


It  was,    by  all  odds,    a  superlative  display. 
Our  sincere  thanks  for  a  superbly  designed,  brilliantly 
mounted,  and  perceptive  look  at  the  very  general  goals 
man  must  achieve  before  he  becomes  a  space  traveler. 

— Editorial,  Newport  News  Daily  Press 
27  October  1959 

On  Saturday  morning,  24  October  1959,  a  little  more  than  a  year 
after  the  metamorphosis  of  the  NACA  into  NASA,  approximately  20,000 
visitors  marched  through  the  gates  of  Langley  Field  to  attend  a  public  open 
house  that  was  being  held  in  conjunction  with  NASA's  First  Anniversary 
Inspection.  The  NACA's  first  anniversary  had  passed  unnoticed;  NASA's 
proved  to  be  a  controlled  mob  scene. 1 

The  crowds  came  at  NASA's  invitation.  Local  newspapers  and  commu- 
nity groups  had  spread  the  word:  for  the  first  time  in  its  42-year  history, 
Langley  Research  Center  was  admitting  curious  outsiders  into  the  previously 
sheltered  sanctuary  of  aeronautical  research.  NASA  scientists,  engineers, 
and  technicians  would  show  the  public  just  what  the  new  space  agency  had 
been  doing  to  launch  their  country  into  space.  Throughout  the  day,  men, 
women,  and  children  streamed  through  the  huge  NASA  aircraft  hangar  as 
well  as  through  two  other  large  buildings  full  of  exhibits  that  represented  a 
cross  section  of  NASA  programs.  Escorting  the  visitors  was  a  handpicked 
group  of  articulate  and  polite  NASA  employees  whose  job  was  to  handle  the 
pedestrian  traffic,  guide  the  visitors  through  the  buildings  included  in  the 
program,  and  explain  the  exhibits. 

The  visitors  moved  "in  fascination"  past  the  many  marvels  on  display.2 
They  saw  helicopters  and  aircraft,  including  a  Chance  Vought  F8U-3  navy 
supersonic  jet  fighter  used  by  NASA  for  sonic-boom  research  over  Wallops 
Island;  a  Vertol  76,  the  world's  first  tilt-wing  aircraft;  a  ground-effect  vehicle 
designed  to  move  over  a  cushion  of  air  that  the  unusual  craft  created  between 
its  base  and  the  ground;  a  display  about  the  possibilities  of  SST  flight 


27 


Space/light  Revolution 

(subsonic  commercial  jet  flights  across  the  Atlantic  had  only  been  made 
for  about  a  year);  a  full-size  mock-up  of  the  air  force/NASA  X-15  rocket- 
powered  research  airplane;  plus  dozens  of  static  and  dynamic  demonstrations 
involving  wind  tunnels,  electrically  powered  models,  electromagnetism, 
research  instrumentation,  as  well  as  several  examples  of  NASA  technical 
reports. 

Towering  above  all  and  attracting  the  most  attention  was  a  large  fleet  of 
space  vehicles  and  rockets.  This  collection  included  a  model  of  the  original 
German  V-2  rocket  engine;  a  full-size  version  of  the  Thor-Able  missile,  which 
had  been  used  to  launch  a  number  of  U.S.  space  probes;  a  19-foot  Discoverer 
satellite  to  be  used  in  polar-orbit  research;  a  full-scale  Little  Joe  rocket  that 
was  part  of  the  Mercury  program;  a  72-foot  Scout  rocket  to  be  used  for 
general  space  research  purposes;  a  six-stage  rocket  vehicle  used  for  reentry 
physics  studies  at  Wallops  Island;  and  a  6-foot  model  of  the  world  with 
orbital  traces  of  the  major  satellites  launched  by  the  United  States. 

The  public  was  so  eager  to  see  these  wonders  of  modern  technology  that 
visitors  had  started  forming  lines  around  the  exhibits  as  early  as  8:00  a.m. 
even  though  the  program  was  not  scheduled  to  begin  until  10:00  a.m.,  and 
they  continued  to  swarm  around  the  exhibits  throughout  the  day.  Most 
of  the  visitors  were  residents  of  the  Peninsula  area,  but  the  license  plates 
on  some  of  the  cars  indicated  that  several  had  traveled  from  more  remote 
parts  of  Virginia  and  a  few  had  come  from  as  far  away  as  Georgia  and 
Tennessee.3  For  the  NASA  Langley  staff,  "The  Nice  NASA  Show  For 
The  People,"  as  one  local  editor  called  it,  was  quite  an  eye-opener.  No 
one  expected  the  general  public  to  be  so  curious  about  NASA's  research 
programs.4 

After  World  War  II,  family  members  and  friends  of  Langley  personnel 
had  been  welcome  on  occasion  to  attend  briefings  and  watch  demonstrations 
"boiled  down"  from  recently  concluded  NACA  inspections  (annual  confer- 
ences for  aeronautical  insiders  only).  Never  before  the  1959  inspection, 
however,  had  Langley  put  on  an  open  house  involving  more  than  just  the 
center's  employees  and  their  families.  Langley  had  neither  a  visitors'  center 
(until  1971)  nor  any  other  regular  means  to  handle  many  outsiders;  none 
was  necessary  given  the  NACA's  low  profile  and  the  limited  public  interest 
in  what  was  going  on  inside  a  place  that  some  locals  referred  to  as  "Sleepy 
Hollow." 

The  unprecedented  public  open  house  came  at  the  end  of  a  week-long 
closed  affair  modeled  after  the  old  NACA  annual  inspections.  Up  to  400 
people  a  day  had  attended  these  NACA  conferences.  Although  they  came 
by  direct  invitation  to  learn  about  NACA  programs,  most  guests  already 
knew  quite  a  bit  about  these  programs  because  conference  attendees  were 
the  patrons  and  clients  of  the  NACA.  Representatives  from  military  aviation, 
the  aircraft  industry,  and  the  airlines,  and  a  few  people  from  government 


28 


The  First  NASA  Inspection 


L-59-8075  L-59-8097 

A  mock-up  of  the  Mercury  space  capsule  appears  to  land  by  parachute  on  Langley's 

"Mercury  Support"  exhibit  at  the  October  1959  event  (top).  At  bottom  left, 
Langley  Director  Henry  Reid  (middle),  former  Langley  researcher  and  soon-to-be- 
named  head  of  the  new  Office  of  Advanced  Research  Programs  (OARP)  at  NASA 
headquarters  Ira  H.  Abbott  (right),  and  an  unidentified  guest  stare  up  at  the  capsule 
mounted  atop  a  model  of  the  Atlas  booster  rocket.  At  bottom  right,  Langley 
Associate  Director  Floyd  L.  Thompson  (middle),  with  Coke  bottle  in  hand,  and 
NASA  Administrator  Glennan  (left)  chat  with  guests  and  associates  in  front  of  the 
globe  showing  the  orbital  traces  of  previously  launched  American  satellites. 

29 


Space/light  Revolution 

and  the  trade  journal  media  had  been  the  only  visitors  invited  to  the  NACA 
inspections.* 

No  one  at  NASA  headquarters  had  been  sure  whether  to  continue  the 
tradition  of  the  NACA  inspection,  which  by  the  1950s  was  rotating  annually 
among  Langley,  Lewis,  and  Ames.  The  inspection  was  such  a  long-running 
show,  having  premiered  at  Langley  in  1926,  and  its  actors,  settings,  and 
stage  directions  were  so  closely  identified  with  the  NACA  that  some  NASA 
officials  wondered  whether  the  event  would  serve  the  interests  of  NASA's  new 
mission.  But  in  the  opinion  of  many  others,  including  Dr.  Hugh  Dryden, 
NASA's  deputy  administrator,  the  inspection  offered  NASA  an  excellent 
means  of  publicizing  what  it  had  accomplished  during  its  first  year  to  achieve 
the  nation's  new  objectives  in  aeronautics  and  space.  "Prom  a  publicity 
point  of  view,"  read  one  NASA  Langley  document  that  outlined  the  general 
purpose  of  the  proposed  inspection,  "the  exhibits  will  present  to  the  audience 
not  only  our  aims  and  objectives,  but  the  research  background  that  led  to 
the  'present-day'  and  future  space  developments."  In  other  words,  NASA 
could  make  the  point,  both  directly  and  indirectly,  that  "pioneering  'in- 
house'  research  is  a  first  prerequisite  to  successful  aeronautic  and  space 
developments."5 

Although  this  emphasis  on  in-house  capabilities  did  not  match 
Keith  Glennan's  agenda  for  NASA  (Glennan  wanted  to  see  more  research 
being  done  by  outside  contractors),  the  overall  objective  of  the  plan  per- 
suaded the  administrator.  He  decided  that,  in  October  1959,  NASA  would 
hold  its  First  Anniversary  Inspection,  a  sort  of  public  show-and-tell  event. 

Because  NASA  was  a  new  agency  with  different  objectives  and  a  much 
wider  scope  than  its  predecessor,  a  few  things  about  the  inspection  were 
to  be  done  differently.  Not  only  was  NASA  to  have  an  open  house  for 
the  general  public,  it  must  also  invite  several  foreign  guests.  While  the 
NACA  had  discouraged  their  attendance,  NASA  had  vested  programmatic 
interests  in  (and  mandated  legal  obligations  to)  foreign  nations,  which  meant 
that  some  foreign  scientists,  diplomatic  representatives,  and  members  of  the 
foreign  press  corps  had  to  be  invited  to  attend.  At  NASA  headquarters, 
the  Office  of  International  Programs,  under  Henry  E.  Billingsley,  and  the 
Office  of  Space  Flight  Development,  under  Abe  Silverstein,  were  in  charge 
of  issuing  these  invitations. 

Although  NASA  had  to  aggressively  pitch  its  program  to  the  taxpayers, 
which  meant  packaging  it  as  attractively  as  possible,  the  1959  inspection 
was  virtually  the  same  ritual  that  the  NACA  had  always  orchestrated  for 
the  visitors.  After  registering  at  the  base  gymnasium  starting  at  8:00  a.m., 
the  guests  moved  to  an  introductory  session  in  the  base  theater  from  8:50  to 


Some  headquarters  officials  did  not  like  the  name  "inspection,"  which  had  been  in  use  since  the 
1940s.  They  argued  that  it  did  not  accurately  convey  what  happened  in  the  program.  They  suggested 
"exhibition,"  "observance,"  "annual  meeting,"  and  a  number  of  other  substitutes,  but  none  of  these 
names  was  adopted. 

30 


The  First  NASA  Inspection 


L-59-8081 

Administrator  Glennan  spends  a  few  minutes  in  front  of  the  Mercury  capsule  exhibit 
with  Walter  Bonney,  NASA 's  first  director  of  the  office  of  public  information. 
Bonney,  who  had  worked  for  the  NACA  from  1949  to  1958,  never  found  much 
favor  in  Glennan's  employ.  Glennan  criticized  Bonney  harshly  for  his  outdated, 
NA  CA  approach  to  the  public  information  field. 

9:00  a.m.  and  from  there  went  to  a  brief  technical  program  in  the  cavernous 
test  section  of  the  Full-Scale  Tunnel.  Pinned  to  the  coat  of  every  guest  was 
an  identification  badge  with  the  person's  name  and  tour  group. 

For  the  extended  tour  of  the  laboratory,  Langley  continued  the  old 
NACA  practice  of  dividing  the  guests  into  color-coded  groups,  in  this 
case  into  10  groups  of  no  more  than  40  persons  each.  Each  group  had 
its  own  bus  with  a  color-coded  sign  in  the  window,  its  own  escorts  and 
attaches,  its  own  schedule  to  keep,  and,  at  least  in  the  minds  of  the 
inspection  organizers,  its  own  personality.  NASA  management  wanted  a 
mix  of  people  in  every  group,  but  it  also  wanted  the  group  members  to  be 
compatible.  As  expected,  the  gold  group  included  dignitaries  and  VIPs. 
The  brown  and  tan  groups  had  the  majority  of  the  journalists,  and  the 
pink  group  included  the  few  women  who  were  invited.  The  red  group 
comprised  most  of  NASA's  leaders.  On  the  first  day  of  the  inspection, 
Tuesday,  20  October,  Langley  hypersonics  specialist  John  V.  Becker  was 
the  guide  for  the  red  group,  which  included  Robert  R.  Gilruth,  head  of 
the  new  Space  Task  Group  (STG);  NASA  Administrator  Glennan;  NASA 
Deputy  Administrator  Dryden;  NASA  Executive  Secretary  John  F.  Victory; 
NASA  Goddard  Director  Harry  J.  Goett;  NASA  Ames  Director  Smith  J. 
DeFrance;  NASA  Flight  Research  Center  Chief  Paul  F.  Bikle;  Wallops 

31 


Spaceflight  Revolution 


L-59-7828 

Four  Langley  secretaries  serving  as  hostesses  for  the  inspection  take  a  look  inside 
the  cockpit  of  a  full-size  mock-up  of  the  X-15. 

Station  Engineer-in-Charge  Robert  L.  Krieger;  plus  several  lesser  officials 
from  NASA  headquarters.  Also  in  the  group  were  a  few  important  men 
from  the  aerospace  industry,  the  airlines,  and  the  armed  forces.6 

Although  some  NASA  personnel  came  to  the  inspection  as  guests,  most 
came  to  Langley  to  report  on  the  progress  of  the  work  at  their  respective 
centers.  NASA  Lewis  sent  an  exhibit  that  demonstrated  the  relative 
merits  of  low-thrust  space  propulsion  systems,  including  chemical,  nuclear- 
hydrogen,  and  electrical  rockets.  NASA  Ames  contributed  a  display  showing 
the  physics  of  high-velocity  impact  in  space  and  the  potential  dangers  of 
meteoroid  collision  with  spacecraft.  For  its  part,  the  NASA  Flight  Research 
Center  at  Edwards  AFB  had  contracted  with  North  American  Aviation 
for  a  mock-up  of  the  X-15  and  of  the  XLR-99  rocket  engine  along  with 
a  dummy  pilot  dressed  in  a  pressure  suit.  The  Jet  Propulsion  Laboratory 
(JPL)  in  Pasadena,  California,  formerly  operated  by  the  California  Institute 
of  Technology,  had  transferred  to  NASA  in  December  1958.  The  laboratory 
sent  a  small  display  and  a  team  of  scientists  to  present  the  story  of  the  Vega 
rocket;  at  the  time  of  the  inspection,  NASA  thought  that  this  three-stage 
booster  would  take  a  number  of  future  vehicles  and  payloads  into  space,  even 
into  lunar  orbit,  but  the  proposed  $65-million  development  program  would 
be  cancelled  only  two  months  after  the  inspection.  The  new  Goddard  Space 
Flight  Center  was  still  a  part  of  the  Naval  Research  Laboratory  (NRL)  at  its 
Anacostia  location  pending  construction  of  an  independent  NASA  facility 

32 


The  First  NASA  Inspection 

at  Greenbelt,  Maryland.  Goddard  contributed  a  display  featuring  several 
examples  of  lightweight  inflatable  structures  that  had  applications  for  use 
in  satellites  and  spaceflight.7 

As  was  becoming  to  the  host  center,  NASA  Langley  presented  by  far 
the  greatest  number  and  variety  of  exhibits.  Langley  staff  built  displays 
and  gave  illustrated  talks  on  many  space  subjects:  the  nature  of  the 
space  environment,  reentry  physics,  and  manned  reentry  vehicles  such  as 
ballistic  capsules,  high-drag  gliders,  and  high  lift-drag  boost-gliders.  Langley 
engineers  also  reported  on  aeronautical  programs,  notably  the  X-15,  Vertol 
76,  and  an  SST  airplane.  Langley  even  supplemented  Ames's  display  of 
high- velocity  impacts  in  space  with  graphic  results  of  its  own  experiments 
on  the  subject. 


Following  the  NACA  Way 

According  to  the  NACA's  policy  of  triennial  rotation  among  its  three 
major  research  centers,  it  was  "by  the  numbers"  Langley's  turn  to  host  the 
1959  inspection.  However,  NASA  probably  would  have  held  the  inspection 
there  regardless  of  the  rotation.  The  assistant  chief  of  the  Full-Scale 
Research  Division  and  Langley's  coordinator  for  the  technical  program,  Axel 
Mattson,  remembers  with  pride: 

There  was  only  one  place  that  could  put  on  that  show.  .  .  .  There  was  no  other  place 
for  it  to  go.    .    .    .   If  it  had  been  someplace  else,  the  overall  presentation  wouldn't 

o 

have  been  as  good,  and  the  emphasis  might  have  been  slightly  different. 

In  other  words,  Langley  had  the  most  experience  in  staging  this  event. 
Langley  was  also  the  oldest  NACA  facility  and  the  NASA  center  closest 
to  Washington,  D.C.,  thus  making  it  convenient  to  congressional  and  other 
powerful  visitors.  Perhaps  most  importantly,  Langley  was  the  place  where 
the  stars  of  the  space  program — the  STG  and  its  astronauts — were  in 
training  for  the  first  U.S.  manned  space  effort,  Project  Mercury. 

Axel  Mattson  was  a  big,  likeable,  and  loquacious  engineer  who  loved  the 
showmanship  and  conviviality  of  past  inspections.  In  the  weeks  prior  to 
the  1959  event,  his  job  was  to  confer  with  the  other  NASA  centers  and  to 
help  them  plan  their  participation  in  the  inspection.  In  the  cases  of  Ames, 
Lewis,  Wallops,  and  the  Flight  Research  Center  at  Muroc,  Mattson's  help 
was  only  minimal  because  the  staffs  at  the  former  NACA  facilities  knew 
what  an  inspection  demanded.  They  understood  the  rigorous  standards  for 
quality  presentations  and  were  ready  for  the  customary  competition  among 
the  centers  for  the  best  exhibits.  All  of  the  centers  "tried  to  out-do  one 
another"  with  the  most  sophisticated  displays  and  demonstrations,  Mattson 
recalls.  "At  least  we  thought  they  were  sophisticated,  let's  put  it  that  way."£ 

33 


Spaceflight  Revolution 

The  1959  Anniversary  Inspection  was  the  first  time  that  all  the  NASA 
facilities  were  participating,  and  those  facilities  included  two  that  had  not 
been  part  of  the  NACA — JPL  and  Goddard.*  Mattson  was  responsible 
for  encouraging  the  staffs  of  these  new  centers  to  develop  appropriate 
and  effective  presentations  for  the  inspection.  "I  had  a  dog  and  pony 
show,"  Mattson  remembers.  "I  took  slides  with  me  from  previous  NACA 
conferences"  to  show  them  what  went  on.  He  assembled  the  initiates  in 
a  conference  room,  making  sure  that  people  "with  enough  horsepower"  to 
make  the  right  things  happen  were  in  the  audience,  and  then  he  briefed 
them  on  what  an  inspection  was  about  and  the  purposes  it  served.10 

Mattson  tried  his  best  to  be  polite  and  not  to  act  arrogant  while 
educating  the  non-NACA  staffs  about  the  do's  and  don'ts  of  an  inspection, 
but  he  still  did  not  receive  a  warm  welcome  at  either  of  the  two  non-NACA 
centers.  In  fact,  at  Goddard's  temporary  home  within  the  NRL,  he  feared 
he  would  "be  tarred  and  feathered."  Typically,  any  organization  that  had 
been  "navy"  had  superb  loyalty  among  its  staff  and  was  very  closed,  even 
resentful  of  outsiders.  In  the  opinion  of  the  Goddard  staff  members,  the 
inspection  "was  just  something  that  the  NACA  did,  and  they  didn't  think 
much  of  it."11 

In  particular,  the  navy  personnel  did  not  like  the  idea  of  rehearsals.  In 
advance  of  NACA  inspections,  staff  members  customarily  rehearsed  their 
talks  in  their  own  research  divisions  and  then  sweated  through  another 
performance  a  week  or  so  before  the  event  as  part  of  a  fully  staged  dress 
rehearsal  with  center  management  and  several  key  officials  from  NACA 
headquarters  as  the  audience.  For  all  the  Washington  office  people  to  come 
down  to  Langley  and  critique  the  inspection  material  was  a  "big  thing." 
Dr.  Dryden,  John  Victory,  and  others  "all  had  a  grand  time  with  that." 
Some  laboratory  employees  complained  privately  about  "having  to  put  on 
a  parade  for  their  parents,"  but  most  had  reconciled  themselves  to  the 
imposition.  By  1959,  NACA  veterans  like  Mattson  saw  the  NACA  practice 
of  rehearsals  as  the  only  way  to  guarantee  the  success  of  such  a  complex 
show.  Mattson  had  to  convince  NASA's  new  partners  of  the  importance  of 
all  the  planning  and  preparations.  The  staff  at  Goddard  was  unimpressed  by 
Mattson 's  explanations.  A  few  of  the  more  indignant  told  Mattson:  "You 
won't  rehearse  me.  My  gosh,  I'm  an  expert,  you  know.  Who's  going  to 
critique  what  I  say?"  But  Mattson  held  his  ground  and  told  them  they 


sk 

The  ABMA  (Army  Ballistic  Missile  Agency)  under  Dr.  Wernher  von  Braun  at  the  Redstone  Arsenal 

in  Huntsville,  Alabama,  did  not  become  a  part  of  NASA  until  their  "shotgun  marriage"  was  consummated 
by  a  vote  of  Congress  in  February  1960,  but  the  decision  to  transfer  the  ABMA  to  NASA  was  actually 
finalized  in  October  1959,  the  month  of  the  first  NASA  inspection.  A  number  of  ABMA  representatives 
attended  the  NASA  inspection.  So,  too,  did  the  mayor  of  Huntsville. 


34 


The  First  NASA  Inspection 


L-59-7831 

In  this  picture  from  the  1959  inspection,  Axel  Mattson  (right)  confers  with  John 
Stack,  a  devoted  airplane  man  who  surely  experienced  mixed  feelings  about  the  affair 
because  of  its  emphasis  on  space  rather  than  aeronautics. 

would  have  to  do  it.  Thinking  back,  Mattson  calls  his  visits  to  the  non- 
NACA  installations  "interesting  sessions,"  and  he  singles  out  the  first  NASA 
inspection  as  "the  most  difficult  inspection  of  them  all  to  put  together." 

Other  NACA  veterans  have  also  commented  on  the  difficulties  of  the  new 
fraternal  relationships  within  NASA.  "There  wasn't  any  love  lost  between 
us,"  remembers  Langley's  Charles  J.  Donlan.  "I  really  shouldn't  say  'love 
lost'  because  the  people  really  didn't  know  one  another."  But  "all  the  NRL 
guys"  came  "kicking  and  screaming  into  this  new  organization"  that  they 
thought  was  "going  to  be  overwhelmed  by  the  NACA  bunch."13  Everyone 
needed  time  to  get  over  these  psychological  barriers  and  realize  that  they 
were  all  working  as  a  team.  A  few  people,  some  say  particularly  at  Goddard, 
were  never  able  to  accept  the  partnership. 

Strained  interaction  among  NASA  centers  represents  a  key  tension  in  the 
story  of  NASA  that  historians  have  not  explored  fully.  In  the  first  NASA 
inspection,  a  vestige  of  the  old  NACA  culture  won  out  over  other  integral 
parts  of  NASA;  in  the  ensuing  years,  the  culture  of  the  NACA  research 
laboratories,  dominant  in  the  early  years  of  NASA,  would  in  many  ways 
be  overwhelmed  and  superseded  by  those  at  the  more  hardware-oriented 
and  operations-oriented  spaceflight  and  spacecraft  centers  in  Huntsville, 
Houston,  and  at  Cape  Canaveral.  This  turnabout,  which  would  have  seemed 
unlikely  in  the  earliest  days  of  NASA,  was  made  inevitable  by  the  large 
manned  spaceflight  programs  of  the  1960s  and  1970s.  The  biggest  bucks 
would  be  spent  on  the  more  industrial  side  of  NASA,  as  they  still  are. 

In  the  end,  everyone  at  Goddard  and  JPL  agreed  to  do  their  part  in  the 
1959  inspection.  As  mentioned  earlier,  Goddard  staff  sent  an  exhibit  that 
featured  four  erectable  space  structures,  but  they  did  so  only  after  Langley 

35 


Spaceflight  Revolution 


L-59-7275 

The  small  exhibit  area  on  space  science  and  technology  provided  by  the  Goddard 
Space  Flight  Center. 


had  proposed  that  Goddard  send  an  exhibit  dealing  with  reentry  physics. 
The  JPL  group  sent  an  exhibit  about  the  soon-to-be-cancelled  Vega  project. 
Both  exhibits  were  prepared  with  the  help  of  outside  design  consultants. 
The  NASA  representatives  sent  to  Langley  with  those  exhibits  were  "awful 
proud"  of  what  they  had  done.  "After  all  the  trials  and  tribulations  of 
getting  them  organized  and  getting  them  going,"  Mattson  states,  "they 
walked  around  like  peacocks"  strutting  their  stuff  and  showing  off  their 
exhibits.14 

Interestingly,  after  getting  the  new  centers  to  cooperate  and  to  do  it  the 
NACA  way,  some  NACA  veterans  still  found  reasons  to  criticize.  "For  my 
money,"  Smith  J.  DeFrance,  the  director  of  Ames,  wrote  to  Henry  Reid,  the 
director  of  Langley: 

the  stops  [on  the  tour]  by  your  group  were  far  superior  to  the  Jet  Propulsion 
Laboratory's  stop  and  especially  the  Goddard  Space  Flight  Center's  stop.  As  you 
know,  both  of  these  were  prepared  by  so-called  specialists  in  the  field  of  exhibition. 
Neither  of  the  stops  came  up  to  the  degree  of  perfection  that  was  demonstrated  by 


your  own  people. 


15 


DeFrance  had  come  to  work  at  NACA  Langley  in  1922;  Reid  had  come 
in  1921.   They  had  followed  the  NACA  way  for  so  long  that  they  found  it 


36 


The  First  NASA  Inspection 

difficult  to  value  any  other.  But  Reid's  answer  did  reflect  an  openness  to  the 
new  NASA  partnerships.  "Letters  are  pouring  in  from  many  of  the  visitors," 
he  wrote  DePrance,  "and  I  feel  that  this  inspection  has  certainly  been  very 
much  worthwhile,  not  only  because  of  the  impression  made  on  people  outside 
our  organization  but  also  the  impression  made  on  many  of  our  new  members 
of  the  organization."  Despite  the  problems  convincing  new  members  of  the 
importance  of  an  inspection,  Reid  summed  up  the  experience  as  positive: 
"We  were  indeed  very  fortunate  in  having  the  excellent  teamwork,  even  from 
our  new  organization,  JPL."  The  teamwork  of  Goddard,  to  the  extent  that 
it  materialized,  Reid  did  not  mention.16 


Project  Mercury 

"Ladies  and  gentlemen,  at  this  stop  we  shall  discuss  Project  Mercury," 
announced  the  NASA  engineer  as  another  busload  of  visitors  to  the  1959 
inspection  found  their  way  to  the  cold  metal  folding  chairs  set  up  in  rows 
inside  the  West  Area's  Aircraft  Loads  Calibration  building.  Eight  young 
members  of  the  STG  working  in  teams  of  two  took  turns  giving  this  talk. 
The  script  of  the  presentation  had  been  finalized  just  a  day  or  two  before 
the  inspection  to  ensure  an  up-to-date  report. 

The  STG  speakers  did  not  bother  to  introduce  themselves  (they  had  been 
told  not  to),  and  their  identities  would  not  have  meant  much  to  most  people 
in  the  audience.  They  were  Edison  M.  Fields  and  Jerome  Hammack,  Systems 
Test  Branch;  Elmer  A.  Horton,  Control  Central  and  Flight  Safety  Section; 
Milton  B.  Windier,  Recovery  Operations  Branch;  John  D.  Hodge,  Opera- 
tions Division;  Carl  R.  Huss,  Trajectory  Analysis  Section;  John  E.  Gilkey, 
Engineering  Branch;  and  Norman  F.  Smith,  Engineering  and  Contract  Ad- 
ministration. As  it  turned  out,  some  of  these  men  were  destined  to  play 
major  roles  in  NASA's  subsequent  manned  space  programs.17 

"The  possibility  of  venturing  into  space,"  the  inspection  talk  began,  "has 
shifted  quite  recently  from  the  fantasy  of  science  fiction  to  the  realm  of 
actuality.  Today,  space  flight  is  considered  well  within  the  range  of  man's 
capabilities."  Only  five  days  after  its  establishment,  NASA  had  formed 
the  STG  to  design  and  implement,  as  quickly  as  possible,  a  manned  satellite 
project.  NASA  put  veteran  NACA  researcher  Robert  R.  Gilruth,  the  former 
head  of  Langley's  Pilotless  Aircraft  Research  Division  (PARD),  in  charge; 
based  the  group  at  Langley;  and  named  the  Project  Mercury  after  the  fleet- 
footed  Roman  god  of  commerce,  who  served  as  messenger  of  the  gods.18  The 
speakers  proudly  declared  the  mission  of  Project  Mercury:  to  send  "this 
nation's  first  space  traveler  into  orbit  about  the  earth,"  to  study  "man's 


37 


Space/light  Revolution 


NACA  veteran  Robert  R.  Gilruth  di- 
rected Project  Mercury  from  offices  at 
Langley. 


L-59-57 


capabilities  in  space  flight,"  and  to  assure  "the  safe  return  of  the  capsule 
and  its  pilot  to  the  earth."19 

The  STG  plan  was  to  send  a  small  one-person  spacecraft  into  orbit  using 
the  existing  Atlas  intercontinental  ballistic  missile  as  the  launch  vehicle  and 
a  ballistic  reentry  module  as  the  crew  capsule.  After  a  few  passes  around  the 
earth,  retrorockets  would  fire  to  slow  the  satellite  and  thus  initiate  descent 
from  orbit.  After  reentry  into  the  atmosphere — accomplished  safely  thanks 
to  the  capsule's  blunt  ablative  heat  shield — a  large  parachute  would  deploy 
to  carry  the  capsule  on  its  final  approach  and  land  it  in  the  open  sea.  The 
capsule  and  the  astronaut  would  be  recovered  by  helicopter  and  brought 
home  aboard  a  naval  vessel. 

The  Mercury  plan  was  a  bold  yet  essentially  conservative  engineering 
concept,  and  it  was  to  be  almost  unbelievably  successful.  By  May  1963,  it 
resulted  in  the  successful  launches  of  six  Americans  into  space,  thus  leading 
to  some  two  and  one-half  days  of  flight  time  in  space.  Although  glitches 
and  other  vexing  technical  problems  would  plague  virtually  every  Mercury 
mission,  no  major  accidents  occurred.  "We  were  pretty  lucky,"  one  leader  of 
Project  Mercury  remembers.  "In  retrospect,  we  wouldn't  dare  do  it  again 
under  the  same  circumstances.  But  that's  true  of  most  pioneering  ventures. 
You  wouldn't  dare  fly  across  the  ocean  with  one  engine  like  Lindbergh  did, 
either,  would  you?"20 


38 


The  First  NASA  Inspection 

DEVELOPMENTS  FOR  MANNED  FLIGHT  IN  SPACE 

ISO  MILE  ORBIT 


L-59-1167 

A  diagram  used  at  the  first  NASA  inspection  to  illustrate  the  basic  concept  of  a 
Mercury  man-in- space  mission. 

Without  question,  the  Project  Mercury  stop  was  the  featured  attraction 
of  NASA's  entire  anniversary  show.  In  1959  everyone  around  the  country 
was  obsessed  with  beating  the  Soviets  to  manned  spaceflight,  and  that 
obsession  soon  included  the  men  who  would  actually  pilot  the  spacecraft. 
Introduced  to  the  public  for  the  first  time  in  April  1959,  NASA's  astronauts 
were  not  yet  the  golden  boys  they  eventually  became,  but  with  the  national 
media  already  bearing  down  on  them  and  NASA's  public  affairs  officers 
polishing  the  seven  former  test  pilots'  armor  to  a  blinding  shimmer,  the 
future  knights  of  spaceflight  had  already  acquired  star  quality.  They  were 
national  heroes  before  they  did  anything  heroic.  Some  of  their  luster  was  lost 
in  August  1959,  if  only  temporarily,  when  the  astronauts  sold  the  exclusive 
rights  to  their  personal  stories  to  Time-Life  for  one-half  million  dollars.  To 
most  Americans  this  seemed  an  excessive  amount  of  money;  at  that  time  the 
federal  minimum  wage  was  a  mere  $1  an  hour.  The  resulting  controversy 
over  the  ethics  of  the  deal  was  fueled  largely  by  Life's  legitimately  disgruntled 
competition  and  did  not  really  do  much  to  damage  the  public's  growing  love 
affair  with  their  handsome,  if  not  yet  "launched,"  astronauts.21 

A  few  minutes  into  their  talk  at  the  Project  Mercury  stop,  the  STG 
speakers  dimmed  the  lights  and  showed  a  short  motion  picture  devoted 
to  "the  seven  brave  young  men  who  have  been  chosen  as  the  Mercury 
astronauts."22  First  as  a  group,  then  one  by  one,  the  film  introduced  them, 


39 


Spaceflight  Revolution 


L-71-2971 

The  "Original  Seven":  (left  to  right)  Carpenter,  Cooper,  Glenn,  Grissom,  Schirra, 
Shepard,  and  Slayton. 


just  as  each  had  been  introduced  with  such  flair  during  the  sensational  open- 
ing press  conference  at  NASA  headquarters  on  9  April  1959.  The  "Original 
Seven"  were  Air  Force  Capts.  Leroy  G.  Cooper,  Jr.  (later  called  L.  Gordon), 
Virgil  I.  "Gus"  Grissom,  and  Donald  K.  "Deke"  Slayton;  naval  aviators 
Lt.  Malcolm  S.  Carpenter  (who  preferred  "M.  Scott"),  Lt.  Comdr.  Alan  B. 
Shepard,  Jr.,  and  Lt.  Comdr.  Walter  M.  Schirra,  Jr.;  and  Lt.  Col.  John  H. 
Glenn,  Jr.,  of  the  Marine  Corps.  Everyone  knew  that  one  of  these  men 
would  soon  be  the  first  American,  possibly  the  first  human,  to  venture  into 
space;  one  of  the  seven  was  destined  to  become  the  greatest  technological 
hero  since  Lindbergh. 

The  Mercury  astronauts  were  the  survivors  of  an  extraordinarily  elab- 
orate and  rigorous  search  process  that  the  STG  had  used  to  solicit  appli- 
cations from  and  to  evaluate  candidate  astronauts.  At  the  start  nobody 
knew  what  sort  or  degree  of  skill,  education,  and  training  space  pilots  would 
need.  So-called  specialists  in  crew  selection  proposed  that  NASA  choose  the 
astronauts  from  "people  in  dangerous  professions,  such  as  race  car  drivers, 
mountain  climbers,  scuba  divers,  as  well  as  test  pilots."  But  the  STG  was 
committed  to  the  idea  of  test  pilots  from  the  beginning;  with  just  any  old 
breed  of  daredevil  on  board,  the  delegation  of  critical  flight  control  and 


40 


The  First  NASA  Inspection 

command  functions  to  the  crew  in  the  capsule  would  be  much  more  diffi- 
cult. When  President  Eisenhower  decided  that  astronauts  would  be  chosen 
from  a  military  test-pilot  pool,  Gilruth  and  associates  all  "breathed  a  sigh 
of  relief."23 

A  key  person  in  the  screening  and  final  selection  of  the  Mercury  astro- 
nauts was  Langley's  Charles  J.  Donlan.  Formerly  the  free-lance  technical 
assistant  to  Floyd  Thompson,  Donlan  was  now  serving  as  Gilruth's  deputy. 
Working  on  a  crash  schedule  basis,  Donlan  headed  the  NASA/Department 
of  Defense  (DOD)  team,  which  included  a  psychologist  on  loan  to  NASA 
from  the  National  Science  Foundation.  The  team  established  the  final  seven 
evaluation  criteria: 

1.  Less  than  40  years  old 

2.  Less  than  5' 11"  tall* 

3.  Excellent  physical  condition 

4.  Bachelor's  degree  in  engineering  or  equivalent 

5.  Test-pilot  school  graduate 

6.  Minimum  of  1500  hours  flying  time 

7.  Qualified  jet  pilot 

Another  Langley  man  who  played  a  part  in  the  screening  process  was  Robert 
A.  Champine,  a  veteran  NAG  A  test  pilot  who  knew  what  kind  of  talents 
it  might  take  to  fly  into  space.  Although  not  an  STG  member,  he  was 
part  of  the  small  NASA/DOD  panel  that  evaluated  the  files  of  the  nearly 
600  military  service  test  pilots  who  had  applied  for  the  astronaut  positions. 
Of  the  seven  evaluation  criteria,  experience  as  a  test  pilot  was  clearly  the 
deciding  factor.24 

Ironically,  the  greatest  skepticism  about  the  Mercury  concept  existed 
inside  the  family  of  test  pilots.  Pathbreaking  NACA/NASA  test  pilots  like 
A.  Scott  Crossfield,  Joseph  A.  Walker,  and  even  the  young  Neil  Armstrong, 
who  in  10  years  was  to  become  the  first  man  to  walk  on  the  moon,  were  at 
first  not  in  favor  of  Project  Mercury.  Their  attitude  was  that  the  astronaut 
inside  the  ballistic  spacecraft  was  no  more  than  "Spam-in-a-can."  Charles 
E.  "Chuck"  Yeager,  the  air  force  test  pilot  who  broke  "the  sound  barrier" 
in  1947  in  the  X-l,  expressed  this  prejudice:  "Who  wanted  to  climb  into 
a  cockpit  full  of  monkey  crap?"25  This  was  a  crude  reference  to  the  noble 
primates  (such  as  "Ham"  and  "Enos")  who  flew  in  the  Mercury  spacecraft 
prior  to  the  astronauts  and  who  went  through  some  challenging  and  painful 
experiences  to  make  the  experience  of  humans  safer  and  more  certain. 

By  the  time  of  the  NASA  inspection,  all  seven  Mercury  astronauts  had 
been  in  training  at  Langley  under  the  STG's  technical  supervision  (and 
Langley  AFB's  administrative  care)  for  about  five  months.  Six  of  the  seven 
moved  into  the  area  with  their  families:  Carpenter  and  Cooper  lived  in 


The  absence  of  a  weight  requirement  is  incredible  given  the  demands  of  the  payload  on  the  launch 
rocket's  boosting  power  and  the  tight  squeeze  for  the  passenger  inside  the  Mercury  capsule. 

41 


Space/light  Revolution 

Hampton  just  across  the  tidal  river  from  the  air  force  base;  Grissom,  Schirra, 
and  Slayton  bought  ranch-style  homes  within  a  few  blocks  of  one  another 
in  the  new  Stoneybrook  Estates  subdivision  of  Newport  News;  and  Shepard 
drove  his  white  convertible  through  the  Hampton  Roads  Bridge  Tunnel  each 
day  from  his  family's  home  at  the  Naval  Air  Station  in  Virginia  Beach. 
Glenn  was  the  exception;  while  at  Langley  Field,  he  stayed  in  military  base 
quarters  and  commuted  to  his  home  in  Arlington,  Virginia,  on  weekends  to 
visit  his  wife  and  children.  Already  the  local  press  was  calling  the  astronauts 
"The  Peninsula's  Own"  and  trying  to  satisfy  an  adoring  public's  hunger  for 
even  the  most  mundane  details  of  the  astronauts'  everyday  existence,  such 
as  what  kind  of  fruit  juice  they  drank  for  breakfast.26 

The  film  shown  at  the  Project  Mercury  inspection  stop  said  little  about 
NASA's  selection  of  the  astronauts  and  showed  nothing  about  their  per- 
sonal lives;  it  concentrated  on  illustrating  key  aspects  of  their  training  for 
the  upcoming  Mercury  flights.  In  one  of  the  film's  early  scenes,  the  astro- 
nauts sat  in  a  classroom  listening  to  a  lecture  delivered  by  an  STG 
engineer.  This  lecture  was  one  in  a  series  organized  by  STG  member 
Dr.  Robert  Voas,  the  navy  psychologist  in  charge  of  coordinating  astro- 
naut training.  The  lecture  series  was  designed  to  introduce  formally 
the  astronauts  to  the  Mercury  program.27  Although  not  depicted  in  the 
film,  the  astronauts  also  took  a  short  course  equivalent  to  graduate-level 
study  in  the  space  sciences.  Henry  Pearson,  W.  Hewitt  Phillips,  and 
Clinton  E.  Brown  were  among  those  engineers  with  special  competencies 
in  reentry  physics,  astronomy,  and  celestial  mechanics  and  navigation  cho- 
sen to  teach  the  course. 

While  the  astronauts  learned  a  little  about  everything  pertinent  to  the 
program,  they  were  also  trained  to  specialize  in  particular  technical  areas. 
Carpenter  specialized  in  communications  and  navigation  equipment;  Cooper 
and  Slayton  concentrated  on  the  liaison  with  the  Army  Ballistic  Missile 
Agency  (ABMA,  later  NASA  Marshall  Space  Flight  Center)  and  the  launch 
vehicle  suppliers;  Glenn  focused  on  cockpit  layout;  Grissom  handled  in- 
flight control  systems;  Schirra  was  responsible  for  life-support  systems  and 
pressure  suits;  and  Shepard  followed  tracking  range  and  recovery.  Each 
astronaut  was  then  responsible  for  briefing  the  other  six  periodically  about 
what  he  had  learned.28 

The  inspection  film  of  1959  showed  the  Langley-based  STG  putting  the 
astronauts  through  several  spaceflight  simulation  systems  and  techniques 
to  familiarize  them  with  the  Mercury  capsule  and  evaluate  the  efficacy  of 
astronaut  capsule  control.  By  this  time  in  their  training,  the  astronauts  had 
already  ridden  on  the  end  of  the  50-foot  arm  of  the  centrifuge  at  the  Naval 
Aviation  Medical  Acceleration  Laboratory  at  Johnsville,  Pennsylvania.  The 
film  showed  one  of  the  astronauts  boarding  what  came  to  be  known  among 
the  astronauts  as  "the  wheel"  because  it  resembled  a  medieval  instrument 
of  torture.  Not  even  the  grimacing  face  of  the  astronaut,  as  he  desperately 
tried  to  operate  a  few  manual  controls,  could  communicate  how  miserable 

42 


The  First  NASA  Inspection 

the  experience  actually  was  for  the  rider,  who  was  being  pushed  back  in  the 
seat  as  the  wheel  picked  up  speed,  pinned  there  unable  to  move  either  arms 
or  legs,  breath  forced  out  of  the  lungs,  vision  narrowing  and  darkening,  and 
a  sharp  pain  growing  beneath  the  breastbone.  John  Glenn  recalls,  "At  16 
Gs*  it  took  just  about  every  bit  of  strength  and  technique  you  could  muster 
to  retain  consciousness."29 

The  training  at  Langley  was  a  little  easier,  at  least  physically.  The 
astronauts  made  several  "flights"  in  a  closed-loop  analog  simulator  that  had 
been  developed  by  the  training  devices  section  of  the  STG's  Operations 
Division.  This  simulator  had  a  basic  configuration  similar  to  the  X-15 
attitude  control  system  simulator  that  had  been  built  earlier  at  Langley. 
At  the  time  of  the  October  1959  inspection,  it  contained  a  simple  chair  with 
a  sidearm  controller  and  rudder  pedals.30  A  later  version  would  have  a 
three-axis  controller  and  a  molded  couch  like  those  individually  fitted  for 
each  astronaut  for  the  actual  Mercury  missions.  The  function  of  this  couch, 
which  was  one  of  many  ideas  supplied  by  the  STG's  brilliant  Maxime  Faget, 
was  to  protect  the  astronaut  against  the  high  G-forces  during  launch  and 
reentry.  In  one  scene  of  the  film,  two  of  the  finished  couch  forms  were  visible 
in  the  background;  in  tests  at  the  Johnsville  centrifuge,  such  couches  had 
proved  effective  for  loads  of  more  than  20  Gs.  The  movie  also  featured 
a  sequence  in  which  an  astronaut  used  the  sidearm  controller  to  move  his 
chair  through  various  changes  in  pitch,  roll,  and  yaw,  and  a  scene  showing 
an  overheated  astronaut  in  a  full  pressure  suit  undergoing  what  the  speaker 
called  "elevated  temperature  elevation."31 

"The  Space  Task  Group  has  found  the  seven  astronauts  inspiring  young 
men  with  whom  to  work,"  speakers  told  the  audience.  To  equip  them  with 
the  "detailed  knowledge  and  skills  that  the  pilot  of  a  pioneering  orbital 
space  capsule  must  possess,"  NASA  was  putting  them  through  "an  extensive 
program  of  training,  indoctrination,  and  specialized  education."  And  rest 
assured,  the  speakers  told  the  audience,  the  astronauts  were  preparing  for 
their  upcoming  launches  into  space  "with  an  enthusiasm  and  a  maturity 
that  are  vital  in  a  program  of  such  importance  to  our  nation."32 

The  speakers  did  not  mention  that  the  astronauts  sometimes  felt  they 
were  being  treated  like  guinea  pigs.  This  was  not  the  case  in  their  dealings 
with  the  STG  at  Langley.  As  the  astronauts  later  attested,  the  STG 
treated  them  as  "active  and  valuable  participants  in  the  safe  operation  of 
the  machine."  Bob  Gilruth  and  his  staff  had  been  dealing  directly  with  test 
pilots  in  NACA  aircraft  research  programs  since  before  World  War  II.  These 
years  of  experience  contributed  to  a  relationship  with  the  astronauts  that 
was  built  on  respect.33 

Much  to  the  disappointment  of  many  in  the  audience  at  the  NASA 
open  house,  especially  the  young  people,  the  living,  breathing  astronauts 
were  nowhere  to  be  seen.  Neither  Gilruth  nor  anyone  else  responsible  for 


"G"  is  the  symbol  representing  the  acceleration  due  to  gravity. 

43 


Spaceflight  Revolution 


L-59-4426 

Molded  astronaut  couches  line  the  Langley  model  shop  wall.   The  names  of  the  test 
subjects — Langley  employees — are  written  on  the  backs. 


PROJECT  MERCURY 


L-61-5741 

This  cutaway  drawing  was  used  by  the  STG  to  explain  the  Mercury  ballistic  capsule 
to  visitors  at  the  first  NASA  inspection. 


44 


The  First  NASA  Inspection 


Astronaut  John  Glenn  sits  within  the  cozy  cocoon  of  the  Mercury  spacecraft. 


the  astronauts  wanted  to  add  to  the  astronauts'  already  heavy  schedule 
by  keeping  them  in  front  of  several  thousand  sticky-fingered  and  camera- 
clicking  fans  for  an  entire  Saturday.  The  astronauts'  training  at  Langley 
included  a  rigorous  regimen  of  physical  exercise,  including  skin-diving 
operations  designed  to  simulate  weightlessness  and  the  kind  of  sensory 
disorientation  that  they  might  experience  during  reentry  from  space.  In 
Langley's  large  hydrodynamics  tank  (Building  720)  as  well  as  in  the  brackish 
water  of  the  Back  River,  an  inlet  of  the  Chesapeake  Bay  behind  the  East 
Area,  the  astronauts  were  learning  to  get  out  of  the  space  capsule  as  it 
floated  in  water.  Along  with  the  tiring  training  at  Langley,  the  astronauts 
also  made  trips  to  the  Johnsville  centrifuge;  to  Cape  Canaveral,  where  the 
countdown  for  their  manned  orbital  flights  would  be  made;  as  well  as  to 
the  McDonnell  Aircraft  Corporation  plant  in  St.  Louis,  where  the  Mercury 
capsules  were  being  built. 

Although  the  astronauts  were  excused  by  NASA  from  appearing  at  the 
open  house,  they  had  participated  in  the  inspection  earlier  in  the  week. 
They  were  not  assigned  to  give  speeches  or  conduct  tours,  but  they  were 
asked  to  mix  with  invited  guests  in  the  major  exhibit  hall  within  the  large 
aircraft  hangar,  where  the  makeshift  after-hours  wet  bar  called  "19th  Hole" 
was  set  up  and  most  socializing  occurred. 


45 


Spaceflight  Revolution 


John  Glenn,  who  in  three  years 
would  become  the  first  American 
to  orbit  the  earth  (20  February 
1962),  explains  a  feature  of  the 
Mercury  capsule  to  his  wife,  Annie 
Castor  Glenn,  whom  he  had  known 
since  his  New  Concord,  Ohio,  child- 
hood. 


L-59-7859 


Big  Joe,  Little  Joe 

The  success  of  two  recent  tests  for  Project  Mercury  lent  a  cautiously 
upbeat  mood  to  the  First  Anniversary  Inspection.  Five  weeks  earlier, 
on  9  September  1959,  the  project  reached  an  important  early  milestone 
with  what  the  inspection  speakers  called  the  "highly  successful  firing"  of 
"Big  Joe."  Big  Joe  was  a  one-ton,  full-scale  instrumented  mock-up  of  the 
proposed  Mercury  spacecraft  designed  to  test  the  efficacy  of  the  ablative 
heat  shield  and  the  aerodynamic  stability  of  the  capsule  design.  Speakers  at 
the  Project  Mercury  stop  boasted  that  the  Big  Joe  project  had  not  begun 
until  December  1958  and  was  flying  successfully  only  10  months  later.34 

After  showing  a  short  movie  of  Big  Joe's  launch  atop  an  Atlas  D  booster 
from  Cape  Canaveral,  the  STG  engineers  explained  that  although  the  launch 
was  normal,  the  two  outer  booster  engines  failed  to  jettison  as  planned 
because  of  a  malfunction;  the  capsule- Atlas  combination  rose  to  an  altitude 
of  only  about  100  miles.  This  was  nevertheless  high  enough  for  the  capsule, 
once  separated  from  the  Atlas,  to  fall  back  to  earth  in  conditions  that 
closely  simulated  orbital  reentry.  Another  short  movie  showed  the  shipboard 
recovery  of  the  capsule  by  a  navy  destroyer.  The  STG  speakers  explained 
that  the  recent  Big  Joe  test  not  only  proved  to  be  an  excellent  exercise  for 
the  military  recovery  teams  but  also  provided  data  that  confirmed  that  the 


46 


The  First  NASA  Inspection 

blunt-body  capsule  shape  had  performed  as  predicted  in  NASA  wind-tunnel 
and  other  laboratory  studies.  In  their  words,  the  Big  Joe  test  was  "the  first 
major  step"  in  proving  that  the  Mercury  design  concepts  were  feasible.35 

On  display  in  the  Aircraft  Loads  building  was  the  recovered  capsule; 
alongside  it  was  a  second  Big  Joe  boilerplate  capsule  mounted  on  a  Little 
Joe  booster  mock-up.  NASA  Langley  was  proud  of  Big  Joe.  A  small  group 
of  Langley  technical  service  people  under  STG's  Jack  Kinzler  had  actually 
fabricated  the  capsule's  afterbody,  including  the  upper  heat  shield  and  the 
parachute  deck,  while  another  NASA  group  under  Scott  Simpkinson  at  Lewis 
had  made  the  lower  part  of  the  capsule,  the  instrumentation,  the  controls, 
and  the  rest  of  the  heat  shield.  But  Langley  positively  doted  on  its  Little 
Joe.  Little  Joe  was  an  innovative  solid-fuel  rocket,  one  of  the  earliest  U.S. 
launch  vehicles  based  on  the  principle  of  the  clustered  rocket  engine.  (The 
Soviets  were  already  "clustering"  the  more  complex  and  troublesome  liquid- 
fuel  rocket  engines.)  STG  engineers  Max  Faget  and  Paul  Purser,  then  of 
Langley 's  PARD,  had  conceived  Little  Joe  as  a  space  capsule  test  vehicle 
even  before  the  establishment  of  NASA  and  the  formation  of  the  STG. 
Gilruth  understood  the  importance  of  the  Little  Joe  tests:  "We  had  to 
be  sure  there  were  no  serious  performance  and  operational  problems  that 
we  had  simply  not  thought  of  in  such  a  new  and  radical  type  of  flight 
vehicle."36  A  launch  of  Little  Joe  on  21  August  1959  had  failed,  but  at 
Wallops  Station  on  4  October  1959,  just  two  weeks  before  the  inspection, 
NASA  successfully  fired  one  of  the  "little"  test  rockets  to  an  altitude  of 
about  40  miles  over  the  Atlantic  Ocean  before  intentionally  destroying  it.37 

"Little"  was  relative,  of  course,  because  the  rocket  stood  50  feet 
tall,  weighed  28,000  pounds — the  gross  takeoff  weight  of  a  Douglas  DC-3 
airliner — and  had  a  cluster  of  eight  solid  propellant  engines  that  produced  a 
quarter  of  a  million  pounds  of  thrust  at  takeoff.  Nor  did  "little"  accurately 
describe  Little  Joe's  importance  to  the  Mercury  project.  For  the  4  October 
launch,  neither  the  capsule  nor  the  escape  rocket  had  been  instrumented, 
but  Little  Joe  would  carry  instrumented  pay  loads  to  varying  altitudes,  thus 
allowing  NASA  engineers  to  check  the  operation  of  the  escape  rocket  and 
recovery  systems.  This  they  could  do  from  Wallops  Island  before  proceeding 
to  the  more  expensive  and  difficult  phases  in  the  latter  part  of  the  program 
at  Cape  Canaveral.  In  ensuing  months,  Little  Joe  rockets  (models  I  and 
II)  also  provided  information  on  flight  stresses  as  they  related  to  "biological 
payloads."  The  first  of  these  payloads  was  Sam,  a  7-pound  Rhesus  monkey 
launched  from  Wallops  on  the  nose  of  a  Little  Joe  on  4  December  1959.  Sur- 
viving a  violent  ride  up  and  down  from  a  height  of  55  miles  with  a  parachute 
landing  into  the  Atlantic  Ocean,  Sam  gave  NASA  flight  engineers  a  better 
idea  of  how  human  astronauts  would  fare  during  their  upcoming  Mercury 
flights.38 

To  the  public,  Project  Mercury  looked  to  be  proceeding  smoothly.  The 
major  setback  of  July  1959,  when  the  first  Atlas-Mercury  production  vehicle 
failed  structurally  under  launch  loads  at  the  Cape,  was  not  mentioned 

47 


Space/light  Revolution 


L-59-4946 


Langley  technicians  constructed  the  Little 
Joe  capsules  in-house  in  Langley 's  shops 
(top).  A  crane  swings  a  capsule  into 
place  atop  Little  Joe  in  preparation  for  a 
launch  at  Wallops  Island  (right). 


L-59-5134 


48 


The  First  NASA  Inspection 


Langley's  Little  Joe  rocket  blasting  off 
(left)  from  Wallops  Island  in  the  fall 
of  1959.  Max  Faget  thought  that  Little 
Joe  could  be  made  reliable  enough 
to  carry  a  man,  but  Gilruth  eventu- 
ally scrapped  the  idea,  deciding  to  use 
Redstone  and  Atlas.  Below,  the  Little 
Joe  capsule  is  recovered  at  sea. 


L-59-8427 


L-64-6455 


49 


Space/light  Revolution 

in  Langley's  open-house  presentation.  To  everyone  behind  the  scenes  at 
Langley,  Project  Mercury  was  in  fact  advancing  at  breakneck  speed.  In  the 
period  between  early  October  1958  and  mid-January  1959,  specifications  for 
the  Mercury  capsule  had  been  prepared  and  sent  to  the  aerospace  industry 
with  a  Request  for  Proposals;  the  bidders  had  been  briefed;  all  the  source 
selection  (evaluation  of  proposal)  activity  had  taken  place;  and  the  contract 
had  been  placed.  That  was  not  all.  During  the  same  period,  the  STG 
procured  Atlas  rockets  and  launch  services  from  the  air  force;  worked  out  a 
plan  with  the  army  (and  Wernher  von  Braun's  rocket  team  in  Huntsville)  for 
Redstone  boosters;  drew  up  the  specifications  for  Little  Joe;  tested  escape 
rockets  over  the  beach  at  Wallops;  and  were  in  the  midst  of  a  wide  range 
of  tests  at  Langley.  The  STG  also  had  to  present  technical  reviews  of  the 
project  to  NASA  headquarters  officials  approximately  every  two  months. 
To  do  all  this,  every  member  of  the  STG  worked  holidays,  evenings,  and 
weekends.  "These  were  the  days  of  the  most  intensive  and  dedicated  work 
[by]  a  group  of  people  that  I  have  ever  experienced,"  Gilruth  recalls  proudly. 
This  kind  of  performance  could  have  occurred  only  "in  a  young  organization 
that  had  not  yet  solidified  all  of  its  functions  and  prerogatives."3 

This  performance  could  have  happened  only  in  an  organization  whose 
staff  members  did  not  know — or  care  to  know — the  difference  between  the 
possible  and  the  impossible  until  they  found  out  for  themselves. 


50 


Carrying  Out  the  Task 


There  are  no  billboards  heralding  the  birthplace  of  the 
Nation's  [space]  program.  There  are  no  colorful  ban- 
ners proclaiming  it  the  homebase  for  the  U.S.  's  seven 
astronauts.  Yet  nestled  at  one  end  of  the  historic  Vir- 
ginia Peninsula,  a  small  group  of  buildings  were  the 
setting  for  the  most  penetrating  research  and  develop- 
ment programs  of  our  time.  .  .  .  It  was  here  at  the 
NASA  Langley  Research  Center  that  America  took  its 
first  step  into  space. 

—Virginia  Biggins 
Newport  News  Daily  Press 


For  Bob  Gilruth,  the  chief  operational  officer  of  the  U.S.  manned  space 
program,  NASA's  First  Anniversary  Inspection  meant  only  a  brief  respite 
from  the  torturously  hectic  schedule  he  had  been  following  for  more  than 
a  year.  As  head  of  Project  Mercury,  he  had  given  dozens  of  talks  and  had 
answered  thousands  of  questions  in  the  past  15  months  about  America's 
highly  publicized  enterprise  to  send  a  man  into  space.  He  had  made 
presentations  before  Congress,  to  Dr.  Killian  and  the  rest  of  the  President's 
Science  Advisory  Committee,  and  to  the  senior  staff  of  the  Advanced 
Research  Projects  Agency  (ARPA)  including  agency  heads  Roy  Johnson  and 
Dr.  Herbert  York.*  "Some  of  these  gentlemen  were  not  at  all  enthusiastic 
about  our  plan  to  put  a  man  into  space,"  Gilruth  later  acknowledged.  In 
fact,  Presidential  Science  Adviser  Dr.  George  Kistiakowsky  had  remarked 
with  great  displeasure  that  the  plan  "would  be  only  the  most  expensive 
funeral  man  has  ever  had."1 


The  secretary  of  defense  had  established  ARPA  in  January  1958  to  run  U.S.  space  programs  on  an 
interim  basis  until  NASA  was  established. 

51 


Space/light  Revolution 

But  at  least  during  the  anniversary  inspection  the  pressure  was  off; 
officially,  Gilruth  was  just  one  of  the  guests  touring  with  the  red  group. 
At  the  Mercury  stop,  the  eight  men  from  the  STG  had  to  put  on  the  good 
show  that  everyone  had  come  to  expect,  and  for  once  he  could  sit  back  and 
listen  to  someone  else  do  the  talking. 

For  the  balding  45-year-old  aeronautical  engineer  from  Nashawauk,  Min- 
nesota, Project  Mercury  had  started  in  the  hot  summer  of  1958  while  on 
assignment  in  Washington,  B.C.  Dr.  Hugh  Dryden  had  needed  help  putting 
together  a  plan  and  a  budget  for  the  new  space  agency,  and  Gilruth,  with 
about  20  senior  men  from  Langley  and  the  other  NACA  laboratories,  went 
to  lend  a  hand.  Eisenhower  had  not  yet  given  specific  responsibility  for 
management  of  the  nation's  manned  spaceflight  program  to  the  soon-to-be 
NASA,  nor  had  he  officially  named  Glennan  the  NASA  administrator.  Abe 
Silverstein,  subsequent  head  of  space  projects  at  NASA  headquarters,  had 
not  yet  come  up  with  the  name  "Mercury"  for  the  proposed  manned  satellite 
project.  In  one  large  room  on  the  sixth  floor  of  the  NACA  headquarters, 
Gilruth  and  associates  worked  feverishly  through  the  muggy  midsummer  to 
put  together  a  plan  for  a  man-in-space  program  that  would  be  acceptable 
not  only  to  the  reincarnated  NACA  but  also  to  ARPA,  the  president,  and 
his  scientific  advisers.2 

"In  order  to  do  this,"  Gilruth  remembers,  "I  collected  a  select  group 
of  people  ...  to  form  a  sort  of  task  force."  The  members  of  this  original 
group  included  Langley's  Max  Faget,  head  of  the  Performance  Aerodynamics 
Branch  of  PARD;  Paul  Purser,  head  of  the  High  Temperature  Branch  of 
PARD;  Charles  W.  Mathews,  head  of  the  Stability  and  Control  Branch  of 
the  Flight  Research  Division;  Charles  H.  Zimmerman,  assistant  chief  of  the 
Stability  Research  Division;  and  three  men  from  Lewis.  These  men  were 
called  from  the  10  telephones  specially  installed  in  the  NACA's  big  sixth- 
floor  room  and  were  told  to  "be  in  Washington  tomorrow  afternoon."  As 
Zimmerman  remembers: 

I  said,  well,  what  for?  [The  voice  said,]  "I  can't  tell  you  what  for."  Who  am  I  supposed 
to  see?  [The  voice  said,]  "Just  be  in  the  Washington  office  tomorrow  morning/'  I 
went  to  the  Washington  office  and  I  stayed  there  three  or  four  months.  ...  I  wasn't 
told  anything,  just  be  there.  I  had  to  go  and  tell  my  wife  I'm  going.  [I]  didn't  win  a 

o 

popularity  contest  that  day. 

Gilruth  brought  in  several  other  NACA  engineers  for  consultation  when 
their  expertise  was  needed.  He  called  in  PARD's  top  engineering  designer 
Caldwell  Johnson,  who  had  been  hired  by  the  NACA  as  a  model  builder 
in  1937  at  the  age  of  18;  Johnson's  job  was  to  put  the  first  design  of  the 
Mercury  capsule  on  paper.  The  result  was  an  elegant  series  of  freehand  pen- 
and-ink  sketches  that  artistically  put  many  detailed  engineering  drawings  to 
shame.  Near  the  end  of  the  summer,  two  more  engineers  from  Lewis  and 
one  from  Langley,  Charles  Donlan,  joined  the  group  to  finalize  and  fine-tune 
the  Mercury  plan. 

52 


Carrying  Out  the  Task 

The  work  of  the  task  force  turned  out  well  both  in  the  short  term  and  the 
long  run.  Thinking  back  on  the  substance  of  these  early  talks  about  what 
came  to  be  Project  Mercury,  Gilruth  would  be  impressed  by  how  closely 
the  STG  was  able  to  follow  the  original  plan  of  that  summer:  "We  said  we 
would  use  the  Atlas  rocket;  a  special  space  capsule  with  a  [NACA-proven] 
blunt  heat  shield;  and  parachutes  for  a  landing  at  sea.  All  these  things 
were  to  work  out  very  much  as  we  proposed."4  During  that  hot  summer  of 
1958,  Max  Faget,  Caldwell  Johnson,  and  Lewis's  Andre  Meyer  also  came  up 
with  the  idea  of  an  escape  rocket  to  enable  the  capsule  to  get  away  from  a 
malfunctioning  launch  rocket,  and  Faget  conceived  the  form  of  the  contour 
couch,  which  would  help  to  protect  the  astronauts  against  the  high  G-forces 
during  launch  and  reentry. 

Much  about  the  group's  Mercury  concept  was  not  all  that  new:  the 
aerodynamic  benefits  of  the  blunt-body  shape  had  been  discovered  (at 
least  for  ballistic  nose  cones)  by  H.  Julian  "Harvey"  Allen  and  Alfred  J. 
Eggers  at  NACA  Ames  in  the  early  1950s.5  Since  then,  several  important 
notions  about  ballistic  reentry  vehicles  had  been  germinating  in  the  minds  of 
Gilruth's  colleagues  in  PARE),  notably  in  the  brilliant  one  belonging  to  the 
outspoken  Max  Faget.  (Because  he  was  one  of  the  most  intuitive  researchers 
on  the  Langley  staff,  jealous  colleagues  jibed  that  his  name  stood  for  Fat- Ass 
Guess  Every  Time.)  By  the  launch  of  Sputnik  1,  Faget  had  proposed  that 
a  simple  nonlifting  shape,  if  properly  designed,  could  follow  a  ballistic  path 
when  reentering  the  atmosphere  without  overheating  or  accelerating  at  rates 
dangerous  to  the  astronaut.  Drag  would  slow  the  capsule  as  it  reentered  the 
atmosphere.  Furthermore,  the  shape — though  basically  nonlifting — could 
generate  the  slightest  amount  of  aerodynamic  lift  necessary  to  permit  the 
capsule  to  make  one  or  two  simple  maneuvers  during  reentry.  Faget  had 
made  some  rough  tests  to  prove  this  theory.  From  the  balcony  overlooking 
the  PARD  shop,  he  had  flipped  two  paper  plates  that  had  been  taped 
together  into  the  air.  "I  thought  he  was  crazy  at  first,"  remembers  fellow 
PARD  engineer  J.  Thomas  Markley.  "Max,  what  are  you  doing?"  asked 
Markley  in  amusement.  Faget  answered,  "I  think  these  things  will  really  fly. 
We  really  have  some  lift-over-drag  in  this  thing."6 

A  few  months  after  the  paper-plate  toss,  at  the  last  NACA  Conference  on 
High-Speed  Aerodynamics  held  at  Ames  in  March  1958,  the  feisty  5-foot-6- 
inch  Faget  gave  a  talk  entitled  "Preliminary  Studies  of  Manned  Satellites- 
Wingless  Configuration:  Non-Lifting,"  which  was  coauthored  by  Langley's 
Benjamin  J.  Garland  and  James  J.  Buglia.  In  the  talk  Faget  put  forward 
most  of  the  key  items  that  NASA  would  later  use  in  Project  Mercury:  a 
ballistic  shape  weighing  some  2000  pounds  and  having  a  nearly  flat-faced 
cone  configuration,  small  attitude  jets  for  controlling  the  capsule  in  orbit, 
retrorockets  to  bring  the  capsule  down,  and  a  parachute  for  final  descent. 
"As  far  as  reentry  and  recovery  are  concerned,"  Faget  concluded  his  talk, 
"the  state  of  the  art  is  sufficiently  advanced  so  that  it  is  possible  to  proceed 


53 


Spaceflight  Revolution 

confidently  with  a  manned  satellite  project  based  upon  the  ballistic  reentry 
type  of  vehicle."7 

Not  everyone  was  so  confident.  In  the  wake  of  Sputnik,  several  interesting 
concepts  for  manned  satellites  had  popped  up.  Some  advocates  of  these  al- 
ternatives disdained  Faget's  proposed  ballistic  approach  because,  as  Gilruth 
explained,  it  represented  "such  a  radical  departure  from  the  airplane."8  This 
man-in-the-can  approach  was  too  undignified  a  way  to  fly.  Many  concerned 
with  America's  new  space  program  searched  for  another  plan:  Couldn't  a 
pilot  fly  into  space  and  back  in  some  honest-to-goodness  flying  machine? 
Why  not  doctor  the  X-15  so  a  pilot  could  take  it  into  orbit  and  back  with- 
out burning  up?  Or  why  not  push  to  quickly  build  one  of  the  hypersonic 
gliders  that  had  been  drawn  up  on  paper?  One  of  the  most  innovative  con- 
cepts for  such  a  space  plane,  proposed  by  Langley's  Chuck  Mathews,  called 
for  a  craft  similar  to  NASA's  later  Space  Shuttle.  Mathews'  plane  would 
have  a  circular  wing  and  would  glide  back  from  space  at  a  high  angle  of  at- 
tack. During  reentry,  most  of  the  intense  heat  caused  by  the  friction  would 
therefore  be  confined  to  the  wing's  lower  surface.  Upon  reaching  the  atmo- 
sphere, the  vehicle  would  pitch  over  and  fly  to  a  landing  like  a  conventional 
airplane.9 

Such  concepts  sparked  much  interest  in  the  months  after  Sputnik. 
Gilruth  and  the  rest  of  the  team  planning  for  Project  Mercury  considered 
the  merits  of  each  one  separately.  Several  of  the  ideas  could  have  been 
made  to  work  in  time,  but  the  new  space  agency  did  not  have  time  to  spare. 
Everything  indicated  that  the  Soviets  were  intent  on  launching  a  man  into 
space,  and  the  United  States  was  determined  to  beat  them  to  it.  The  Atlas 
rocket,  the  most  powerful  American  booster  at  the  time,  was  not  capable 
of  lifting  more  than  about  2000  pounds  into  orbit,  which  ruled  out  the  hy- 
personic glider  concepts.  Furthermore,  even  the  Atlas  was  still  horribly 
unreliable.  Only  one  out  of  eight  Atlases  had  been  launched  successfully; 
the  other  seven  had  staggered  off  course  or  blown  up.  If  the  United  States 
wanted  to  win  this  important  second  leg  of  the  space  race,  waiting  for  the 
development  of  a  bigger  and  more  dependable  missile  capable  of  lifting  the 
far  greater  weight  of  a  small  space  plane  did  not  make  sense.  "It  seemed 
obvious  to  our  group,"  Gilruth  would  explain  many  years  later,  "that  only 
the  most  simple  ballistic  capsule  could  be  used  if  manned  spaceflight  were 
to  be  accomplished  in  the  next  few  years." 10 

Several  options  may  have  been  more  technologically  attractive  to  some 
NASA  engineers,  but  Faget's  plan  appeared  the  best  to  achieve  America's 
immediate  space  objectives.  In  some  respects  the  plan  was  an  ungainly 
(some  have  said  unimaginative,  even  ugly)  way  to  send  an  American  into 
space,  yet  in  1959  it  seemed  the  only  way  to  do  so  quickly.  As  Gilruth  would 
say  later,  Project  Mercury 

wasn't  pretty  like  a  flower  or  a  tree.  But  it  had  no  bad  traits.  It  was  designed  as  a 
vehicle  for  a  man  to  ride  in,  and  circle  the  earth.  With  its  blunt  body,  its  retrorockets 


and  parachutes,  it  was  an  elegant  solution  to  the  problem. 


54 


Carrying  Out  the  Task 

But  a  solution  that  was  elegant  in  conception  had  no  guarantee  of  becoming 
a  practical  success.  Once  ARPA  heads  Roy  Johnson  (a  former  General 
Motors  executive)  and  Herbert  York  (a  distinguished  atomic  physicist) 
approved  the  plan  on  7  October  1958  and  NASA  gave  the  go-ahead,  Gilruth 
and  his  people  were  left  with  the  job  of  making  Project  Mercury  work. 


A  Home  at  Langley 

Gilruth  and  associates  returned  to  Langley  Research  Center  from  the 
nation's  capital  in  mid-October  1958  and  immediately  began  to  contend 
with  the  unknown  challenges  of  putting  together  an  organization  that  could 
manage  an  operation  much  bigger,  more  complicated,  and  far  riskier  than 
any  previously  undertaken  by  the  NACA.  In  approving  the  project,  Keith 
Glennan's  comment  had  been,  "All  right.  Let's  get  on  with  it."  Bob  Gilruth 
remembers  that  at  the  time  he  "had  no  staff  and  only  [oral]  orders  to  return 
to  Langley  Field."  When  Gilruth  politely  pressed  the  administrator  for  some 
details  about  how  he  was  to  implement  the  plan  in  terms  of  staffing,  funding, 
and  facilities,  Glennan  reiterated  brusquely,  "Just  get  on  with  it."12 

Gilruth's  yet-to-be-built  organization  was  given  temporary  quarters  at 
Langley,  where  it  would  act,  again  temporarily,  as  a  quasi-independent 
NASA  field  unit  reporting  directly  to  Abe  Silverstein's  Office  of  Space  Flight 
Development  in  Washington.  Though  Langley  lacked  management  control 
over  the  new  group,  the  center's  support  of  the  task  group's  ambitious 
program  proved  remarkably  strong. 

Almost  everything  about  the  initial  organization  and  early  operation 
of  Gilruth's  group  happened  catch- as-catch-can.  Even  the  name  of  the 
STG  itself  suggested  a  makeshift  character,  as  if  NASA  did  not  want  to 
raise  expectations  too  high  about  meeting  the  Soviet  challenge.  One  STG 
member  suggests  that  the  choice  of  the  title  "Space  Task  Group"  amounted 
to  a  "conscious  effort  to  put  the  work  in  proper  perspective  and  avoid 
grandiose  organizational  concepts  at  a  time  when  satellite  development 
experience  was  limited  to  basketball-  and  grapefruit-sized  objects."  The 
timid  nomenclature  might  protect  NASA  if  the  manned  satellite  program 
did  not  work  as  planned.  NASA  could  say  that  only  one  task  failed;  the  rest 
of  NASA's  operation  was  proceeding  nicely. 

Excluding  Bob  Gilruth,  the  most  important  person  behind  the  formula- 
tion of  the  STG  was  Langley's  Floyd  Thompson.  Although  still  nominally 
the  laboratory's  number  two  man,  Thompson  had  been  serving  as  the  di- 
rector for  some  time  because  of  Henry  Reid's  rather  relaxed  approach  to 
his  impending  retirement.  According  to  Gilruth,  Thompson  "was  all  for 
me,  because  he  knew  that  if  we  didn't  succeed,  NASA  wouldn't  succeed." 
He  realized  that  Gilruth  would  need  substantial  center  support  until  the 
slow-grinding  paper  mill  at  NASA  headquarters  made  alternative  provisions. 
Thus,  when  Gilruth  asked  Thompson  how  he  could  get  the  men  and  women 

55 


Space/light  Revolution 

he  needed  for  the  STG,  Thompson  told  him  simply  to  write  a  short  memo- 
randum stating  that  he  had  been  authorized  by  Administrator  Glennan  to 
draft  personnel.  Gilruth  wrote  that  memo  on  3  November  1958  and  per- 
sonally took  it  down  the  hall  to  the  associate  director's  office.  The  letter 
amounted  to  one  brief  paragraph: 

The  Administrator  of  NASA  has  directed  me  to  organize  a  space  task  group  to 
implement  a  manned  satellite  project.  This  task  group  will  be  located  at  the  Langley 
Research  Center  but,  in  accordance  with  the  instructions  of  the  Administrator,  will 
report  directly  to  NASA  Headquarters. 

For  the  project  to  proceed  with  the  utmost  speed,  Gilruth  proposed  to  form 
his  group  around  a  nucleus  of  key  Langley  personnel,  the  majority  of  whom 
had  already  worked  with  him  on  the  project  at  NASA  headquarters. 

Thompson  did  not  want  to  run  the  STG  himself,  because  he  recognized 
that  a  quasi-independent  person  like  Gilruth,  not  a  center  director,  was 
"the  best  guy  to  do  it."15  At  the  same  time,  Thompson  wanted  Gilruth,  a 
personal  friend,  to  have  a  circle  of  bright  and  trustworthy  individuals  around 
him.  In  particular,  Thompson  felt  Gilruth  should  have  a  good,  solid  deputy, 
so  he  gave  him  Donlan,  his  own  energetic  assistant.*  For  the  past  seven 
or  eight  years  Donlan  had  been  enjoying  the  enviable  job  of  probing,  at  his 
own  discretion,  into  different  areas  of  the  laboratory's  research  programs 
and  acting  as  its  technical  conscience.  "Thompson  thought  Gilruth  needed 
me,  because  Bob  liked  to  play  around  with  ideas  and  not  pay  too  much 
attention  to  the  actual  running  of  the  technical  functions,"  Donlan  states. 
So,  "for  the  first  time  in  [my]  professional  career,"  Thompson  told  Donlan, 
"[I]  am  going  to  make  a  recommendation."  Thompson  asked  Donlan  to  join 
the  STG  as  Gilruth's  deputy.16 

Gilruth's  terse  memo  created  a  rapidly  expanding  core  group  of  space 
pilgrims.  According  to  one  cynic,  these  pilgrims  were  like  those  who  came 
to  America  on  the  Mayflower,  "considering  how  many  people  tell  you  they 
were  in  it."17  But  Gilruth  asked  by  name  for  the  transfer  of  only  36 
Langley  personnel  plus  10  engineers  from  Lewis  laboratory.  Lewis  provided 
rocket-engine  and  electronic  engine-component  specialists — the  experts  in 
aerospace  propulsion  systems  that  Langley  lacked. 

Fourteen  of  the  36  Langley  personnel  belonged  to  PARD.  This  major 
and  quasi-independent  division  of  the  laboratory  had  been  headed  for  a 
time  in  the  early  1950s  by  Gilruth.  The  work  of  PARD  had  always 
required  the  management  of  flight  operations  (albeit  pilotless  ones)  and 
had  dabbled  with  hardware  development.  While  studying  the  aerodynamics 
of  various  missiles  and  missile  nose-cone  configurations  during  the  past 


>k 

Later  on,  Thompson  would  "feel  an  obligation"  to  bring  Donlan  back  to  Langley,  making  him 
Langley's  associate  director  in  March  1961.  Donlan  stayed  on  as  associate  director  (later  renamed 
deputy  director)  until  May  1968  when,  at  the  request  of  the  NASA  administrator,  Donlan  transferred 
to  NASA  headquarters  and  became  the  deputy  associate  administrator  for  Manned  Space  Flight. 

56 


Carrying  Out  the  Task 

few  years,  PARD  engineers  had  established  launch  procedures  at  Wallops 
Island,  experimented  with  the  principles  of  rocket  staging,  developed  key 
technologies  for  missile  guidance  and  control  systems,  and  built  or  refined 
sensitive  instrumentation  for  telemetry  studies.  They  had  also  supported 
manned  satellite  proposals  from  the  Defense  Department.  In  1957  and  early 
1958,  before  ARPA/NASA  approval  for  Project  Mercury,  PARD  engineers 
had  given  research  support  for  Project  MISS,  the  unfortunate  acronym 
of  the  "Man-in-Space-Soonest  project,"  an  air  force  concept  for  simple 
manned  orbital  flights  that  in  some  technical  respects  presaged  the  Mercury 
concept.  This  early  work  in  support  of  the  manned  satellite  proposals  had 
taken  the  PARD  engineers  into  such  areas  as  space  environmental  controls, 
communications  systems,  and  heat-shield  technology.  Having  had  this 
experience,  many  members  of  PARD  were  not  as  concerned  as  other  Langley 
employees  about  the  possible  compromise  of  traditional  laboratory  research 
functions  implicit  in  heavy  involvement  in  Project  Mercury.  In  terms  of 
technological  expertise  and  organizational  culture,  PARD  people  were  the 
most  naturally  inclined  at  Langley  to  become  involved  in  the  planning  and 
management  of  NASA's  manned  spaceflight  program.18 

Of  the  remaining  22  STG  staff  members  recruited  from  Langley,  10  were 
from  research  divisions  other  than  PARD;  4  had  been  working  in  the 
Fiscal  Division,  central  files,  or  in  the  stenographic  pool;  and  8  were 
either  secretaries  in  PARD,  stenographers,  or  "computers"  (operators  of 
the  calculating  machines).  Thompson  agreed  to  give  Gilruth  all  the  people 
he  asked  for,  save  one:  a  young  electrical  engineer,  William  J.  Boyer.  The 
Instrument  Research  Division  (IRD)  wanted  to  keep  Boyer,  and  he  was  not 
anxious  to  be  transferred.  The  head  of  that  division,  Edmond  C.  Buckley, 
finally  found  a  satisfactory  replacement  in  Howard  C.  Kyle. 

Most  of  the  original  STG  crew  signed  up  voluntarily;  they  were  young, 
relatively  unestablished,  and  they  relished  the  challenge.  At  ages  45  and 
42,  respectively,  Gilruth  and  Donlan  were  experienced  enough  to  recognize 
the  difficulties  of  the  job  ahead,  but  many  of  their  subordinates  were  naive 
about  the  ways  of  the  world  and  did  not  consider  the  serious  hazards  facing 
them.  Jack  Kinzler,  a  skilled  master  craftsman  in  the  West  Area  machine 
shop,  recalls  that  he  had  grown  "so  consumed  with  space"  after  Sputnik 
that  he  just  dropped  everything  when  Gilruth  called  him  to  join  the  group. 
After  accepting  the  transfer,  Kinzler  then  had  a  devil  of  a  time  fighting  off 
a  swarm  of  excited  co-workers  who  wanted  to  move  to  the  STG  with  him. 
When  the  21-year-old  Lewis  engineer  Glynn  Lunney  heard  about  what  the 
STG  was  doing,  he  thought,  "Gee,  that  looks  like  it  would  be  a  hell  of  a  lot 
of  fun — let's  go  do  that!"  Carl  Huss  and  Ted  Scopinski  worked  at  the  same 
desk  in  the  Aircraft  Loads  Laboratory  in  Langley 's  West  Area.  The  two 
engineers  recall  one  day  in  late  1958,  after  they  had  heard  so  much  about 
the  STG  from  former  co-worker  John  P.  Mayer:  "[We]  looked  at  each  other 
and  asked  why  we  didn't  transfer  over  to  the  Space  Task  Group.  So  we 
did."19 

57 


Spaceflight  Revolution 

Wild  enthusiasm  might  have  been  confined  to  the  young  and  inexpe- 
rienced, but  strong  passion  for  Project  Mercury  was  not.  Donlan  looked 
upon  the  manned  satellite  project  "as  a  pioneering  effort  of  a  type  that 
comes  along  only  about  once  in  a  half  century."  To  him,  the  project  offered 
a  moment  in  history  that  would  be  "similar  to  aviation  when  Lindbergh 
flew  the  ocean."  He  never  doubted  that  he  should  join  the  STG:  "I  had 
to  participate  in  what  I  instinctively  felt  would  be  a  breathtaking  opera- 
tion, and  I  decided  to  do  so  without  much  thought  as  to  the  long-range 
possibilities."20  In  the  end,  his  time  with  the  STG  (November  1958-May 
1961)  did  not  hurt  his  career.  When  he  resigned  his  position  as  the  STG's 
number  two  man,  he  rejoined  the  Langley  operation  as  Floyd  Thompson's 
associate  director. 

The  rest  of  Langley's  senior  staff  was  not  as  easily  impressed  by  the  man- 
in-space  program.  With  the  exception  of  the  two  men  from  the  director's 
office,  only  one  member  of  Langley's  senior  staff  joined  the  STG:  Charles 
Zimmerman,  assistant  chief  of  the  Stability  Research  Division.  Zimmerman 
was  not  keen  about  the  assignment.  "It  was  a  traumatic  experience  as  far  as 
I  was  concerned,"  Zimmerman  remembers.  After  spending  a  hectic  summer 
in  Washington  with  Gilruth's  planning  group,  he  said,  "The  hell  with  this." 
He  got  in  touch  with  Henry  Reid  and  told  him  that  he  wanted  to  come  back 
to  Langley.  After  taking  a  week  off  to  vacation  in  Canada,  he  returned  to 
Langley  Field.  "I  got  back  home  on  Friday  and  was  going  to  go  to  work 
on  Monday,"  Zimmerman  recalls,  but  that  Friday  night  a  colleague  came  to 
break  the  news  that  Zimmerman  had  been  assigned  to  the  Mercury  group. 
"So,  there  I  was  in  it  again."21  Once  more,  Zimmerman  had  to  put  aside 
his  precious  airplane  work.* 

At  51,  Zimmerman  was  the  old  man  of  the  STG;  several  of  the  others 
were  young  enough  to  be  his  children.  He  had  started  his  career  at  NACA 
Langley  in  1929,  only  two  years  after  Lindbergh's  transatlantic  flight,  and 
like  many  NACA  researchers  of  his  generation,  he  was  not  comfortable  with 
the  idea  of  moving  away  from  aeronautics  into  the  management  of  a  large 
manned  space  program.  For  Zimmerman  and  most  other  senior  Langley  staff 
members,  the  excitement  of  the  program  was  not  enough  to  compensate  for 
the  headaches  and  perhaps  even  the  career  risks  associated  with  moving 
outside  the  comfortable  confines  of  aeronautical  research.  Perhaps  the 
country's  interest  in  manned  spaceflight  was  just  a  passing  fancy,  some  of 
the  older  men  thought.  Project  Mercury  had  been  authorized,  but  nothing 
else  up  to  this  point  had  been.  Throwing  in  with  the  lot  of  the  "space 
cadets"  meant  accepting  a  great  many  technological,  political,  institutional, 
and  personal  career  unknowns,  t  If  the  initial  series  of  Mercury  launches 
came  off  successfully,  the  manned  space  program  would  probably  continue 


Zimmerman,  famous  for  the  XF5U  "flying  flapjack,"  which  he  designed  for  Vought  during  the  1940s, 
had  been  busy  for  a  number  of  years  trying  to  make  the  conventional  airplane  into  a  VTOL  machine. 
*    Space  cadet  is  an  expression  of  derision  taken  from  a  popular  American  television  show  of  the  1950s. 

58 


Carrying  Out  the  Task 

in  some  form,  and  it  might  even  be  expanded,  but  late  in  1958  no  one  could 
be  any  more  sure  about  that  than  they  could  be  about  the  outcome  of  the 
upcoming  1960  presidential  election,  on  which  so  much  about  the  course  of 
the  U.S.  space  program  would  ultimately  depend. 

With  the  exception  of  the  graying  triumvirate  of  Zimmerman,  Gilruth, 
and  Donlan,  the  entirety  of  Langley's  senior  management  stayed  where  they 
were  in  the  organization  and  continued  what  they  had  been  doing.  At  least 
a  few  of  the  senior  staff  also  privately  advised  their  juniors  to  do  the  same. 
One  member  of  the  STG  remembers  that  his  division  chief  tried  to  persuade 
him  not  to  accept  the  transfer  to  the  STG.  "You  don't  want  to  ruin  your 
career,"  the  division  chief  told  him.  "There's  nothing  going  to  come  of  this, 
and  you're  going  to  be  hurt  by  it."  Manned  spaceflight,  he  warned,  was  just 
a  fad.22 

Many  veteran  employees  felt  that  "it  just  wasn't  the  Langley  way"  to 
implement  big  projects  like  Mercury.  The  laboratory  had  flourished  for 
more  than  40  years  by  doing  research,  not  by  implementing  things.23  It 
had  remained  strong  and  autonomous  by  developing  its  own  competencies 
and  by  doing  nearly  everything  that  involved  research  in-house,  but  Project 
Mercury  was  to  be  based  on  considerable  work  that  was  contracted  out 
to  industry.  The  people  responsible  for  the  contract  work  would  have  to 
cover  many  new  fronts:  they  had  to  prepare  space  capsule  specifications; 
evaluate  contractor  proposals,  then  monitor  the  awarded  contracts;  procure 
Redstone  rockets  from  the  army  and  Atlas  rockets  from  the  air  force;  arrange 
for  launch  services;  coordinate  recovery  operations;  and  so  on.  Skeptics 
feared  that  members  of  the  STG  would  be  so  caught  up  in  the  urgency 
of  managing  contract  work  and  in  refereeing  contractor  haggling  sessions 
(much  to  his  chagrin,  Zimmerman  became  chief  of  the  STG's  Engineering 
and  Contract  Administration  Division)  that  they  would  not  be  conducting 
much  research,  if  any.  Becoming  bureaucrats  rather  than  staying  technical 
personnel  was  a  fate  too  horrible  to  ponder.  To  this  day,  Bob  Gilruth  holds 
his  forehead  when  remembering  how  Langley  colleagues  would  approach 
him  during  the  heyday  of  Mercury  not  to  inquire  whether  he  had  had  any 
good  ideas  recently  but  rather  to  ask  snidely,  "Well,  have  you  let  any  good 
contracts  today?"24  His  old  NACA  associates  might  have  envied  Gilruth 
the  publicity  he  was  receiving,  but  they  did  not  envy  him  his  work. 

Gilruth's  senior  colleagues  who  did  not  want  to  join  the  STG  did 
follow  Floyd  Thompson's  example  of  helpfulness  and  energetically  supported 
NASA's  manned  satellite  project  through  traditional  research  avenues. 
"At  the  outset  of  the  program,  Langley  threw  all  of  its  resources  behind 
the  infant  STG,"  Thompson  reflected  in  1970,  "providing  technical  and 
administrative  support  informally  as  required,  just  as  though  the  STG  was 
a  part  of  Langley  and  not  a  separate  organization."25  Besides  providing 
extensive  support  for  the  development  and  implementation  of  the  Big  Joe 
and  Little  Joe  projects,  dozens  of  center  personnel  conducted  experimental 
studies  aimed  at  evaluating  the  performance  of  the  Mercury  spacecraft  at 

59 


Space/light  Revolution 

launch,  in  space,  during  reentry,  and  during  its  ocean  recovery.  Dozens  of 
others  became  involved  in  engineering,  shop,  instrumentation,  and  logistic 
support  for  much  of  the  STG's  own  in-house  testing. 

For  example,  in  1959  a  battery  of  wind-tunnel  tests  using  scale  models 
of  the  Mercury  capsule  and  capsule-booster  combinations  had  helped  to 
provide  needed  data  about  lift,  drag,  static  stability,  trajectories,  heat 
transfer,  heat-shield  pressures,  and  afterbody  pressures;  only  after  hundreds 
of  these  tests  would  the  shape  and  appearance  of  the  Mercury  capsule  be 
refined  and  finalized.  At  Wallops,  engineers  had  mounted  small  models  of 
the  Mercury  capsule  on  the  tips  of  research  rockets,  launched  them  through 
the  complete  speed  range  predicted  for  the  proposed  spaceshot,  and  collected 
thousands  of  data  points  about  the  capsule's  structural  integrity,  tumbling 
characteristics,  and  reentry  dynamics.  With  the  military's  assistance, 
Langley  researchers  also  tested  the  reliability  of  the  capsule  parachute 
system  and  determined  the  optimum  altitude  at  which  to  deploy  the  drogue 
chute.  Prom  a  C-130  Hercules  transport  that  had  been  loaned  to  NASA 
by  the  U.S.  Air  Force  Tactical  Air  Command  at  Langley  Field,  full-scale, 
one-ton  models  of  the  Mercury  capsule  prepared  at  Langley  were  dropped 
from  an  altitude  of  10,000  feet  into  the  Atlantic  Ocean  off  Wallops  Island. 
Motion  pictures  from  cameras  in  T-33  chase  jets  were  used  to  make  a  detailed 
engineering  study  of  the  capsule's  motions  during  descent  and  the  impact 
forces  on  it  when  smacking  into  the  sea.  Langley  personnel  also  conducted 
other  impact  studies  by  dropping  small  models  of  the  space  capsule  at  30 
feet  per  second  (21.6  miles  per  hour)  into  the  Hydrodynamics  Division's 
Water  Tank  No.  I.26 

While  the  numerous  aerodynamic,  structural,  materials,  and  component 
tests  were  going  on  at  the  center,  Langley  representatives  were  arranging 
a  schedule  for  wind-tunnel  tests  at  the  air  force's  Arnold  Engineering 
Development  Center  in  Tullahoma,  Tennessee,  and  a  team  of  non-STG  staff 
members  was  being  assembled  to  travel  around  the  world  to  plan  Project 
Mercury's  global  tracking  network,  the  responsibility  for  which  NASA 
headquarters  had  just  assigned  to  the  research  center  at  the  STG's  request 
in  February  1959.  In  addition  to  this  colossal  effort,  Langley  engineers 
and  technicians  were  developing  the  simulators  and  spaceflight  procedure 
trainers  for  the  Mercury  astronauts  who  had  just  been  entrusted  to  the  STG. 
By  opening  day  of  the  NASA  inspection  in  October  1959,  Langley  had  sent 
six  months'  worth  of  weekly  reports  to  NASA  headquarters  about  the  great 
volume  of  work  being  done  in  support  of  Mercury.  Of  the  laboratory's  1150 
employees,  119  of  them  (about  10  percent)  had  been  working  full-time  on 
the  project  in  recent  months. 

In  the  year  following  the  STG's  establishment,  between  October  1958 
and  October  1959,  some  250  people  were  added  to  the  original  STG; 
more  than  half  came  from  Langley's  staff.  Many  of  the  key  people  who 
moved  from  Langley  to  the  STG  brought  with  them  important  experience 
in  flight-test  research.  Floyd  Thompson  wanted  to  give  Gilruth  a  strong 

60 


Carrying  Out  the  Task 


L-59-336 


L-59-3397 

The  Mercury  spacecraft  and  booster  rockets  underwent  extensive  testing  in  Langley 
wind  tunnels.  The  full-size  capsule  is  mounted  in  the  Full-Scale  Tunnel  (top).  A 
one-sixth  scale  model  of  the  Mercury  capsule  is  tested  in  Langley 's  7  x  10- Foot 
High-Speed  Tunnel  to  determine  the  effect  of  escape  system  power  on  the  capsule 's 
stability  (bottom). 

61 


Spaceflight  Revolution 


L-59-4335 

The  Redstone  booster  carrying  the  spacecraft  is  mounted  for  testing  in  Langley's 
Unitary  Plan  Wind  Tunnel. 


L-60-76 


In  impact  studies  conducted  in  the  Back  River  behind  Langley's  East  Area,  the 
astronauts  practiced  the  dangerous  maneuver  of  getting  out  of  the  space  capsule  as 
it  floated  in  water. 


62 


Carrying  Out  the  Task 

cohort  that  understood  "flying  men" — pilots,  that  is — not  just  the  flying 
of  pilotless  models.  "Tommy  wanted  to  make  sure  that  there  were  enough 
flight  guys  involved  in  this  venture,"  Donlan  remembers.27  Fortuitously, 
NASA  headquarters  recently  had  made  a  decision  to  limit  Langley's  flying 
and  had  transferred  most  of  its  flight  research  activities  to  the  NASA 
center  at  Edwards  AFB.  This  decision  disappointed  Langley  researchers  and 
made  them  ready  to  jump  at  the  chance  to  get  involved  with  the  manned 
space  program.  Consequently,  several  top-notch  Langley  flight  researchers 
became  part  of  the  STG.  Along  with  Gilruth  (also  a  former  NACA  flight 
research  engineer),  Walter  C.  Williams,  former  director  of  the  NACA  Flight 
Research  Center  in  California,  and  Christopher  C.  Kraft,  Jr.,  and  Charles 
W.  Mat  hews,  both  standouts  in  Langley's  Flight  Research  Division,  became 
the  heart  of  the  Project  Mercury  flight  operations  team. 


The  Tracking  Range 

Of  all  the  Langley  efforts  in  support  of  Project  Mercury,  by  far  the 
biggest,  the  most  difficult  to  carry  out  logistically,  and  the  most  adven- 
turesome was  the  Mercury  tracking  range  project.  NASA  flight  operations 
officers  and  aeromedical  specialists  wanted  to  have  almost  constant  radio 
contact  with  the  Mercury  astronauts.  To  maintain  communication  with  the 
spacecraft  as  it  circled  the  earth,  NASA  had  to  create  a  worldwide  commu- 
nications and  tracking  network. 

In  the  early  days  of  Project  Mercury,  NASA  really  did  not  know  what 
sort  of  tracking  network  was  needed  to  monitor  its  spacecraft.  Those  frontier 
days  of  the  manned  space  program  before  the  operation,  let  alone  the  very 
idea,  of  a  "mission  control"  center  are  hard  to  remember.  Over  the  last  three 
decades,  the  public  has  grown  familiar  with  the  drama  and  the  emotionally 
charged  "electricity"  of  the  control  center  amphitheater.  This  amphitheater, 
with  its  tidy  rows  of  communications  consoles,  computerized  workstations, 
and  its  front  wall  covered  with  a  large  electronic  map  of  the  world,  became 
thought  of  as  the  brain  and  nerve  center  of  a  NASA  spaceflight  mission. 
Here,  in  what  one  NASA  astronaut  has  called  a  "temple  of  technology," 
worked  the  middle-aged  men  in  white  shirts  and  dark  neckties — the  flight 
controllers  who  wore  the  headphones  and  the  worried  looks  as  they  talked 
to  the  astronauts  in  the  spacecraft  and  made  the  split-second,  life-or-death 
decisions  about  whether  to  "abort"  or  "go  for  orbit."28 

This  stage  for  the  high  drama  of  "space  theater"  did  not  exist  prior 
to  Project  Mercury.  The  flight  tests  of  the  most  experimental,  high-speed 
airplane  had  not  required  the  development  of  a  ground-control  facility  as 
sophisticated  as  mission  control.  Even  at  a  pioneering  place  like  Edwards 
AFB,  the  role  of  the  flight  experts  on  the  ground  had  involved  little  more 
than  "getting  the  airplane  into  the  best  possible  mechanical  condition, 
spelling  out  the  day's  test  objectives  for  the  pilot,  and  retrieving  data  from 

63 


Space/light  Revolution 

the  instrumentation  after  the  plane  landed."29  During  the  flight  itself,  flight 
operations  people  talked  to  the  pilot  in  moderation;  for  the  most  part,  they 
quelled  their  curiosity,  shaded  their  eyes,  strained  anxiously  to  follow  the 
flashing  metal  arrow  through  the  sky,  and  left  the  pilot  to  his  own  devices. 

At  first,  the  STG  envisioned  little  more  than  this  rather  passive  mode 
of  flight  control  for  the  Mercury  spacecraft:  checking  it  out  before  launch, 
maintaining  a  voice  link  with  the  astronaut  to  see  how  things  were  going,  but 
letting  the  astronaut  and  the  automatic  in-flight  systems  do  the  rest.  After 
reflecting  seriously  on  the  immense  task  before  them,  that  vision  changed. 
"I  don't  know  how  to  describe  it  exactly,"  explains  Glynn  Lunney  of  the 
original  STG,  "but  we  began  to  realize  that,  'Hey,  we're  going  to  fly  this 
thing  around  the  world!'  '  In  that  instant  of  stark  realization  came  the 
feeling  that  certain  critical  decisions  about  a  spaceflight — such  as  whether 
to  abort  immediately  after  launch,  to  use  the  escape  rocket,  or  to  blow  up 
a  maverick  rocket  before  it  dug  a  big  hole  into  downtown  Cocoa  Beach — 
could  be,  and  should  be,  controlled  from  the  ground.  Out  of  this  conviction 
came  the  concept  of  a  ground  room  with  not  just  a  person  talking  to  the 
astronaut,  but  many  people  analyzing  tracking  and  telemetry  data  on  the 
status  of  the  launch  vehicle  and  the  spacecraft.30  Already  by  the  time  of 
the  first  NASA  inspection  in  October  1959,  the  STG  was  calling  this  room 
the  Mercury  "Control  Center"  and  was  moving  rapidly  to  have  one  built  at 
Cape  Canaveral. 

As  the  vague  and  open-ended  possibilities  of  Mercury  flight  operations 
and  mission  control  became  more  clearly  defined,  the  STG  decided  that  to  be 
out  of  communication  with  the  astronauts  during  their  spaceflights  for  very 
long  would  be  neither  wise  nor  safe.  The  STG's  flight  operations  people 
and  more  conservative  aeromedical  specialists  argued  over  the  maximum 
amount  of  time  they  could  be  out  of  contact  with  the  astronauts.  The 
physicians  were  "horrified  at  the  casualness"  of  one  suggestion  that  in-flight 
communications  with  the  astronauts  could  be  handled  like  commercial  air 
traffic  control,  with  the  pilot  only  reporting  to  the  ground  every  15  to  30 
minutes.31  The  doctors,  intent  on  continuous  and  complete  monitoring  of 
the  astronaut's  vital  physiological  and  mental  responses  to  the  unknown 
demands  of  spaceflight,  did  not  like  the  idea  of  gaps  in  communication 
lasting  for  any  appreciable  length  of  time.  Without  the  resolution  of  this 
internal  debate,  engineers  could  not  establish  design  parameters  necessary 
for  proceeding  with  the  global  tracking  network.  In  the  end,  the  STG 
decided  that  a  tracking  network  was  needed  in  which  gaps  in  communication 
lasted  no  more  than  10  minutes.32 

Fathoming  the  immensity  of  what  had  to  be  done  to  establish  this 
network  took  time.  Initially  some  naive  Langley  engineers  believed  that 
whatever  tracking  stations  were  needed  by  the  Mercury  team  to  provide 
"real-time"  tracking  data  could  be  provided  simply  by  mounting  radar  sets 
on  rented  air  force  trucks  that  could  be  stationed  at  sites  around  the  world. 
But  after  giving  the  matter  careful  thought,  the  communications  experts 

64 


Carrying  Out  the  Task 

"began  to  realize  that  it  wasn't  good  enough  to  have  isolated  radar  sets: 
the  people  back  at  the  Control  Center  needed  a  network  of  linked  stations, 
capable  of  receiving,  processing,  and  reacting  to  a  variety  of  voice,  radar, 
and  telemetry  data."33 

Thus  began  a  Promethean  task  because  1960  was  a  different  technologi- 
cal age — especially  in  terms  of  communications.  An  instantaneous  telephone 
call  around  the  world  was  not  yet  possible.  The  only  long-range  communi- 
cation, from  continent  to  continent,  was  by  undersea  telegraph  cable,  and 
most  of  these  cables  had  been  laid  at  the  turn  of  the  century  by  the  British. 
That  is  not  to  say  a  remarkable  telecommunications  network  did  not  exist. 
Over  the  years  the  British,  among  others,  had  built  up  an  amazing  global 
system  involving  tens  of  thousands  of  miles  of  submarine  cables  as  well  as 
vast  distances  covered  by  wireless  communications,  but  the  day  of  instan- 
taneous electronic  communication  around  the  world  had  not  yet  arrived. 
Its  arrival  depended  largely  on  the  launch  of  communications  satellites  like 
Telstar,  which  the  infant  space  programs  at  that  time  were  making  possible. 
For  NASA  staff  to  have  the  type  of  communications  necessary  for  control 
of  the  Mercury  spacecraft  and  for  assistance  to  the  astronauts,  they  had  to 
build  their  own  global  system. 

Creating  this  global  network  was  a  job  that  NASA  Goddard  Research 
Center  could  not  do  from  its  temporary  quarters  at  Anacostia.  Also, 
Goddard  people  were  still  responsible  for  the  Minitrack  Network  that  had 
been  set  up  for  the  Project  Vanguard  satellite,  so  they  were  busy  tracking 
the  unmanned  satellites  that  were  then  being  launched.  This  existing  system 
was  not  suitable  for  tracking  the  orbit  of  the  Mercury  spacecraft  because 
the  system  had  been  laid  out  north-to-south  (along  the  75th  meridian), 
whereas  STG  studies  had  concluded  that  the  best  orbital  path  of  the 
Mercury  spacecraft  would  be  west-to-east  along  the  equator.  Minitrack, 
even  in  combination  with  other  existing  commercial,  scientific,  and  military 
communications  networks,  had  far  too  many  "bare  spots"  to  provide  the 
comprehensive  global  coverage  required  for  Mercury. 

The  STG  was  unable  to  take  on  this  job  because  its  manpower  was 
already  stretched  to  the  limit;  STG  staff  could  not  bear  the  additional  load 
of  setting  up  an  ambitious  new  tracking  and  communications  net  that  had  to 
reach  completely  around  the  world.  "There  was  just  no  way  [for  the  STG]  to 
build  the  spacecraft  as  well  as  the  ground  tracking  network,"  says  William 
J.  Boyer,  the  fellow  from  Langley's  IRD  whose  transfer  to  the  original  STG 
had  been  short-circuited  by  his  division  chief  in  November  1958.  Boyer, 
who  became  one  of  the  most  active  members  of  the  Langley  team  that 
built  the  Mercury  tracking  range,  remembers  that  Howard  Kyle,  the  IRD 
engineer  who  was  named  to  replace  him  on  the  STG,  was  the  first  to  come 
to  this  conclusion.  Kyle,  without  any  trouble,  persuaded  STG's  Chuck 
Mathews  of  the  impossibility;  Mathews  in  turn  convinced  Bob  Gilruth;  and 
Gilruth  asked  Floyd  Thompson  whether  Langley,  with  NASA  headquarters' 
approval,  could  take  on  this  additional  heavy  responsibility.35 

65 


Spaceflight  Revolution 

Once  again,  Thompson  wanted  to  do  everything  he  could  to  make  Project 
Mercury  a  success.  So  in  February  1959,  he  called  in  his  assistant  director, 
Hartley  Soule,  and  they  put  together  an  ad  hoc  team  that  came  to  be 
known  as  TAGIU  (pronounced  "Taggy-you" ) ,  which  stood  for  the  Tracking 
and  Ground  Instrumentation  Unit.  Heading  the  temporary  unit  was  Soule 
himself,  who  was  deemed  the  tracking  range  project  director.  G.  Barry 
Graves,  Jr.,  the  head  of  IRD's  Pilotless  Aircraft  Research  Instrumentation 
Branch,  was  to  handle  the  detailed  management  of  the  tracking  network 
project  from  a  special  TAGIU  office,  and  Paul  H.  Vavra,  Graves's  colleague 
in  the  IRD  branch,  was  to  assist.  The  unit  was  placed  within  IRD  on 
an  organizational  chart.  No  one  really  knew  how  much  work  faced  them: 
members  of  TAGIU  were  told  initially  that  their  work  would  be  part-time 
and  add  only  slightly  to  their  regular  duties.  But  as  Vavra  notes,  "a 
few  weeks  later  we  were  in  the  space  program  night  and  day  and  never 
thought  about  our  other  jobs."36  As  with  everything  else  concerning  Project 
Mercury,  TAGIU  progressed  rapidly.  On  30  July  1959,  NASA  awarded  the 
contract  for  the  creation  of  an  integrated  spacecraft  tracking  and  ground 
instrumentation  system  to  Western  Electric  Company  and  its  four  major 
subcontractors:  Bell  Telephone  Laboratories  of  Whippany,  New  Jersey,  for 
system  engineering,  engineering  consultations,  and  command  and  control 
displays;  the  Bendix  Corporation  of  Los  Angeles  and  Towson,  Maryland,  for 
radar  installation,  ground-to-air  communications,  telemetry,  and  site  display 
equipment;  Burns  and  Roe  of  Long  Island  for  site  preparation,  site  facilities, 
construction,  and  logistic  support;  and  International  Business  Machines 
Corporation  of  New  York  for  computer  programming,  simulation  displays, 
and  computers.37  Monitoring  the  contract  involved  the  expenditure  of 
nearly  $80  million  and  extensive  negotiation  with  other  federal  agencies, 
private  industry,  and  representatives  of  several  foreign  countries.  However, 
in  June  1961,  less  than  two  years  after  awarding  the  contract,  Langley 
looked  on  with  pride  as  the  power  for  the  around-the-world-in-an-instant 
communications  system  was  turned  on  for  the  first  time. 

Working  on  the  global  tracking  range  took  Langley  personnel  farther 
away  from  the  comfortable  confines  of  their  wind  tunnels  than  any  other 
aerospace  project  ever  had  before,  or  has  since.  In  the  two-year  period 
between  the  awarding  of  the  contract  and  the  initiation  of  the  tracking 
operations,  a  team  of  engineers  and  technicians  from  NASA  Langley  traveled 
tens  of  thousands  of  miles  to  some  of  the  most  remote  places  on  earth.  They 
went  to  oversee  the  building  of  an  ambitious  network  that  when  completed 
stretched  from  the  new  Mercury  Control  Center  at  Cape  Canaveral  to  18 
relay  stations  spanning  three  continents,  seven  islands,  and  two  ocean-bound 
radar  picket  ships.  Along  its  way  around  the  world,  the  network  utilized  land 
lines,  undersea  cables  and  radio  circuits,  special  computer  programs  and 
digital  data  conversion  and  processing  equipment,  as  well  as  other  special 
communications  equipment  installed  at  commercial  switching  stations  in 
both  the  Eastern  and  Western  hemispheres.  The  network  involved  range 

66 


Carrying  Out  the  Task 


George  Barry  Graves,  Jr.,  head  of  the 
Pilotless  Aircraft  Research  Instrumenta- 
tion Branch  of  Langley's  IRD,  handled 
the  detailed  management  of  the  Mercury 
tracking  network  from  a  special  office 
within  the  ad  hoc  TAGIU. 
L-61-659 


stations  in  such  faraway  and  inaccessible  places  as  the  south  side  of  the 
Grand  Canary  Island,  120  miles  west  of  the  African  coast;  Kano,  Nigeria,  in 
a  farming  area  about  700  rail-miles  inland;  Zanzibar,  an  island  12  miles  off 
the  African  coast  in  the  Indian  Ocean;  a  place  called  Woomera,  amid  the 
opal  mines  in  the  middle  of  the  Australian  outback;  and  Canton  Island,  a 
small  atoll  about  halfway  between  Hawaii  and  Australia. 

"It  was  quite  mind-boggling  to  realize  that  you're  living  in  Hampton, 
Virginia,  and  you  were  getting  tickets  to  change  planes  in  the  Belgian  Congo 
to  go  to  Kenya  and  from  there  on  to  Zanzibar,"  exclaims  Bill  Boyer.  Boyer 
traveled  with  Barry  Graves's  small  "management  team,"  which  negotiated 
with  foreign  governments  and  picked  the  tentative  sites  for  the  Mercury 
tracking  stations.  In  Madrid  his  team  sat  for  four  weeks  waiting  for  the 
Spanish  government  to  grant  permission  to  go  to  the  Canary  Islands.  On 
their  way  through  central  Africa,  in  the  Belgian  Congo,  group  members 
moved  cautiously  past  threatening  gun-toting  rebels  who  were  fighting 
against  European  colonial  rule.  "We  would  pick  the  tentative  sites  based  on 
the  technical  criteria  established  by  the  Space  Task  Group,"  Boyer  states, 
"and  then  we'd  go  around  to  the  telecommunications  people  in  those  foreign 
countries  to  get  as  much  advice  and  assistance  from  them  as  we  could." 
The  TAGIU  team  looked  into  the  logistics  of  particular  sites:  Where  would 
NASA  people  eat?  Where  would  they  sleep?  How  would  they  be  supplied? 
What  were  the  capabilities  of  the  local  construction  companies?  After 
addressing  these  questions,  the  management  team  would  move  on,  and  a 

67 


Spaceflight  Revolution 


WPfCAL 


AND 
ACQUISITION  AID 


L-61-5740 

^4  layout  of  a  typical  Project  Mercury  tracking  site  as  conceptualized  by  Graves's 
outfit  in  1961. 


"technical  team"  would  move  into  the  recommended  site.  This  larger,  follow- 
on  team  would  then  conduct  a  detailed  study  to  determine  whether  the  site 
met  technical  criteria:  could  NASA  construct  the  buildings  it  needed,  and 
were  the  materials  easily  available?  The  technical  team  would  then  make  a 
final  recommendation  about  the  proposed  site.38 

As  with  so  many  other  rushed  and  complicated  operations  of  the  early 
manned  space  program,  much  about  the  multimillion-dollar  Mercury  track- 
ing network  could  have  gone  wrong.  Instead,  it  worked  like  a  charm,  track- 
ing the  spacecraft  with  a  high  degree  of  accuracy.  In  the  words  of  Edmond 
C.  Buckley,  the  former  IRD  head  at  Langley  who  by  the  time  of  the  first 
Mercury  orbital  flight  by  John  Glenn  was  the  director  of  tracking  and  data 
acquisition  at  NASA  headquarters,  the  network  "worked  better  than  it  could 
have  in  the  most  optimistic  dreams."39  For  example,  as  the  system  tracked 
the  spacecraft  from  the  Bermuda  station  on,  NASA  found  that  the  "residu- 
als," that  is,  the  comparison  of  the  computed  predicted  path  and  the  actual 
path  as  determined  by  each  location,  differed  in  most  cases  by  less  than  1000 

68 


Carrying  Out  the  Task 

feet  and  in  some  cases  by  less  than  100  feet.  These  figures  compared  favor- 
ably with  the  ability  of  tracking  systems  of  that  day  to  report  the  location 
of  naval  ships  crossing  the  oceans. 

The  creation  of  this  unprecedented  and  highly  successful  worldwide 
ground  instrumentation  and  tracking  network  required  the  services  of  many 
members  of  the  Langley  staff  beyond  those  formally  part  of  TAGIU.  Three 
Langley  organizations  (as  well  as  several  outfits  at  Wallops  Island)  played 
major  roles  in  establishing  the  network:  IRD,  which  helped  to  guide  the 
design  of  the  electronic  systems;  the  Engineering  Service  Division,  which 
assisted  in  the  selection  of  sites  and  the  coordination  and  monitoring  of  the 
station  construction;  and  the  Procurement  Division,  which  negotiated  the 
huge  contract  and  maintained  constant  liaison  with  the  prime  contractor, 
Western  Electric,  and  its  associates.  Thanks  to  this  extensive  effort, 
NASA  was  able  to  have  the  kind  of  direct  and  comprehensive  contact 
with  the  astronauts  and  their  spacecraft  that  the  flight  operations  and 
medical  experts  believed  was  necessary.  As  Edmond  Buckley  remarked,  in  a 
masterpiece  of  understatement,  Langley  "can  take  a  well-deserved  bow."40 


Shouldering  the  Burden 

Nothing  was  more  important  to  the  stated  objectives  of  the  American 
space  program  by  the  early  1960s  than  Project  Mercury,  but  supporting  the 
program  was  still  a  burden  on  Langley  Research  Center.  Gilruth  admits 
that  the  days  of  a  rapidly  expanding  Mercury  program  must  have  been 
"particularly  difficult  for  Langley"  because  Gilruth's  need  for  good  people 
was  such  that  he  "could  not  help  but  continue  to  recruit"  from  the  center. 
Faced  with  Gilruth's  personnel  demands,  Thompson  bargained  with  him. 
"Okay,  Bob.  I  don't  mind  letting  you  have  as  many  good  people  from 
Langley  as  you  need  . . .  but  for  every  one  that  you  want  to  take  . . .  you  must 
also  take  one  that  I  want  you  to  take."  Prom  that  day,  whenever  Gilruth 
recruited  a  person  for  the  STG,  he  also  took  a  person  that  Thompson  was, 
for  one  reason  or  another,  eager  to  transfer.41 

Thompson  became  the  center  director  in  May  1960,  and  Henry  Reid 
moved  on  to  become  his  titular  senior  adviser.  Aware  that  certain  Langley 
staff  members  were  not  productive  in  their  present  positions,  the  crafty 
Thompson  wanted  to  make  room  in  his  organization  for  some  new  blood. 
Langley  had  found  ways  to  make  room  in  the  past,  notably  in  the  1940s 
when  several  wagonloads  of  its  people  had  moved  west  to  colonize  the 
newly  created  NACA  centers  in  Ohio  and  California.  The  founding  of  new 
laboratories  such  as  Ames  and  Lewis,  and  now  the  STG,  enabled  the  center 
director  to  transfer  out  restless  souls  and  nonproductive  old-timers  along 
with  the  people  who  were  crucial  to  the  success  of  the  new  operation.  These 
transfers  allowed  for  the  influx  of  fresh  and  dynamic  young  people  that 
Langley  continually  needed  to  remain  a  productive  laboratory. 

69 


Spaceflight  Revolution 

While  Langley's  support  for  Project  Mercury  continued  to  expand,  so  too 
did  the  size  and  experience  of  the  STG.  With  Langley's  help,  the  STG's 
capacity  for  handling  its  own  technical  and  administrative  affairs  increased 
dramatically.  By  the  time  Thompson  officially  became  the  director,  he 
and  his  senior  staff  recognized  that  Langley's  ad  hoc  parental  role  in  the 
Mercury  program  needed  further  definition.  According  to  Thompson,  the 
time  had  come  "to  replace  the  informal  free-wheeling  and  somewhat  chaotic 
working  arrangements  with  orderly  procedures."  A  formalizing  of  relations 
was  needed  to  "clearly  identify  the  respective  responsibilities  of  the  two 
organizations"  and  to  establish  more  distinct  channels  for  authorizing  and 
conducting  business.  Otherwise,  too  many  more  of  Langley's  own  precious 
capabilities  would  be  carved  off  for  the  STG.42 

But  Thompson's  thoughts  about  Langley's  proper  relationship  to  the 
STG  were  ambivalent.  On  the  one  hand,  a  voice  within  Thompson  told 
him  to  follow  the  advice  that  Hugh  Dryden  had  been  giving  him  about 
Project  Mercury:  "Support  it,  but  don't  let  it  eat  you  up."  By  that  Dryden 
meant  that  the  director  of  a  research  laboratory  should  not  neglect  his 
basic  research  programs  because  of  the  center's  appetite  for  any  one  big 
project,  however  delectable  it  might  seem.*  As  soon  as  possible,  Dryden 
warned,  the  STG  needed  to  become  part  of  a  laboratory  devoted  just  to 
spaceflight  development.  Dryden  knew  that  in  a  technical  environment 
where  a  "research  function"  and  a  "development  function"  tried  to  coexist, 
the  development  function  would  always  win  out  (as  it  would  later  do  when 
Langley  managed  the  Viking  project).  If  Langley  kept  the  STG,  Dryden 
worried,  the  center  would  inevitably  lose  many  of  its  most  capable  people 
to  development.  Without  its  expertise  in  research,  NASA  would  turn  into  a 
shadow  of  its  former  self  and  something  less  than  what  the  country  needed 
it  to  be. 

Moreover,  Thompson  was  plagued  by  some  troubling  questions:  What 
happens  when  "the  development"  reaches  completion?  How  are  the  "devel- 
opment people"  brought  back  effectively  into  the  general  research  program, 
or  do  these  people  just  continue  to  look  for  things  to  develop?  The  only  way 
to  truly  ensure  the  priority  of  the  center's  research  function  was  to  move 
the  STG  away  from  Langley  completely,  but  by  the  early  1960s  so  much  of 
NASA  Langley's  identity  was  tied  up  with  the  success  of  Project  Mercury 
and  the  publicity  glow  surrounding  its  astronauts  that  Thompson  and  oth- 
ers at  Langley  were  not  at  all  sure  they  wanted  to  lose  the  STG  to  some 
other  facility.  The  STG  was  so  important  to  the  national  mission,  so  many 
resources  were  being  devoted  to  it,  and  the  American  public  was  becoming 
so  fascinated  with  astronauts  and  the  prospect  of  manned  spaceflight,  that 
even  the  most  clearheaded  researchers  at  Langley  were  turning  a  little  misty 
over  the  center's  involvement  in  Project  Mercury.  At  Langley  the  number 


?k 

Although  Hugh  Dryden  supported  Project  Mercury,  he  was  in  truth  no  great  fan  of  the  emphasis 
NASA  placed  on  it. 

70 


Carrying  Out  the  Task 

of  "envious  people  who  didn't  want  to  leave  their  own  jobs  but  who  liked  to 
bask  in  the  [STG's]  limelight"  was  growing.43  Mercury  was  a  mushrooming 
project  that  was  suddenly  making  national,  even  international,  news.  The 
local  press  was  sending  reporters  out  regularly  to  the  center — something 
that  had  never  happened  before.  The  attention  was  a  lot  to  lose. 

Thompson  was  less  alarmed  by  the  risks  of  supporting  the  STG  and 
Project  Mercury  than  Dry  den,  although  he  claims  to  have  understood  them 
well.  Thompson  was  willing  to  gamble  that  the  STG  would  help  Langley 
more  than  harm  it.  In  the  long  run,  Thompson  argued,  "the  broader 
demands  imposed  by  a  space  program  added  to  an  existing  aeronautics 
program"  would  make  the  research  role  more  important  to  the  country  than 
it  had  ever  been  before.  To  carry  out  the  space  program  while  continuing  to 
stimulate  the  aircraft  industry  and  support  commercial  and  military  aviation 
required  more  fundamental  research,  not  less.44 

A  voice  inside  Thompson  told  him  that  the  STG  should  become  an  official 
part  of  Langley;  into  the  early  1960s,  this  voice  of  aggrandizement,  not 
Dryden's  of  caution,  dominated  much  of  Thompson's  thinking  and  some 
of  his  behind-the-scenes  activities  and  management  decisions  pertaining  to 
Project  Mercury.  "He  wanted  to  combine  the  STG  with  Langley  and  have 
Langley  manage  it,"  recalls  Laurence  K.  Loftin,  Jr.,  one  of  Thompson's 
closest  associates  from  the  time.  "He  wanted  to  run  the  whole  damn 
thing."45 

However,  in  a  research  culture  with  deep  NACA  roots  like  Langley's, 
not  everyone  felt  that  supporting  the  STG  was  an  acceptable  risk.  These 
feelings  were  reflected  in  such  mundane  matters  as  board  hearings  about 
promotions.  Originally  the  STG  went  through  the  regular  Langley  board 
for  promotions,  but  some  STG  members  felt  "they  didn't  get  a  fair  deal" 
that  way.  For  example,  candidates  for  promotions  who  had  done  jobs  such  as 
the  preparation  of  Mercury  training  manuals  were  "considered  unfavorably" 
by  Langley  people  who  felt  that  the  production  of  a  traditional  research 
report  was  a  much  more  important  achievement.  Feelings  about  this  "unfair 
treatment"  eventually  grew  so  strong  that  the  STG  decided  to  create  its 
own  promotions  board  to  sidestep  those  at  Langley  who  felt  that  writing 
training  manuals  amounted  to  "clambake  work"  and  was  not  "worth  that 
kind  of  money."46 

Funding  was  at  the  root  of  some  of  the  senior  staff's  concerns.  They 
worried  that  the  STG  might  absorb  so  much  of  the  center's  research 
capability  that  NASA  headquarters  would  reduce  its  support  for  Langley's 
independent  research  function.  The  tail  would  start  wagging  the  dog. 
Most  members  of  the  STG  were  too  busy,  ambitious,  or  imprudent  to 
discourage  this  notion.  Some  STG  members  believed  that  they  would 
continue  conducting  research  while  proceeding  with  Project  Mercury.  If  that 
happened,  some  at  Langley  worried,  NASA's  and  the  country's  support  for 
independently  funded  research  at  the  center  might  be  badly,  and  perhaps 


71 


Spaceflight  Revolution 

even  fatally,  compromised.  Langley  might  turn  into  a  place  that  handled 
big  projects  while  remaining  no  more  than  semiactive  in  research.47 

Only  very  gradually  and  reluctantly  did  Langley  management  and  the 
conflicted  Floyd  Thompson  come  to  feel  that  something  had  to  be  done 
to  cut  the  apron  strings  that  connected  Langley  to  the  STG.  Certain 
productive  steps  were  taken  by  NASA  headquarters  in  1959  and  1960  to 
strengthen  the  STG's  own  organization  and  management  and  reduce  its 
dependence  on  Langley  for  administrative  and  technical  support.48 

One  of  the  steps  taken  to  distinguish  the  STG  operation  from  that  of 
the  larger  research  center  simply  involved  office  space  and  physical  facilities. 
Pressed  for  space,  Langley  had  assigned  the  STG  initially  to  the  second  floor 
of  the  Unitary  Plan  Wind  Tunnel  building  in  the  West  Area.  But  before 
long,  Langley  relocated  Gilruth  and  his  staff  to  facilities  in  the  East  Area. 
Two  factors  behind  the  move  were  the  need  to  expand  and  the  desire  to 
find  a  cluster  of  offices  where  the  growing  STG  could  work  as  a  consolidated 
team,  but  a  third  seems  to  have  been  the  prejudice  of  the  Langley  senior 
staff  against  locating  research  and  development  functions  so  close  together 
within  the  confines  of  the  same  center. 

In  the  East  Area,  the  STG  went  to  work  inside  two  of  the  oldest  buildings 
at  the  center;  they  had  been  constructed  nearly  40  years  earlier,  before  the 
laboratory's  formal  opening  in  1920.  Building  104  (later  renumbered  NASA 
no.  586)  was  the  old  Technical  Services  building;  to  make  room  for  the 
STG,  some  of  Langley 's  systems  and  equipment  engineering  people  had  to 
vacate  their  dusty  premises.  Building  58  (later  renumbered  NASA  no.  587) 
had  served  as  Langley's  main  office  from  1920  until  the  new  headquarters 
building  opened  in  the  West  Area  right  after  World  War  II;  in  the  center 
telephone  directory,  this  once  important  building  on  Dodd  Boulevard,  the 
former  home  of  Langley's  engineer-in-charge,  was  still  referred  to  as  the 
Administration  building.  In  1959  the  sturdy  two-story,  red-brick  structure 
housed  the  East  Area's  cafeteria,  a  group  of  stenographers,  the  center's 
editorial  division,  as  well  as  most  of  the  personnel,  employment,  and 
insurance  offices.  To  accommodate  members  of  the  STG,  some  but  not 
all  of  these  office  operations  were  moved  to  buildings  in  the  West  Area. 

The  rapidly  expanding  STG  eventually  took  over  most  of  NASA's 
buildings  in  the  East  Area,  as  well  as  several  adjacent  air  force  facilities.  But 
the  STG  remained  hungry  for  space.  Langley  management  had  to  release  a 
few  buildings  in  the  West  Area  for  STG  use.  For  instance,  Building  1244 
became  a  staging  area  where  technicians  refurbished  the  boilerplate  capsules 
that  were  used  for  drop  tests  in  the  nearby  Back  River,  and  Building  1232 
was  turned  into  an  STG  fabrication  shop  where  prototype  capsules  were 
inspected  and  assembled. 

Members  of  the  STG  did  not  complain  much  about  the  patchwork  nature 
of  these  quarters  because  the  group  was  housed  at  Langley  only  temporarily, 
pending  transfer  to  a  permanent  base  of  operation.  Abe  Silverstein,  the 
head  of  the  Office  of  Space  Flight  Development,  planned  to  move  the  STG 

72 


Carrying  Out  the  Task 

to  Goddard  when  the  facility  for  the  new  spaceflight  center  in  Greenbelt, 
Maryland,  was  completed.  Although  located  at  Langley,  the  STG  had  been 
reporting  directly  to  Silverstein's  office  in  Washington,  but  this  arrangement, 
like  housing  the  group  at  Langley,  was  a  temporary  expedient  until  a  more 
permanent  arrangement  could  be  established.49 

The  management  logic  behind  the  transfer  of  the  STG  into  the  Goddard 
organization  came  from  Silverstein:  a  focused  little  organization  like  the 
STG  might  be  capable  of  running  the  technical  part  of  its  operation, 
but  in  terms  of  handling  budgetary  matters,  looking  after  swelling  fiscal 
and  procurement  responsibilities,  and  supplying  material  and  housekeeping 
support,  the  STG  needed  all  the  help  it  could  get.  NASA  did  not  have 
the  resources  to  build  a  complete  organization  around  a  solitary  task  force 
carrying  out  a  single  project,  no  matter  how  important  the  project.  It 
made  more  sense  to  place  the  task  force  inside  an  existing  organization 
already  having  a  complete  range  of  capabilities — but  not  as  overburdened 
with  responsibilities  as  Langley. 

Most  members  of  the  STG  disliked  Silverstein's  plan.  They  did  not 
want  to  move  to  a  suburb  of  the  nation's  congested  capital  city,  and  they 
were  a  little  bitter  over  what  they  viewed  as  a  lack  of  appreciation  for  the 
magnitude  of  their  work.  The  manned  spaceflight  program  would  be  only 
one  of  several  projects  at  the  new  Goddard  center.  If  Gilruth  and  the  rest 
of  his  STG  could  have  had  their  way,  they  would  have  preferred  to  stay 
at  Langley  and  continue  the  close  relationship  with  the  center  that  both 
sides  had  found  workable  from  late  1958  on.  In  spite  of  the  heavy  drain 
on  his  center's  manpower  and  facilities  and  the  justifiable  fears  about  what 
such  a  big  space  project  might  do  to  divert  and  distort  essential  research 
capability  strengths,  Floyd  Thompson  ultimately  would  have  preferred  to 
keep  the  entire  manned  space  program  at  Langley.  Such  were  the  personal 
and  institutional  temptations  that  came  with  the  spaceflight  revolution  and 
its  "big  technology." 

The  STG,  however,  was  made  formally  a  part  of  Goddard  on  1  May  1959, 
which  was  Goddard's  official  opening  day.  Although  still  housed  at  Langley 
and  separated  from  the  new  spaceflight  center  by  more  than  100  miles  of 
Tidewater  Virginia,  the  STG  became  the  Manned  Spacecraft  Division  of 
Goddard,  with  Gilruth  serving  as  the  new  center's  assistant  director  for 
manned  satellites  while  remaining  the  director  of  Project  Mercury. 

In  the  beginning,  everyone  had  thought  that  Bob  Gilruth  would  be  the 
director  of  Goddard  and  that  the  new  space  center  would  be  not  only  the 
place  for  manned  spaceflight  but  also  for  all  of  NASA's  space  science  activity. 
As  Charles  Donlan  remembers,  "When  Dr.  Dryden  gave  Gilruth  his  first 
title,  it  wasn't  'Director  of  Project  Mercury,'  it  was  'Assistant  Director 
of  Goddard.'  '  The  thought  was  that  Gilruth  would  be  the  director.  In 
fact,  in  the  fall  of  1958,  Gilruth  and  Donlan,  figuring  they  were  going  to 
be  the  director  and  deputy  director  of  this  new  Goddard  center,  "went  up 
and  looked  over  the  place  and  what-not."  Donlan  recalls,  "We  spent  some 

73 


Spaceflight  Revolution 

time  thinking  about  how  we  would  organize  it."  In  the  meantime,  however, 
Project  Mercury  was  bubbling  along  at  a  very  fast  rate.  "Silverstein  was 
anxious  to  get  Goddard  moving,  and  he  knew  that  Gilruth  was  going  to  be 
tied  up  with  Mercury,"  Donlan  explains.  So  Silverstein  brought  in  Harry 
Goett,  a  friend  he  used  to  work  with  in  Langley's  Full-Scale  Tunnel  in  the 
old  days  before  Goett  moved  to  the  Ames  laboratory  in  the  early  1940s  and 
before  he,  Silverstein,  moved  to  Lewis.  "This  upset  Gilruth  very  much," 
Donlan  recalls,  "but  nevertheless  he  decided  he'd  rather  work  on  the  manned 
program  than  spend  his  time  organizing  a  new  center."50 

In  truth,  the  STG  always  acted  quite  independently  of  Goddard's  control. 
Harry  Goett,  the  figurehead  director  of  the  STG's  operation,  and  some  of  his 
associates  visited  Langley  almost  weekly  and  were  always  received  "politely 
but  noncommittally."  Goett  told  Donlan  and  others  flatly  that  "he  knew 
what  the  situation  was" :  Goddard's  control  over  the  STG  was  pro  forma  and 
that  most  STG  members,  from  Gilruth  on  down,  felt  some  contempt  for  the 
contrived  relationship.  Fortunately,  the  awkward  "paper"  arrangement  did 
not  last  long  enough  for  hard  feelings  to  develop  on  either  side.51 

By  the  time  of  President  Kennedy's  May  1961  commitment  to  landing 
astronauts  on  the  moon,  everybody  in  NASA  realized  that  the  manned 
space  program  was  never  going  to  be  just  a  division  of  some  other  center. 
Silverstein  and  others  at  NASA  headquarters  finally  decided  to  break  off  the 
STG  as  a  completely  separate  entity,  away  not  only  from  Goddard  but  also 
from  Langley. 

The  fate  of  the  STG,  however,  ultimately  came  to  rest  in  the  hands  of 
powerful  people  beyond  the  control  of  Langley  or  the  STG — or  even  NASA 
headquarters.  Influential  people  representing  vested  political  and  economic 
interests  were  maneuvering  behind  the  scenes  to  build  a  manned  spacecraft 
center  in  Texas.  The  principal  players  behind  the  Texas  plan  were  Vice- 
President  Lyndon  B.  Johnson,  the  nation's  number  two  man  in  the  executive 
office  but  number  one  space  enthusiast;  Representative  Olin  E.  Teague 
of  College  Station,  Texas,  the  third-ranking  member  of  the  House  Space 
Committee;  and  Albert  H.  Thomas,  chairman  of  the  House  Independent 
Offices  Appropriations  Committee,  a  powerful  link  in  the  legislative  chain 
that  reviewed  NASA's  annual  budget  requests.52  In  September  1961,  after 
months  of  unsettling  rumors  (often  denied  by  NASA)  that  the  STG  would 
be  moving  to  a  large  and  expensive  new  facility  in  Texas,  and  despite 
outspoken  criticism  of  the  alleged  backstage  chicanery  expressed  by  the 
outraged  politicos  and  newspapers  serving  the  equally  vested  interests  of 
the  Commonwealth  of  Virginia,  NASA  announced  that  the  STG  would  in 
fact  be  moving  from  Langley  to  a  1620-acre  site  at  Clear  Lake,  some  25 
miles  south  of  Houston,  which  just  so  happened  to  be  in  Albert  Thomas's 
own,  hurricane-torn,  congressional  district.53 

"Now  what's  behind  this  need  for  relocation?"  asked  one  editorial  in  the 
Newport  News  Daily  Press.  And  the  questions  kept  coming: 


74 


Carrying  Out  the  Task 

What  is  needed  that  we  don't  have,  or  can't  get,  right  here  where  the  Space 
Task  Group  was  conceived  and  developed?.  .  .What  is  wrong  with  research  facilities 
presently  located  in  the  Langley  Research  Center  area?  Some  of  this  'back  40'  could 
be  conditioned  for  space  probe  progress  and  closely  related  to  the  existing  complex 
of  laboratories,  facilities,  and  manpower. 

The  simple  one- word  answer,  the  local  media  sourly  reported,  was  "politics." 
This  angered  area  residents.  In  their  minds,  the  activities  of  the  STG  and 
Langley  Research  Center  were  "interwoven."  To  tear  them  apart  was  not 
only  "a  terrible  waste  of  time  and  money"  but  was  also  tantamount  to 
kidnapping  a  brainchild.54 

Many  of  the  STG  members  were  unhappy  as  well.  "I  was  so  upset  about 
going  to  Texas,"  one  STG  engineer  still  remembers  with  indignation,  "I 
wouldn't  even  let  them  send  me  the  free  subscription  to  their  goddanged 
newspaper."  But,  once  the  decision  was  made,  nothing  could  be  done 
about  it  short  of  leaving  the  manned  space  program.  A  native  and  lifelong 
resident  of  Hampton,  Caldwell  Johnson,  who  had  just  built  a  beautiful  new 
waterfront  home,  sums  up  the  predicament:  "I'd  eat  my  heart  out  if  I  stayed 
here  and  let  all  these  other  guys  come  to  Houston  and  do  this.  I  would've 
kicked  myself  fifty  thousand  times."55  In  the  frenetic  period  during  late 
1961  and  early  1962  when  thousands  of  preparations  for  the  first  Mercury 
orbital  flight  still  had  to  be  made,  Caldwell  and  700  other  engineers  and 
their  families  packed  their  belongings  and  drove  the  1000-plus  miles  to  East 
Texas. 

Although  no  one  at  Langley  was  happy  to  see  the  STG  go,  many  sighed 
with  relief  when  the  group  finally  left.  "It  would  have  been  a  great  mistake 
to  have  had  the  STG  stay  at  Langley,"  argues  Charles  Donlan  in  retrospect. 
According  to  Donlan,  who  by  the  time  of  the  move  to  Texas  was  back  at 
Langley  as  Floyd  Thompson's  associate  director,  once  the  decision  was  made 
that  the  STG  would  go  someplace  else,  Thompson  and  everybody  else  felt 
that  "it  was  for  the  best,  because  if  it  had  stayed  it  would  have  overwhelmed 
the  center."56 

The  move  helped  Langley  almost  immediately.  As  a  compensation  for 
the  loss  of  the  STG,  NASA  approved  a  $60-million  expansion  of  Langley 
and  authorized  the  center  to  hire  several  hundred  new  employees  to  replace 
the  departing  STG  members.  Hugh  Dryden,  who  had  been  looking  out  for 
the  interests  of  the  center  at  NASA  headquarters,  was  in  part  responsible 
for  these  boons.  "That  was  the  best  thing  that  could  have  happened,"  says 
Donlan  about  the  authorization  to  hire,  because  one  of  the  most  important 
resources  for  creative  thinking  at  a  research  laboratory  is  a  supply  of  young 
minds.  "We  got  the  cream  of  the  crop  of  many  of  the  best  kids  coming  out 
of  the  universities,"  Donlan  remembers.  Thanks  to  the  STG's  departure, 
Langley  received  a  healthy  infusion  of  the  "fresh  blood"  Thompson  wanted, 
and  instead  of  it  all  flowing  into  space  project  work,  most  of  it  was  channeled 
into  the  general  research  areas.5'  It  was  a  development  that,  on  balance, 

75 


Space/light  Revolution 

pleased  Langley's  senior  management  and  made  them  less  regretful  over  the 
STG's  leaving. 

The  experience  of  having  had  the  STG  at  Langley  also  helped  to  clarify 
management's  thinking  about  the  proper  relationship  between  projects  and 
fundamental  research  and  helped  a  few  to  understand  better  that  all  projects 
eventually  reach  a  dead  end.  Donlan  remembers  the  policy  started  after  the 
STG  moved  away  from  Langley:  "Whenever  a  new  guy  came  in,  we  never 
put  him  in  a  project.  [We  would]  put  him  in  one  of  the  research  divisions 
and  let  him  work  there  for  a  few  years.  If  a  researcher  then  wanted  to  try 
something  else,  fine,  stick  him  in  a  project."58 

A  management  philosophy  that  called  for  a  mix  of  experience  was  healthy 
for  the  overall  NASA  operation,  especially  because  it  enhanced  the  in-house 
capability  of  the  field  centers.  People  assigned  to  projects  did  not  have  to  do 
research  work,  meaning  that  they  could  devote  their  time  to  the  job  at  hand. 
But  the  breadth  and  depth  of  problem-solving  experience  gained  during  the 
required  period  in  major  research  divisions  almost  always  immeasurably 
helped  scientists  and  engineers  if  and  when  they  did  become  involved  in  the 
management  of  a  project. 

Although  the  new  management  philosophy  solved  some  problems,  the 
tension  and  ambivalence  created  by  supporting  development  work  would 
persist  at  Langley  well  beyond  Project  Mercury.  The  same  tension  would  be 
present  through  the  Apollo  program,  the  Viking  project,  the  Space  Shuttle 
program,  the  space  station  program,  and  beyond.  Because  of  Sputnik 
and  the  ensuing  space  race,  development  projects  would  always  be  a  part 
of  Langley,  and  the  conflicting  feelings  surrounding  them  would  never  go 
away.  Buried  deep  inside  those  feelings  was  the  final  and  most  worrisome 
irony  of  all,  which  Hugh  Dryden  tried  to  make  Floyd  Thompson  recognize: 
everything  about  the  space  program  in  the  long  run  could  turn  out  to  be  ad 
hoc  except  research.  No  one  from  the  NACA  except  the  clairvoyant  Hugh 
Dryden  anticipated  this  outcome  of  the  spaceflight  revolution,  and  no  NACA 
veterans  would  be  pleased  by  it. 


The  End  of  the  Glamour  Days 

It  took  about  nine  months,  until  mid-June  1962,  for  the  STG  in  its 
entirety  to  complete  the  move  to  the  new  $60-million  facility  south  of 
Houston.  For  Gilruth  and  associates  this  period  was  busy  and  difficult. 
At  the  same  time  that  they  were  clearing  their  desks  and  packing  their 
files,  families,  and  household  belongings  for  the  western  trek  from  Langley, 
they  were  also  doing  the  thousands  of  things  that  had  to  be  done  to  make 
John  Glenn's  February  1962  Mercury- Atlas  6  flight  (America's  first  manned 
orbital  flight)  and  Scott  Carpenter's  May  1962  Mercury-Atlas  7  flight  the 
great  successes  that  they  turned  out  to  be.59 


76 


Carrying  Out  the  Task 

Thanks  to  President  Kennedy's  May  1961  commitment  to  the  lunar 
landing  program,  the  STG  (renamed  the  Manned  Spacecraft  Center  in 
November  1961)  was  also  gearing  up  to  meet  the  demands  of  what  was 
now  being  called  Project  Apollo.  Although  several  ideas  for  lunar  missions 
had  been  circulating  at  Langley  and  the  other  NASA  centers  for  some  time, 
NASA  did  not  yet  know  how  to  send  an  astronaut  to  the  moon,  how  to  land 
him  on  its  surface  and  return  him  safely,  or  how  to  do  all  three  by  the  end  of 
the  decade  as  President  Kennedy  wanted.  Many  crucial  decisions  had  to  be 
made  quickly  about  the  lunar  mission  mode,  and  the  overworked  manned 
spaceflight  specialists  of  the  STG,  when  they  found  the  time  and  energy, 
were  asked  to  help  make  those  decisions. 

Project  Mercury  came  to  an  end  in  the  early  summer  of  1963,  following 
the  successful  orbital  flights  of  astronauts  Walter  A.  Schirra  (Mercury- 
Atlas  8)  in  October  1962  and  L.  Gordon  Cooper  (Mercury- Atlas  9)  in  May 
1963.  As  the  project  drew  to  a  close,  Bob  Gilruth  wrote  a  letter  to  Floyd 
Thompson,  thanking  his  old  friend  for  all  the  help  that  Langley  had  given 
the  STG  over  the  past  four  years.  "It  is  fitting  that  the  Manned  Spacecraft 
Center  express  its  sincere  appreciation  to  the  Langley  Research  Center 
for  the  invaluable  part  that  the  Center  has  played  in  our  initial  manned 
space  flight  program,"  Gilruth's  letter  stated.  "The  Manned  Spacecraft 
Center  owes  much  to  Langley,  since  . . .  Langley  was  really  its  birthplace." 
Specific  contributions  that  Langley  had  made  to  Project  Mercury  were 
"too  numerous  to  detail  completely"  but  briefly,  they  included  assistance  in 
the  Big  Joe  program;  implementation  of  the  Little  Joe  program;  planning 
and  implementation  of  the  tracking  and  ground  instrumentation  system; 
numerous  aerodynamic,  structural,  materials,  and  component  evaluation 
and  development  tests;  engineering,  shop,  instrumentation,  and  logistic 
support  for  much  of  the  STG  in-house  testing;  and  administrative  support 
and  office  space  during  the  period  from  late  1958  until  mid- 1962  when  the 
STG  completed  its  move  to  Houston.  In  conclusion  Gilruth  wrote,  "As  you 
can  see,  all  elements  of  the  Langley  Center  provided  major  assistance  to 
Project  Mercury,  and  we  are  deeply  grateful  for  this  help."60 

The  local  public  also  wanted  to  express  its  gratitude.  On  Saturday 
morning,  17  March  1962,  more  than  30,000  shouting  and  flag-waving 
residents  of  the  Peninsula  lined  a  25-mile  motorcade  route  through  the  cities 
of  Hampton  and  Newport  News.  The  huge  crowds,  swelling  to  10  and  20 
people  deep  in  some  places,  came  to  salute  the  country's  seven  original 
astronauts,  one  of  whom,  Marine  Lt.  Col.  Glenn,  had  just  made  the  first 
American  orbital  flight  into  space  on  20  February.  Area  residents  wanted 
to  show  the  people  of  NASA  Langley  Research  Center  just  how  much  they 
appreciated  Langley's  effort  to  launch  the  first  Americans  into  space. 

Frequent  cries  of  "Good  work,  John,"  "You're  one  of  us,  Gus,"  and 
similar  encouraging  messages  to  the  seven  smiling  astronauts  followed  the 
impressive  motorcade  throughout  its  meandering  trip  from  Langley  AFB 
to  Darling  Memorial  Stadium  in  downtown  Hampton.  Inside  the  stadium, 

77 


Space/light  Revolution 


In  his  capacity  as  the  astronauts'  public  affairs  officer,  Lt.  Col.  "Shorty"  Powers 
(sitting  in  the  back  seat  of  the  convertible  with  his  wife)  introduced  the  astronauts 
one  by  one  to  the  enthusiastic  crowd.  (Photo  by  Fred  D.  Jones.) 


Astronaut  Walter  M.  Schirra  arrives  at  Darling  Stadium  for  the  rally;  seven  months 
later,  on  3  October  1962,  he  would  become  the  fifth  American  to  be  launched  into 
space.  (Photo  by  Fred  D.  Jones.) 


78 


Carrying  Out  the  Task 


The  mayor  of  Newport  News, 
Va.,  presents  Robert  R.  Gilruth, 
head  of  Project  Mercury,  with 
a  token  of  his  citizens'  esteem 
(left).  Below,  wild  cheers  greet 
astronaut  John  Glenn  and  his 
wife,  Annie.  From  the  out- 
set of  the  manned  space  pro- 
gram, "The  Marine,"  Lt.  Col. 
Glenn,  had  been  the  public 's  fa- 
vorite astronaut.  (Photos  by 
Fred  D.  Jones.) 


79 


Space/light  Revolution 

5000  people  waited  anxiously  in  brisk  50-degree  weather  for  the  arrival  of 
the  parade.  Beneath  the  speakers'  stand  in  the  middle  of  the  football  field, 
where  Manned  Spacecraft  Center  Public  Affairs  Officer  John  A.  "Shorty" 
Powers  would  introduce  the  astronauts  and  Governor  Albertis  S.  Harrison 
would  deliver  the  featured  speech  in  praise  of  them,  the  huddled  spectators 
watched  in  anticipation  as  a  red,  white,  and  blue  banner  fluttered  in  the 
strong  breeze;  the  banner  read:  "HAMPTON,  VA.,  SPACETOWN  U.S.A." 

Behind  the  astronauts  in  the  procession  of  40  open  convertibles  rode 
Gilruth,  Thompson,  and  several  prominent  Langley  researchers  and  senior 
staff  members.  Like  the  astronauts,  the  engineers  smiled  broadly  and  waved 
vigorously  to  the  crowd  while  receiving  lusty  cheers  from  the  throng.61  Help- 
ing to  launch  the  astronauts  into  space  had  altered,  in  a  fundamental  way, 
the  public's  perception  of  who  these  men  were  and  what  they  did.  Instead  of 
NACA  Nuts — those  shadowy  figures  whom  the  public  had  mostly  ignored — 
they  had  become  NASA  Wizards,  the  technological  magicians  who  were 
making  the  incredible  flights  of  mankind  into  space  a  reality. 

Things  had  moved  full  circle.  On  a  previous  Saturday  morning  three 
years  earlier,  NASA  Langley  for  the  first  time  in  its  history  had  opened 
wide  its  gates  and  played  host  to  the  people  of  the  Hampton  area.  Now 
the  people  of  Hampton  were  returning  the  favor.  For  them,  the  glamour  of 
having  the  nation's  first  seven  space  pilots  living  and  working  in  their  midst 
had  been  wonderful.  Losing  them  and  the  rest  of  the  STG  to  Texas  was  a 
bitter  pill  to  swallow.*  Thirty  years  later,  long  after  changing  the  name  of 
busy  Military  Highway  to  "Mercury  Boulevard"  and  dedicating  the  bridges 
of  Hampton  in  honor  of  the  astronauts,  area  residents  still  reminisce  fondly 
about  "the  good  old  days  in  Hampton  and  Newport  News"  when  "those 
brave  astronauts"  lived  in  their  neighborhoods,  ate  in  their  restaurants,  and 
drove  down  their  streets  and  across  their  bridges.62 


jfc 

A  headline  of  the  Newport  News  Daily  Press,  24  Sept.  1961,  read,  "See  Here,  Suh!    What  Does 
Texas  Have  That  Hampton  Doesn't?" 

80 


Change  and  Continuity 


What  we  should  do  is  retain  our  competence  and 
contract  out  our  capacity. 

—Floyd  L.  Thompson,  director  of 
Langley  Research  Center 

In  the  working  partnership  between  universities,  in- 
dustry, and  government  . . .  each  of  the  three  has  re- 
tained its  traditional  values.  .  .  .  I  believe  that  each 
has  become  stronger  because  of  the  partnership. 

— James  E.  Webb,  NASA  administrator 


Born  in  1898,  the  offspring  of  another  century,  Floyd  Thompson  was  62 
years  old  when  he  took  over  officially  from  Henry  J.  E.  Reid  as  the  Langley 
director  in  May  1960.  When  defending  the  interests  of  his  beloved  Langley, 
Thompson  could  definitely  play  the  part  of  a  stubborn  old  curmudgeon. 

He  played  it  particularly  well  in  July  1963  at  a  press  conference  called 
by  NASA  Administrator  James  E.  Webb,  President  Kennedy's  appointed 
successor  to  Keith  Glennan.  In  his  office  at  NASA  headquarters,  Webb 
announced  the  appointment  of  Earl  D.  Hilburn,  a  former  vice-president 
and  general  manager  of  the  Electronics  Division  of  the  Curtiss- Wright 
Corporation,  as  a  new  deputy  associate  administrator.  Webb  reported 
to  the  assemblage  that  Hilburn  would  now  be  responsible  for  the  general 
management  of  all  the  NASA  facilities  that  were  not  manned  spaceflight 
centers — in  other  words,  Hilburn  would  manage  Ames,  Lewis,  Langley, 
Goddard,  Wallops,  JPL  in  Pasadena,  and  the  Flight  Research  Center  at 
Edwards  AFB.  After  introducing  Hilburn,  Webb  turned  to  Thompson,  who 
had  been  invited  to  Washington  just  for  the  occasion,  stuck  out  his  chin,  and 
asked  Langley's  director  what  he  thought  of  the  news.  Thompson  answered 
loud  enough  for  everyone  to  hear,  "Well,  Langley  has  been  around  a  long 

81 


Space/light  Revolution 

time,  and  I  suspect  it  will  be  around  a  lot  longer  no  matter  what  you  people 
up  here  do."1 

This  arrogant  reply  stemmed  from  Thompson's  devotion  to  Langley's 
long  tradition  of  independence  and  freedom  from  bureaucratic  headaches 
and  political  machinations  in  Washington.  Why  should  the  director  of  a 
research  center  be  overly  impressed  by  the  news  that  another  bureaucrat  was 
joining  the  organization  in  Washington?  Although  most  NASA  personnel 
in  the  audience  knew  Thompson  well  enough  not  to  be  stunned  by  his 
comment,  they  were  still  surprised  that  he  would  make  it  in  Webb's  own 
office  and  with  reporters  present.  Neither  Webb  nor  any  other  NASA  officials 
present  would  ever  forget  the  incident.  For  many  of  them,  it  was  just  another 
instance  of  a  prideful  Langley  trying  to  go  its  own  way. 

But  Thompson's  answer  revealed  more  than  just  pride;  it  demonstrated 
his  conviction  that  some  essential  continuity  at  NASA  must  be  sustained 
amid  the  rapid  changes  taking  place  for  the  space  race.  Whatever  man- 
agement changes  or  reforms  NASA  headquarters  made  in  the  affairs  of  the 
research  laboratories,  Langley  would  continue  to  do  its  job. 

The  Langley  center  director  was  no  thoughtless  institutional  reactionary. 
Thompson  had  shown  by  his  nurturing  of  the  STG,  if  not  by  his  comments 
to  headquarters  officials,  that  he  was  no  foot-dragger  when  it  came  to 
supporting  and  promoting  the  space  program.  Not  for  a  moment  would 
he  try  to  stop  the  spaceflight  revolution  from  happening  at  Langley;  rather, 
in  his  own  cautious  and  pragmatic  ways  he  would  advocate,  encourage,  and 
even  delight  in  NASA's  ambitious  objectives.  In  the  changes  in  the  modus 
vivendi  of  the  NASA  laboratory  that  were  taking  place  in  the  early  1960s, 
Thompson  recognized  an  elevated  level  of  excitement  and  commitment, 
a  new  degree  of  freedom,  and  an  unprecedented  opportunity  for  building 
unique  capabilities  that  went  beyond  the  constraints  of  traditional  NACA- 
style,  in-house  research  activities.  After  the  fever  of  supporting  the  space 
race  had  passed,  Langley,  Thompson  believed,  would  emerge  not  weakened, 
but  strengthened. 

Sometime  in  the  early  1960s,  Thompson  invented  a  motto  to  capture 
what  he  wanted  Langley's  new  operational  philosophy  to  be:  "We  should 
retain  our  competence  and  contract  out  our  capacity."2  By  that  Thompson 
meant  that  a  NASA  center  forfeited  none  of  its  own  capabilities  by  sharing 
some  of  its  work  with  outsiders.  Langley  should  nurture  the  industrialization 
of  research  and  development  (R&D),  which  had  been  taking  place  at  an 
increasing  rate  in  America  since  the  end  of  World  War  II.  Langley  would 
not  be  losing  control  over  its  own  destiny  by  farming  out  some  of  its 
responsibilities  to  American  business  and  industry  while  taking  on  certain 
duties  that  went  beyond  the  traditional  in-house  research  function.  Instead, 
Langley  would  benefit  because  it  would  now  be  able  to  focus  on  the  genesis  of 
valuable  scientific  and  technological  ideas,  take  its  own  potential  to  the  limit, 
and  accomplish  important  tasks  that  could  not  be  done,  as  well  any  other 
way.  Moreover,  in  spreading  its  wealth  to  contractors,  NASA  would  not  just 


Change  and  Continuity 


Langley  Director  Floyd  Thompson 
might  have  done  more  to  obstruct 
the  spaceflight  revolution  at  Lang- 
ley  if  he  had  known  how  much  the 
essential  character  of  his  research 
center  was  going  to  be  changed  by 
it. 


L-63-3879 


be  putting  together  a  national  team  to  beat  the  Soviets  in  the  space  race 
but  would  also  be  invigorating  the  aerospace  industry  and  strengthening 
the  country's  economy.  NASA's  new  style  of  managing  large  endeavors 
might  even  demonstrate  a  cooperative  means  by  which  other  national  and 
international  needs,  such  as  the  alleviation  of  poverty,  could  be  met. 

Thompson's  slogan  was  his  own,  but  the  broader  message  belonged 
to  bigger  thinkers.  Whether  Thompson  was  conscious  of  it  or  not,  the 
phrase  "retain  our  competence  and  contract  out  our  capacity"  echoed  a 
more  sweeping  vision  of  his  time,  one  that  was  an  essential  part  of  John 
F.  Kennedy's  "New  Frontier."  An  ambitious  program  of  space  exploration 
would  make  the  United  States  an  overall  healthier  society,  Kennedy  declared 
during  the  1960  presidential  campaign.  "We  cannot  run  second  in  this  vital 
race.  To  insure  peace  and  freedom,  we  must  be  first.  .  .  .  This  is  the  new 
age  of  exploration;  space  is  our  great  New  Frontier."  A  successful  space 
program  would  help  in  the  ultimate  defeat  of  communism  by  showing  to  all 
peoples  of  the  world  the  great  things  a  western  democratic  and  capitalistic 
society  can  do  when  its  resources  are  effectively  mobilized.  NASA  would 
make  manifest  the  essential  superiority  of  the  American  way  of  life.3 


83 


Space/light  Revolution 


L-64-6455 

President  Kennedy's  (actually  Vice- President  Lyndon  B.  Johnson's)  choice  for 
NASA  administrator,  James  E.  Webb,  sits  between  Langley  Director  Floyd 
Thompson  (left)  and  the  NACA's  retired  executive  secretary,  John  F.  Victory 
(right),  at  the  1964  NASA  Inspection  held  at  Langley.  Although  Thompson  had 
some  memorable  run-ins  with  Webb,  the  two  were  much  alike.  Both  were  country 
boys  (Thompson  from  Michigan,  Webb  from  North  Carolina)  with  rumpled  collars, 
corn-pone  accents,  and  down-home  homilies.  They  were  also  highly  intelligent,  com- 
plex, and  cunning. 


A  dynamic  new  union  of  science,  technology,  and  government — or  what 
NASA  Administrator  Webb  called  the  "university-industry-government 
complex" — would  lead  the  charge  in  this  campaign  for  a  better  world. 
NASA,  above  all  other  national  institutions,  would  help  to  forge  this  re- 
lationship, which  would  serve  as  the  means  for  winning  the  contest  with  the 
Soviet  Union,  for  solving  pressing  social  and  economic  problems  at  home 
and  abroad,  and  for  accelerating  the  pace  of  progress  in  the  human  com- 
munity. The  elaborate  teamwork  necessary  for  spaceflight  programs  would 
force  some  major  changes  and  adjustments  in  the  workings  of  existing  insti- 
tutions, but  the  end  result  would  be  for  the  best.  Traditional  values,  those 
worth  keeping,  would  be  retained.  But  through  the  new  partnership,  each 
of  the  team  members,  including  NASA,  would  become  stronger.4 

Given  his  democratic  leanings  and  his  position  of  responsibility  within 
the  space  program,  not  even  a  stubborn  old  curmudgeon  like  Thompson 

84 


Change  and  Continuity 

was  outside  the  rising  tide  of  thinking  about  how  the  world  was  changing 
and  how  even  successful  places  like  NASA  Langley  would  also  have  to 
change  if  they  were  to  contribute  to  and  be  a  vital  part  of  the  new  order. 
Thompson  could  be  brusque  with  Webb,  as  at  the  Hilburn  press  conference, 
but  in  technological  spirit  he  and  the  NASA  administrator  stood  on  common 
ground.  "Every  thread  in  the  fabric  of  our  economic,  social,  and  political 
institutions  is  being  tested  as  we  move  into  space,"  Webb  stated  in  a  1963 
speech  on  the  meaning  of  the  space  program. 

Our  economic  and  political  relations  with  other  nations  are  being  reevaluated.  Old 
concepts  of  defense  and  military  tactics  are  being  challenged  and  revised.  Jealously 
guarded  traditions  in  our  educational  institutions  are  being  tested,  altered,  or  even 
discarded.  Our  economic  institutions — the  corporate  structure  itself — are  undergoing 
reexamination  as  society  seeks  to  adjust  itself  to  the  inevitability  of  change. 

Thompson,  in  his  much  less  publicized  talks  around  Langley,  often  echoed 
the  same  sentiments.  Not  even  the  oldest  and  best  American  institutions 
could  go  on  as  before,  unaffected,  in  light  of  the  technological  revolution  that 
was  taking  place  as  humankind  moved  into  space.  Even  a  place  like  NASA 
Langley  would  have  to  make  some  major  adjustments,  and  Thompson  knew 
it — no  matter  what  curt  remark  he  might  make  to  the  contrary. 


The  Organization 

Apart  from  meeting  the  sizable  personnel  requirements  of  the  STG, 
Langley  laboratory  initially  did  not  change  much  to  meet  the  growing 
demands  of  the  nascent  space  age.  Some  new  boxes  were  drawn  on  the 
organization  charts,  and  a  few  old  ones  eliminated.  Some  existing  divisions 
and  branches  received  new  names  and  experienced  reorganizations,  and 
a  few  significant  new  research  sections  and  branches  were  built  around 
emerging  space  disciplines  (for  example,  the  Space  Applications  Branch 
of  the  Full- Scale  Research  Division  created  in  December  1959  and  the 
Magnetoplasmadynamics  Branch  of  the  Aero-Physics  Division  created  in 
May  1960).  Several  major  project  offices  also  came  to  life  at  the  laboratory 
in  the  early  1960s,  but,  for  the  most  part,  everything  about  the  formal 
structure  of  the  laboratory  remained  the  same  as  before.  Thompson  and  his 
senior  staff  believed  that  the  organization  of  the  laboratory  for  its  general 
applied  aerodynamic  research  under  the  NACA  in  the  late  1950s  would 
serve  the  new  combination  of  aeronautics  and  space  equally  well.  When 
Langley's  diversified  capabilities  needed  to  be  focused  on  mission  plans 
or  specific  program  goals,  ad  hoc  task  forces,  steering  committees,  study 
groups,  and  other  "shadow  organizations"  that  usually  did  not  appear  on 
the  organization  charts  were  created. 

The  organization  chart  of  1962  shows  the  continuity  in  Langley's  struc- 
ture from  the  1950s  into  the  1960s  even  though  four  years  had  passed  since 

85 


Space/light  Revolution 

the  changeover  of  the  NACA  to  NASA  and  one  year  had  passed  since 
President  Kennedy  had  committed  the  country  to  a  manned  landing  mission 
to  the  moon  by  the  end  of  the  decade.  In  the  summer  of  1962,  Langley 
Research  Center  consisted,  as  it  had  since  the  mid-1950s,  of  three  major  re- 
search directorates.  Heading  each  directorate  was  an  assistant  director.  This 
person  was  responsible  for  overseeing  the  work  being  done  in  the  subsidiary 
research  divisions.  In  1962  each  research  directorate  had  three  research  di- 
visions, for  a  total  of  nine  at  the  center.  Within  the  nine  research  divisions 
were  some  50  branches,  plus  a  number  of  sections,  offices,  facilities,  shops, 
and  testing  units.  Typically,  a  division  numbered  between  100  and  150  full- 
time  research  professionals.  In  the  management  formula,  3  nonprofessionals, 
that  is,  secretaries,  mechanics,  data  processors  and  the  like,  were  needed  to 
support  one  researcher.  That  did  not  mean  that  every  division  employed 
300  to  450  support  people;  none  in  fact  did.  The  research  divisions  instead 
received  much  of  their  nonprofessional  assistance  from  two  supporting  di- 
rectorates. One  of  these  directorates,  "technical  services,"  employed  the 
mechanics,  modelmakers,  electricians,  and  other  technicians  necessary  for 
keeping  the  shops,  testing  facilities,  and  the  rest  of  the  infrastructure  of  the 
research  operation  alive.  The  other  supporting  directorate,  "administrative 
services,"  handled  fiscal  matters,  personnel  affairs,  the  photo  lab,  the  library, 
and  the  publications  office  as  well  as  the  rapidly  increasing  requirements  for 
procurement.6 

Until  early  in  1962,  the  research  directorates  did  not  have  names  or 
any  official  designation;  on  the  organization  charts  were  three  boxes  simply 
labeled  "Office  of  Assistant  Director"  with  no  way  to  distinguish  them, 
apart  from  knowing  who  the  particular  assistant  director  was  and  what 
divisions  he  directed.  In  February  1962,  Director  Thompson  and  Associate 
Director  Charles  J.  Donlan  decided  to  remedy  this  situation.  There  were 
three  directorates,  they  thought,  so  why  not  call  them  "Group  1,"  "Group 
2,"  and  "Group  3."7 

Named  to  head  Group  1  at  the  time  of  this  nominal  reorganization  was 
Clinton  E.  Brown,  formerly  the  chief  of  the  Theoretical  Mechanics  Division. 
This  division  was  one  of  several  smaller  Langley  divisions  that  in  the  early 
1960s  were  focusing  on  the  study  of  lunar  missions.  Brown  replaced  Hartley 
A.  Soule,  who  retired.  The  new  Analysis  and  Computation  Division  (ACD), 
whose  chief  was  Paul  F.  Fuhrmeister,  was  part  of  Group  1.  This  division  was 
established  in  January  1961  by  combining  the  Analytical  and  Computation 
Branch  of  the  Theoretical  Mechanics  Division  with  the  Data  Systems  Branch 
of  IRD.  The  goal  of  ACD  was  "to  allow  more  effective  management  at  the 
Center  in  the  development  and  utilization  of  data  systems  for  data  reduction 
services  and  for  theoretical  analysis  requirements."8  Also  within  Brown's 
group  were  IRD,  headed  by  electrical  engineer  and  future  assistant  director 
Francis  B.  Smith,  and  the  Theoretical  Mechanics  Division  (in  June  1963, 
renamed  the  Space  Mechanics  Division),  led  by  Dr.  John  C.  Houbolt,  the 
champion  of  the  lunar-orbit  rendezvous  concept  for  Project  Apollo. 

86 


Change  and  Continuity 


II  h 
II  ii 


ill! 


is  !§!§ 

Og       <ir2^ 

II  full 


i   -e^ 


87 


Space/light  Revolution 


L-60-5232 


L-64-9585 


L- 77-3336 

The  heads  of  Groups  1,  2,  and  3:  Clinton  E.  Brown  (top  left),  Eugene  C.  Draley 
(top  right),  and  Laurence  K.  Loftin,  Jr.  (bottom).  When  these  groups  were  baptized 
in  February  1962,  the  three  men  had  been  working  at  Langley  for  a  combined  62 
years. 


Change  and  Continuity 

Heading  Group  2  was  Eugene  C.  Draley,  who  had  been  serving  as  an 
assistant  director  since  November  1958.  Within  this  directorate  was  Joseph 
A.  Shortal's  (Class  of  1929,  Texas  A&M)  Applied  Materials  and  Physics 
Division,  the  reincarnation  of  PARD,  which  had  been  dissolved  in  December 
1959.  PARD,  created  near  the  end  of  World  War  II,  had  developed  the 
methods  of  rocket-model  testing  at  Wallops  and  had  provided  instrumented 
flight  data  at  transonic  and  supersonic  speeds  important  for  the  design 
of  the  country's  postwar  high-speed  jets  and  ballistic  missiles.  Led  in  its 
early  years  by  Bob  Gilruth,  the  old  PARD  had  served  almost  unwittingly 
as  Langley's  training  ground  for  the  space  age.  One  year  after  the  birth 
of  NASA,  and  in  view  of  the  changed  programs  and  responsibilities  of 
PARD,  Langley  had  changed  its  name  to  the  Applied  Materials  and  Physics 
Division.9  The  Dynamic  Loads  Division,  headed  by  I.  Edward  Garrick,  an 
applied  mathematician  who  had  graduated  from  the  University  of  Chicago 
in  1930,  and  the  Structures  Research  Division,  headed  by  MIT  aeronautical 
engineer  (Class  of  1942)  Richard  R.  Heldenfels,  were  also  in  Group  2. 

Laurence  K.  Loftin,  Jr.,  a  mechanical  engineer  who  came  to  work  at 
Langley  in  1944  after  graduating  from  the  University  of  Virginia,  had  served 
as  the  technical  assistant  to  Floyd  Thompson  since  December  1958.  When 
Henry  Reid  relinquished  his  duties  on  20  May  1960,  Loftin  began  working 
for  the  laboratory's  director.  On  24  November  1961  Loftin  replaced  John 
Stack  as  Langley's  third  assistant  director  when  Stack  moved  up  to  take 
charge  of  the  agency's  aeronautical  programs  at  NASA  headquarters.  In 
practice,  Loftin  served  also  as  Langley's  director  for  aeronautics.  When, 
four  months  later,  Thompson  assigned  group  numbers  to  the  directorates, 
Loftin  remained  in  charge  of  what  was  called  from  then  on  Group  3. 

Group  3  was  home  to  the  Aero-Physics  Division,  headed  by  hypersonics 
expert  John  V.  Becker  (M.S.  in  aeronautical  engineering,  New  York  Univer- 
sity, 1935).  The  roots  of  this  division  went  back  to  the  old  Compressibility 
Research  Division  of  the  late  1940s  and  1950s,  in  which  NACA  researchers 
had  studied  the  vexing  problems  of  high-speed  flight  in  new  wind  tunnels 
and  other  unique  test  facilities.  In  December  1958,  Langley  had  redesignated 
this  division  the  Supersonic  Aerodynamics  Division.  But  this  name,  which 
Becker  and  others  did  not  like  because  it  did  not  capture  the  range  of  re- 
search areas  covered  by  the  division's  work,  did  not  last  long.  Seven  months 
later,  after  another  reorganization,  it  was  rebaptized  the  Aero-Physics  Divi- 
sion, a  title  that  then  lasted  until  the  major  organizational  shake-up  brought 
on  in  1969  and  1970  by  Thompson's  successor  as  center  director,  Edgar  M. 
Cortright. 

The  second  division  in  Group  3  was  the  Aero-Space  Mechanics  Division, 
led  by  Philip  Donely,  an  aeronautical  engineer  who  had  graduated  from 
MIT  in  1931.  Like  a  few  other  parts  of  Langley,  this  division  was  cre- 
ated not  long  after  the  establishment  of  NASA,  during  the  reorganization  of 
September  1959.  Essentially,  the  Aero-Space  Mechanics  Division  combined 
two  older  aeronautical  research  groups:  the  Flight  Research  Division  and 

89 


Spaceflight  Revolution 


L-64-6446 

On  the  Langley  tarmac  in  May  1964  to  welcome  Raymond  Bisplinghoff,  director  of 
the  Office  of  Advanced  Research  and  Technology  (OART)  at  NASA  headquarters, 
are,  left  to  right,  Floyd  Thompson,  Langley  director;  Raymond  Bisplinghoff;  T. 
Melvin  Butler,  chief  administrative  officer;  Eugene  C.  Draley,  head  of  Group  2;  and 
Laurence  K.  Loftin,  head  of  Group  3. 

the  Stability  Research  Division,  both  of  which  dated  to  the  late  1930s.  As 
with  many  other  changes  at  the  time,  the  establishment  of  the  Aero-Space 
Mechanics  Division  reflected  the  snowballing  of  space-related  research  activ- 
ities at  Langley  and  the  de-emphasis  on  aeronautics.  In  a  directorate  such 
as  this,  where  aeronautics  always  had  been  the  byword,  center  management 
was  reclassifying  activities  to  show  how  even  the  airplane  flight  research 
groups  were  tackling  critical  problems  in  the  new  regime  of  space.  Nobody 
in  Donely's  division  much  liked  the  new  name,  because  it  eliminated  the 
word  "flight"  to  add  the  word  "space."  Effective  30  June  1963,  after  a  reor- 
ganization, Donely's  division  became  the  Flight  Mechanics  and  Technology 
Division,  a  redesignation  that  stuck  until  the  Cortright  reorganization,  when 
the  political  advantages  of  calling  everything  "space" -this  or  "space" -that 
had  mostly  passed. 

The  third  division  in  Loftin's  group  was  the  Full-Scale  Research  Divi- 
sion, which  comprised  several  large  aeronautics  groups  clustered  around  the 
laboratory's  larger  wind  tunnels.  This  division  began  in  the  early  1940s 
but  had  recently  expanded  in  May  1961  with  the  addition  of  the  former 
Unitary  Plan  Wind  Tunnel  Division  as  one  of  its  major  research  branches. 
Aeronautical  engineer  Mark  R.  Nichols,  a  1938  graduate  of  the  Alabama 


90 


Change  and  Continuity 

Polytechnical  Institute  (later  Auburn  University),  led  this  division  through 
the  1960s. 

Not  surprisingly,  all  three  assistant  directors  and  all  nine  division  chiefs, 
as  well  as  the  director  and  associate  director,  were  former  employees  of 
the  NACA.  The  average  age  of  these  14  men  in  1962  was  just  over  44. 
When  Lindbergh  made  his  famous  transatlantic  crossing  in  1927,  they  were 
young  boys.  Many  of  them  remembered  the  flight  of  "Lucky  Lindy"  as  a 
seminal  event  in  their  lives,  launching  them  toward  professional  careers  in 
aeronautics.  Only  two  of  them,  ACD's  Fuhrmeister  and  IRD's  Smith,  had 
not  worked  at  NACA  Langley  during  World  War  II,  but  they  arrived  only 
a  few  years  later. 

A  few  important  changes  in  the  structure  of  the  organization  occurred 
after  1962.  A  handful  of  new  assistant  directors  would  be  assigned.  In 
October  1965,  IRD  would  be  split  into  two  divisions:  a  new  IRD  and  a 
brand  new  Flight  Instrumentation  Division.  Both  divisions  would  belong 
to  Group  1.  In  the  spring  of  1964,  a  fourth  major  research  directorate, 
the  Office  for  Flight  Projects,  was  formed  to  accommodate  the  growing 
number  of  special  projects  at  the  laboratory.  Under  this  office  was  placed 
the  Flight  Reentry  Programs  Office,  which  handled  Project  Fire,  the  Lunar 
Orbiter  Project  Office,  the  Manned  Orbiting  Research  Laboratory  (MORL) 
Studies  Office,  the  Scout  Project  Office,  and  the  Applied  Materials  and 
Physics  Division  (the  old  PARD).  The  first  assistant  director  of  this  new 
Office  for  Flight  Projects  was  Gene  Draley,  who  moved  over  from  Group  2. 
Replacing  Draley  as  head  of  Group  2  was  Dr.  John  E.  Duberg  (Ph.D.  from 
the  University  of  Illinois,  Class  of  1948).  Duberg  was  responsible  for  a 
directorate  comprising  only  the  Dynamic  Loads  Division  and  the  Structures 
Research  Division.  The  Applied  Materials  and  Physics  Division,  which  for 
its  entire  history  had  been  the  maverick  in  Langley's  overall  organization, 
moved  over  with  Draley  to  Flight  Projects.10  Curiously,  this  fourth  research 
directorate  was  not  called  "Group  4." 


Thompson's  Obscurantism 

Langley's  organization  charts  did  not  reveal  the  substance  of  the  labora- 
tory operation.  In  keeping  with  a  long-standing  tradition  of  obscurantism 
fathered  by  George  W.  Lewis,  the  NACA's  politically  shrewd  director  of  re- 
search in  Washington  from  1919  to  1947,  Langley  Directors  Henry  Reid  and 
Floyd  Thompson  never  made  the  structure  of  the  laboratory  too  apparent. 
If  they  had,  they  thought,  then  outsiders — and  that  category  of  suspicious 
people  included  Langley's  own  superiors  at  NACA/NASA  headquarters  in 
Washington — would  be  able  to  interfere  with  what  was  going  on  inside  the 
laboratory.  Micromanagement  was  something  that  the  directors  of  the  field 
centers  and  their  research  staffs  definitely  did  not  want.11 


91 


Space/light  Revolution 

"Thompson  was  a  great  one  for  saying  that  you  couldn't  be  too  sensible 
about  this  kind  of  stuff,"  remembers  Larry  Loftin,  assistant  director  for 
Group  3.  He  wanted  to  "keep  things  confused  so  that  the  people  at 
headquarters  wouldn't  really  know  what  was  going  on."  Thus,  Langley's 
formal  organization,  following  the  NACA  way,  was  kept  deliberately  vague. 
Loftin  remembers  one  instance  from  the  early  1960s  when  a  concerned 
Bernard  Maggin  from  the  Office  of  Aeronautical  and  Space  Research  in 
NASA  headquarters  asked  Thompson  outright  how  many  people  were 
working  on  space  projects  under  William  J.  O'Sullivan,  Jr.,  in  Langley's 
Applied  Materials  and  Physics  Division.  Thompson  just  looked  at  Maggin 
grimly  (to  some  colleagues,  the  Langley  center  director  was  known  as  "The 
Grim  Reaper")  and  said,  "I'm  not  going  to  tell  you."12 

And  he  never  did  tell  Maggin.  Thompson  could  get  away  with  veiling 
the  organization  because  of  his  many  years  with  the  NACA,  the  outstanding 
reputation  of  Langley  Research  Center  both  inside  and  outside  the  agency, 
and  the  power  Langley  wielded  early  on  within  NASA.  This  policy  of  ob- 
scurantism, however,  was  not  something  that  headquarters  liked  or  wanted 
continued  much  longer;  it  was  not  to  be  carried  on  by  Thompson's  successor. 
Edgar  M.  Cortright,  the  headquarters  official  named  by  NASA  Administra- 
tor Jim  Webb  in  March  1968  to  replace  Thompson,  believed  that  it  would  be 
to  Langley's  advantage  if  headquarters  had  a  more  detailed  understanding 
of  the  laboratory  operation.  So,  in  1969  and  1970,  when  he  put  Langley 
through  what  was  the  most  sweeping  and  traumatic  reorganization  in  its 
then  more  than  50-year  history,  Cortright  made  certain  that  the  titles  in  all 
the  boxes  on  the  organization  charts  indicated  exactly  what  staff  members 
did.  This  was  just  what  Thompson  had  avoided. 

Another  hallmark  of  Thompson's  management  style  was  generating 
spirited  competition  among  his  research  divisions.  He  did  not  want  any  one 
group  to  have  all  the  research  opportunities  in  a  given  technical  area.  No 
one  group  should  be  doing  all  the  reentry  heating  work,  all  the  space  station 
design,  or  all  the  supersonic  research.  Monopolies  such  as  that,  though  they 
might  seem  to  prevent  duplication  of  efforts,  bred  complacency.  Better  to 
have  several  research  groups  tackling  the  same  set  of  problems  from  different 
angles. 

This  philosophy  of  creative  research  through  friendly  competition  led 
to  the  formation  of  shadow  organizations  and  invisible  lines  of  organiza- 
tional communication  and  responsibility  within  Langley — a  process  that 
would  become  known  to  management  theorists  by  the  1990s  as  "nonlinear" 
thinking.  For  example,  besides  serving  as  assistant  director  for  Group  3, 
Loftin  also  was  responsible  for  all  the  aeronautics  efforts  at  the  laboratory— 
that  included  all  of  the  aeronautical  work  in  the  Structures  Research  Divi- 
sion, which  was  technically  under  the  auspices  of  Gene  Draley's  Group  2. 
As  part  of  his  everyday  duties,  Loftin  had  to  review  and  approve  all  the 
important  paperwork  related  to  the  aeronautical  activity  of  someone  else's 
directorate.13 

92 


Change  and  Continuity 

This  arrangement  did  create  some  tension  but  frequently  resulted  in  a 
positive  outcome.  "There  was  enormous  technical  competition  between  the 
divisions  at  Langley,"  remembers  Israel  Taback,  a  longtime  member  of  IRD 
who  came  to  work  at  the  laboratory  in  the  early  1940s  and  stayed  into  the 
1980s.  "People  would  fight  with  each  other  over  technical  details.  That  was 
all  very  healthy.  The  end  result  was  a  battle  of  ideas.  Ideas  that  had  merit 
tended  to  float  to  the  surface.  The  good  ideas  won." 14 


The  Sinking  of  Hydrodynamics — and  Aeronautics? 

Only  one  major  research  division  completely  disappeared  at  Langley 
during  the  first  years  of  the  spaceflight  revolution:  Hydrodynamics.  This 
division  had  done  pioneering  work  in  the  field  of  waterborne  aircraft  research 
since  1930.  Langley  management  decided  to  dissolve  Hydrodynamics  in 
late  December  1959  and  reassign  its  roughly  four  dozen  personnel  to  other 
divisions.  Many  of  its  staff  members  went  to  Dynamic  Loads,  which  dated 
back  to  the  old  Aircraft  Loads  Division  of  World  War  II  and  had  specialized 
in  the  study  of  such  problems  as  aeroelasticity,  flutter,  buffeting,  ground 
wind  loads,  gust  loads,  and  aircraft  noise.  In  recent  months,  however, 
Dynamic  Loads,  like  most  other  Langley  divisions,  had  been  taking  on  work 
related  to  Project  Mercury  and  the  space  program.15 

With  the  group  that  moved  to  Dynamic  Loads  went  the  continued 
responsibility  for  operating  the  High-Speed  Hydrodynamics  Tank,  a  2177- 
foot-long,  8-foot-wide,  and  5-foot-deep  towing  test  basin.  This  long  concrete 
water  channel  was  located  in  the  far  West  Area  of  Langley  Field  alongside 
the  Dynamic  Loads  Division's  Landing  Loads  Track.*  In  the  High-Speed 
Hydrodynamics  Tank,  NAG  A  researchers  in  the  mid-1950s  had  evaluated 
the  performance  of  floats  for  the  navy's  Martin  YP6M-1  Seamaster  jet- 
propelled  flying  boat.  They  had  worked  to  develop  retractable  "hydro- 
skis"  for  the  navy's  experimental  little  XF2Y-1  Sea  Dart  jet  fighter  built 
by  Convair  (still  to  this  day  the  only  supersonic  seaplane  ever  to  fly). 
In  addition,  they  had  searched  for  a  way  to  provide  water-based  aircraft 
with  the  combat  air  performance  of  comparable  land-based  planes.  These 
investigations  contributed  information  essential  to  the  design  of  several 
experimental  military  vehicles  including  a  "panto-base"  airplane,  a  proposed 
amphibious  type  that  could  operate  from  concrete  runways,  grass,  mud, 
snow,  sandy  beaches,  or  even  from  seaplane  ramps  and  floating  rafts.16 

Those  members  of  the  Hydrodynamics  Division  who  did  not  move  to  Dy- 
namic Loads  became  members  of  the  Full-Scale  Research  Division.  This  was 
the  largest  single  division  at  the  laboratory,  and  it  was  essentially  composed 
of  aeronautical  researchers  who  staffed  the  larger  wind  tunnels.  John  B. 


The  Landing  Loads  Track  was  an  outdoor  facility  that  simulated  aircraft  landing  loads  and  motions 
through  the  braking  and  impact  of  a  catapult-launched  test  carriage  onto  a  hard  runway-like  surface. 

93 


Space/light  Revolution 

Parkinson,  Hydrodynamic's  ever-faithful  chief  and  the  1957  winner  of  the 
first  Water-Based  Aviation  Award  given  by  the  Institute  of  Aeronautical 
Science,  was  reassigned  to  this  division.  Parkinson  had  worked  in  Hydro- 
dynamics since  coming  to  Langley  in  1931.  He  accepted  with  reluctance 
his  new  assignment  as  "Aeronautical  Research  Scientist,  Aerodynamics," 
which  then  Associate  Director  Floyd  Thompson  invented  for  him.  In  that 
position,  Parkinson  was  to  help  in  program  planning  and  serve  as  "the  Cen- 
ter's consultant  for  the  consideration  of  future  vehicles  that  operate  on  or 
in  the  water  as  part  of  their  mission  and  other  future  vehicles  for  which  wa- 
ter landing  or  other  hydrodynamic  requirements  affect  and  modify  design 
requirements."  As  Parkinson  would  no  longer  be  a  division  chief,  Langley 
had  to  request  an  "excepted  position"  for  him  from  the  civil  service  that 
would  allow  him  to  retain  his  present  salary  of  $15,500.  Within  two  years 
of  the  dissolution  of  Hydrodynamics,  Parkinson  left  Langley  for  a  job  over- 
seeing the  management  of  aerodynamics  research  in  the  Office  of  Advanced 
Research  and  Technology  (OART)  at  NASA  headquarters.17 

Parkinson  and  his  colleagues  took  with  them  to  the  Full- Scale  Research 
Division  the  responsibility  for  maintaining  what  had  always  been  Hydrody- 
namics premier  facility,  "Tank  No.  1,"  a  unique  2900- foot  indoor  seaplane 
towing  basin  on  the  shore  of  the  Back  River  in  the  East  Area.  This  tank 
was  designed  in  1930  by  NAG  A  civil  engineer  Starr  Truscott,  who  according 
to  Langley  lore  was  a  descendant  of  the  Wild  West  outlaw  Belle  Starr  and 
a  veteran  of  the  construction  of  the  Panama  Canal.  The  NACA's  original 
hydrodynamics  research  program  had  begun  in  Tank  No.  1  when,  29  years 
before,  Truscott,  Parkinson,  and  fellow  engineers  had  employed  it  to  test 
floats  that  were  eventually  used  on  several  American  seaplanes,  including 
the  Sikorsky  twin-float  "Amphibian,"  which  set  speed  records  in  the  1930s. 
Data  gained  from  work  in  this  facility  also  contributed  to  the  development 
of  the  famous  Clipper  flying  boats,  the  romantic  ocean-hoppers  that  before 
World  War  II  had  trailblazed  air  routes  and  carried  hundreds  of  paying  pas- 
sengers over  all  the  oceans  of  the  world.  In  the  big  water  tank,  the  NACA 
had  studied  the  design  characteristics  of  most  American  floatplanes  and  the 
performance  of  nearly  all  the  early  U.S.  Navy  flying  boats  that  would  be 
used  for  air-sea  rescue,  antisubmarine  patrol,  and  troop  transport  in  World 
War  II.  In  the  enlarged  version  of  the  tank  (it  was  lengthened  to  its  full 
2900  feet  from  an  original  2000  feet  in  1937)  and  in  its  1800-foot-long  lit- 
tle brother,  Tank  No.  2  (built  adjacent  to  it  in  1942),  Langley  engineers 
discovered  ways  to  ease  the  shock  on  a  landplane  when  crash-landing  or 
ditching  in  the  water.  Both  tanks  were  equipped  with  an  overhead  elec- 
tric carriage  from  which  a  dynamic  model  could  be  suspended  and  towed 
at  up  to  80  miles  per  hour,  which  was  sufficient  to  make  a  model  take  off 
from  the  water  and  fly  at  scale  speed.  As  the  model  was  moving  along 
the  surface,  researchers  took  motion  pictures  and  recorded  measurements 
demonstrating  the  aircraft's  stability,  controllability,  water  resistance,  drag, 
and  spray  characteristics.  The  tanks  were  equipped  with  catapult  devices, 

94 


Change  and  Continuity 


L-59-8685 

Aerial  view  of  Langley's  East  Area.  The  largest  building  on  the  shore  of  the  Back 
River  is  the  Full-Scale  Tunnel;  the  long  building  seeming  to  run  from  the  top  of  the 
tunnel  is  Tank  No.  1. 

for  the  study  of  the  free-launched  landing  characteristics  of  airplanes  and 
with  mechanical  wave-makers,  for  the  simulation  of  takeoff  and  landing  in 
rough  water.18 

On  the  eve  of  the  dissolution  of  the  Hydrodynamics  Division,  researchers 
in  Tank  No.  1  were  studying  the  characteristics  of  revolutionary  VTOL 
machines  over  water.  They  were  even  investigating  the  requirements  of  a 
supersonic  seaplane  and  a  prototype  "ground-effect"  machine,  a  platform- 
like  vehicle  that  could  hover  and  move  just  above  the  ground  by  creating 
a  cushion  of  supporting  air  between  it  and  the  ground  surface.  Nobody, 
not  even  the  U.S.  Navy,  was  interested  enough  in  the  research  going  on 
in  Langley's  Hydrodynamics  Division  to  ask  NASA  to  keep  it  alive.  Two 
ambitiously  experimental  Martin  YP6M-1  Seamaster  jet  seaplanes  had 
recently  been  lost  due  to  design  failures;  the  navy  was  about  to  terminate 
its  entire  flying-boat  program;  and  Martin,  one  of  the  most  dedicated 
builders  of  flying  boats,  was  on  the  verge  of  moving  into  the  guided  missile 
business.19  Langley's  Hydrodynamics  Division,  historic  as  it  was,  had 
apparently  outlived  its  usefulness.  Tank  No.  2  had  already  been  deactivated 
in  April  1958  after  16  years  of  continuous  use.  Beginning  on  the  first  working 
day  of  1960,  historic  Tank  No.  1  would  be  placed  on  standby  status,  with 
no  operating  personnel  regularly  assigned  to  it.  It  became  an  abandoned 
facility  that  was  to  be  used  "only  to  meet  the  requirements  of  such  special 
needs  as  they  might  arise."20  Shortly  thereafter,  NASA  would  give  complete 
control  over  the  tank  to  the  navy. 


95 


Spaceflight  Revolution 


L-61-6509 

A  Langley  engineer  prepares  a  model  of  the  proposed  air  force  X-20  Dyna-Soar 
aerospace  plane  for  testing  in  Tank  No.  2  in  1961. 


Langley  management  explained  its  decision  to  abolish  the  Hydrodynam- 
ics Division  by  pointing  to  "the  declining  need  for  hydrodynamics  research 
as  it  applies  to  seaplanes  and  other  water-borne  aircraft."  Although  that 
justification  was  apparently  legitimate,  it  was  only  half  the  story.  The  other 
half  was  that  the  exigencies  of  NASA's  space  program  were  sweeping  over 
Langley  like  a  tidal  wave  and,  in  this  case,  engulfing  an  entire  aeronautics- 
oriented  division  whose  activities,  facilities,  and  reason  for  being  suddenly 
seemed  antiquated.  It  did  not  matter  that  the  division  had  been  contributing 
to  Project  Mercury  by  making  studies  of  the  water  landing  characteristics  of 
the  capsule;  it  was  better  to  get  rid  of  the  division  and  make  its  staff  more 
clearly  a  part  of  the  new  regime  of  space.  "In  view  of  the  changing  nature  of 
the  nation's  research  programs,"  conceded  Langley  Director  Henry  Reid,  "it 
is  felt  that  the  experienced  personnel  of  the  Hydrodynamics  Division  could 
best  be  utilized  by  transferring  them  to  the  staffs  of  divisions  which  have 
assumed  increased  space  research  responsibilities  in  recent  months." 

As  indicated  by  the  elimination  of  the  Hydrodynamics  Division  in  1959, 
Langley  management  was  doing  everything  it  could  to  transform  Langley 
into  an  R&D  center  ready-made  for  the  space  age.  But  aeronautical 
engineers  and  their  passion  for  airplanes  and  other  winged  flight  vehicles 
did  not  completely  disappear  at  the  center.  Floyd  Thompson  was  not  about 
to  let  aeronautics  die  at  the  historic  NACA  facility  where  he  had  worked 

96 


Change  and  Continuity 


Aeronautics  and  Space  Work  as  Percentages  of 
Langley's  Total  Effort,  1957-1965 


Effort 

1957 

1958 

1959 

1960 

1961 

1962 

1963 

1964 

1965 

Hypersonics 
Supersonics 
Subsonics 
Special  Types 

6 

12 
5 

8 

9 
16 
6 
9 

6 
13 

8 
9 

9 

13 

5 

7 

6 
16 
5 

7 

5 

15 
4 

7 

7 
12 
3 
4 

9 
12 
3 
3 

8 
10 
3 
3 

Aeronautics  Total 
Space  Total 

31 
69 

40 
60 

36 

64 

34 

66 

34 

66 

31 
69 

26 

74 

27 
73 

24 
76 

Source:      "Distribution    of   Effort"    pie    charts    in    folder    labeled    "Research    Effort," 
Laurence  K.  Loftin,  Jr.,  Collection,  Langley  Historical  Archives. 


since  before  the  Lindbergh  flight  and  where  so  many  ideas  important  to  the 
progress  of  American  aviation  had  been  born.22 

The  place  of  aeronautics  at  Langley  was  nevertheless  to  change  signifi- 
cantly in  the  wake  of  Sputnik.  For  the  NACA  to  metamorphose  successfully 
into  NASA,  aeronautics,  out  of  political  necessity,  had  to  give  up  the  center 
stage  that  it  had  enjoyed  for  over  five  decades  so  that  an  overnight  sensation 
could  now  dazzle  in  the  spotlight.  The  astronaut  rocketing  into  the  dark- 
ness of  space  would  now  get  top  billing;  the  aviator  flying  through  the  wild 
blue  yonder,  and  the  engineers  and  scientists  who  made  that  flight  possible, 
would  play  the  part  of  supporting  actors.  Already  by  the  spring  of  1958, 
aeronautics  at  Langley  made  up  only  40  percent  of  the  total  work  done  at 
the  center.  By  1965  aeronautics  would  plummet  to  its  lowest  point,  a  measly 
24  percent.  The  space  program  was  outshining  older  stars. 

For  aviation  enthusiasts,  this  turn  of  events  proved  traumatic.  Veteran 
aeronautical  engineer  Raymond  L.  Bisplinghoff,  who  directed  the  OART  at 
NASA  headquarters  from  1962  to  1966,  put  it  mildy  when  in  a  1983  memoir 
he  stated  that  the  formation  of  NASA  had 

a  dramatic,  and  at  first  deleterious,  influence  on  the  on-going  program  of  aeronautical 
research.  The  new  space  tasks  were  often  under  scientists  who  worked  on  a  space 
problem  for  one  week  then  switched  back  to  aeronautics  the  next  week.  .  .  . 
The  massive  priority  which  the  country,  from  the  president  on  down,  placed  on 
eclipsing  the  Russian  lead  in  spaceflight  had  a  profound  influence  on  the  NACA 
aeronautical  staff  as  they  assumed  positions  in  the  new  agency.  Many  took  advantage 
of  opportunities  to  move  to  higher  grades  and  levels  of  responsibility  in  space 
activities.  As  a  result,  many  moved  from  aeronautical  research  tasks  to  space  program 


management  tasks. 


23 


97 


Spaceflight  Revolution 

Others,  such  as  Langley's  fiery  director  of  aeronautics,  John  Stack,  were  so 
sure  that  the  first  "A"  in  NASA  was  being  erased  forever  that  they  decided 
to  leave  the  space  agency  entirely.  At  the  time,  especially  after  NASA's 
annual  R&D  budget  for  aeronautics  fell  below  a  million  dollars  in  1962,  these 
disillusioned  aviation  enthusiasts  could  not  have  known  how  extensively,  or 
how  successfully,  NASA  would  rebuild  its  aeronautics  programs  following 
its  major  buildup  for  space. 

In  the  late  1950s  and  early  1960s,  all  that  the  aviation  enthusiasts  could 
think  about  was  the  overwhelming  dominance  of  space  over  aeronautics. 
In  private,  many  Langley  aeronautical  engineers  held  NASA's  manned 
spaceflight  programs  in  contempt,  especially  the  quest  to  land  men  on  the 
moon,  believing  it  to  be  the  height  of  dishonesty  for  their  organization  to 
undertake  such  a  mission,  even  if  it  could  be  done,  when  it  was  not  worth 
doing.  John  V.  Becker,  a  talented  Langley  researcher  who  by  the  late  1940s 
had  already  shifted  his  attention  to  hypersonics  and  the  possibilities  of  an 
evolutionary  progression  into  space  via  transatmospheric  vehicles  like  the 
X-15,  remembers  that  his  longtime  colleague  John  Stack  was  "not  really 
much  interested  in  the  reentry  problem  or  in  space  flight  in  general."  For 
Stack,  even  the  X-15  was  a  program  barely  worth  supporting,  and  he  did  so 
"with  only  the  semblance  of  the  notorious  promotional  fire  he  could  generate 
if  he  was  really  interested."24 

John  Stack  and  his  team  of  aeronautical  engineers  reserved  their  enthu- 
siasm for  advanced  high-speed  military  jets  and  for  a  viable  commercial 
SST.  As  Becker  remembers  about  his  volatile  colleague,  Stack  developed 
"a  hostile,  adversary  attitude  towards  Space,  perhaps  because  it  threat- 
ened to  drain  resources  that  otherwise  might  belong  to  aeronautics."  When 
the  Apollo  program  was  established  in  1961,  Stack  told  Becker,  "I  don't 
buy  this  'to  the  Moon  by  noon'  stuff."  Unimpressed  by  the  great  size  and 
complexity  of  the  booster  rockets,  he  compared  von  Braun's  Saturns  to 
the  impressive  but  very  stationary  Washington  Monument  and  sided  with 
some  early  but  abortive  attempts  inside  NASA  to  find  viable  air-breathing 
aircraft-like  launch  systems  for  the  manned  space  missions.  According  to 
Becker,  Stack,  even  after  leaving  NASA  in  1962  for  an  executive  position 
with  Republic  Aviation,  "continued  to  favor  advanced  aircraft  as  opposed 
to  space  projects."25 

Most  members  of  the  Stack  team,  as  well  as  many  of  Langley's  other 
aviation  enthusiasts,  felt  exactly  the  same  way.  The  hard-core  aeronautical 
engineers  in  the  years  following  Sputnik  were  in  Mark  R.  Nichols'  Full- Scale 
Research  Division  and  in  Philip  Donely's  Flight  Mechanics  and  Technology 
Division,  both  of  which  were  part  of  Group  3.  Inside  the  wind  tunnels  and 
flight  hangars  of  these  two  divisions,  torrid  love  affairs  with  aerodynamics, 
with  high  lift/drag  ratios,  with  satisfactory  flying  and  handling  qualities, 
and  with  the  comely  shapes  and  exciting  personalities  of  airplanes  and 
helicopters  continued  to  flourish  long  after  the  formation  of  the  STG.  Far  too 
numerous  to  count  or  name  them  all,  the  strongest  adherents  to  aeronautics 

98 


Change  and  Continuity 


L-59-5435 


//  accomplished  high-speed  aerodynam- 
icist  John  Stack  looks  disgruntled  in 
his  staff  photo  from  1959,  it  may  be 
because  of  the  growing  predominance 
of  the  space  program  at  Langley. 


during  the  1960s  can  be  spotted  simply  by  looking  at  an  organization  chart 
or  thumbing  through  the  Langley  phonebook  noting  who  belonged  to  these 
divisions.  From  top  to  bottom,  these  men  were  the  "aero  guys." 

And  they  were  not  happy.  In  the  aftermath  of  the  Sputnik  crisis,  "there 
was  a  real  strong  emphasis  on  getting  people  out  of  aeronautics  and  into 
space,"  remembers  Mark  Nichols,  the  Full-Scale  Research  Division  chief.  In 
fact,  Nichols  himself  was  moved.  In  1959,  Floyd  Thompson  put  Nichols 
in  charge  of  Langley's  first  space  station  committee,  choosing  him,  one  of 
the  laboratory's  most  die-hard  aeronautical  engineers,  as  a  lesson  to  all 
others.  Laurence  K.  Loftin,  Jr.,  a  devoted  aeronautical  researcher  and 
aerodynamic  nutter  expert,  also  found  himself  immersed  in  planning  for  both 
space  stations  and  lunar  missions  in  the  early  1960s.  As  this  substitution 
pattern  became  clear,  the  air-minded  at  Langley  found  themselves  in  "an 
adversarial  mode  with  management,  which  was  always  trying  to  take  our 
people  and  put  them  into  space."  Nichols  and  his  buddies  looked  "for  ways 
of  resisting  this,"  but  were  not  successful.26 

No  one  was  unhappier  with  this  development  than  Langley's  number  one 
"aero  guy,"  John  Stack.  No  one  at  Langley  grew  more  disgruntled  over 
what  he  believed  the  space  agency  was  doing  to,  and  not  /or,  the  country's 
precious  aeronautical  progress.  Stack,  the  brilliant  and  outspoken  head  of 
aeronautics,  set  the  tone  for  the  numerous  dissatisfactions  of  the  air-minded 
engineers  at  the  center  during  the  first  years  of  the  space  revolution.  This 
was  true  especially  after  he,  one  of  the  most  decorated  and  powerful  men 
at  the  laboratory,  started  to  lose  out  in  some  infighting  within  the  Langley 
front  office.  Most  notably,  in  March  1961,  Thompson  made  Charles  Donlan, 

99 


Spaceflight  Revolution 


L-65-2870 

One  of  the  aeronautical  passions  at  Langley  in  the  late  1950s  and  early  1960s  was 
variable  wing  sweep,  a  technology  by  which  an  airplane 's  wings  could  be  mechanically 
adjusted  to  different  sweep  angles  to  conform  to  either  subsonic,  transonic,  or 
supersonic  flight  requirements.  In  this  photo  from  May  1965,  a  wind-tunnel  engineer 
checks  the  mounting  of  a  scale  model  of  the  General  Dynamics  F-lllA,  the  air 
force's  version  of  the  nation's  first  variable-sweep  fighter.  The  F-lllA  first  flew  in 
December  1964;  the  navy  version,  the  F-111B,  made  its  initial  flight  in  May. 

and  not  Stack,  Langley's  associate  director.  A  man  of  rare  accomplishments 
and  visionary  ambitions,  Stack  was  not  accustomed  to  being  passed  over. 
Not  even  NASA's  heavy  involvement  in  the  national  SST  program  could 
keep  Stack  working  for  the  space  agency. 

However  preoccupied  NASA  became  in  the  1960s  with  space-related  mat- 
ters, at  Langley  aeronautics  research  continued  and  resulted  in  outstanding 
contributions  to  everything  from  hypersonic  propulsion  to  the  handling  qual- 
ities of  general  aviation  aircraft.*  One  reason  for  the  unexpected  degree 
of  success,  ironically,  was  the  fact  that  aeronautics  did  not  receive  much 
attention  from  NASA  management  or  from  the  public  at  large.  The  Apollo 
program  and  all  its  related  activities  so  consumed  NASA  headquarters  that 
it  let  the  aeronautical  engineers  do  as  they  pleased.  In  this  sense,  the  aero- 


Originally,  I  planned  to  include  a  long  chapter  dealing  specifically  with  aeronautics.  As  the 
study's  thesis  evolved,  however,  I  realized  that,  although  the  spaceflight  revolution  certainly  affected  the 
aeronautics  efforts  in  many  significant  ways,  I  could  not  do  justice  to  the  complete  history  of  aeronautics 
at  Langley  during  the  1960s  within  the  confines  of  an  already  long  book.  So,  I  decided  not  to  cover  the 
aeronautical  programs  and  leave  them  for  separate  treatment  at  some  later  date,  perhaps  by  someone 
other  than  myself. 

100 


Change  and  Continuity 


L-70-1386 

In  this  photo  from  1970,  a  technician  readies  a  model  of  Langley's  own  pet  Super- 
sonic Commercial  Air  Transport,  known  as  SCAT  15F  ("F"  for  fixed  wing),  for 
testing  in  the  Unitary  Plan  Wind  Tunnel. 

nautical  work  at  Langley  kept  much  the  same  personality  and  flavor  as  dur- 
ing the  NACA  era  when  the  engineers,  not  the  bureaucrats  in  Washington, 
had  been  in  charge. 

But  oddly,  in  retrospect,  the  rearguard  of  aviation  enthusiasts  at  Langley 
in  the  1960s  in  some  ways  resembled  the  vanguard  of  the  spacefiight 
revolution.  As  much  as  the  air-minded  hated  NASA's  emphasis  on  space 
exploration,  these  air-minded  engineers  and  scientists  nevertheless  became 
equally  caught  up  in  their  own  dreams  of  monumental  new  accomplishments. 
Perhaps  it  was  just  the  nature  of  the  revolutionary  times  brought  on  by 
Sputnik  and  President  Kennedy's  New  Frontier  to  think  so  grandly  and  to 
feel  that  the  old  limitations  no  longer  applied. 

NASA's  aeronautical  engineers  had  their  own  Apollo  program  in  the 
1960s:  the  design  of  the  most  revolutionary  aircraft  ever  built — a  commer- 
cial supersonic  airliner  capable  of  flying  two  or  even  three  times  the  speed  of 
sound  and  crossing  the  Atlantic  from  New  York  to  London  or  Paris  in  a  few 
hours.  This  dream  compared  favorably  with  the  lunar  landing  because  an 
SST  would  have  such  immense  economic,  political,  and  social  significance 
that  it  would  change  how  humankind  traversed  the  face  of  the  earth.  The 
Apollo  program  would  accomplish  nothing  similar.  Neither,  of  course,  would 
the  aeronautical  engineers'  national  SST  program  because  the  U.S.  Congress 
killed  it  in  1971.  In  this  sense,  too,  the  space  cadets  emerged  "one  up,"  for 


101 


Space/light  Revolution 

they  had  their  spectacular  moment  with  the  manned  lunar  landings;  the 
"aero  guys"  never  did. 


Growth  Within  Personnel  Ceilings 

Despite  the  dissolution  of  the  Hydrodynamics  Division  and  the  wane  of 
aeronautics,  Langley's  formal  organization  did  not  change  significantly  in 
the  early  1960s.  This  was  in  part  because  the  center  did  not  grow  much 
bigger.  By  the  changeover  to  NASA,  Langley  Research  Center  was  already 
a  large  operation.  It  had  greatly  expanded  during  its  NAG  A  history  from 
a  few  small  buildings  in  an  isolated  corner  of  the  military  base  prior  to 
World  War  II  to  a  710-acre  complex  on  both  sides  of  the  air  force  runways. 
It  was  now  an  establishment  that  included  30  major  wind  tunnels  and 
laboratories  and  whose  replacement  worth  to  the  federal  government  was 
estimated  at  nearly  $150  million.  In  1958  the  center  paid  approximately 
$6  million  in  operating  expenditures,  including  nearly  $2  million  just  for 
electric  power.  Its  annual  payroll  stood  at  $22  million.  Its  full-time  civil 
service  staff  numbered  about  3300,  of  whom  approximately  one-third  were 
engineers,  scientists,  mathematicians,  and  other  professional  people.27 

With  the  transition  to  NASA,  the  size  of  the  Langley  staff  actually 
became  a  little  smaller  before  it  grew  any  larger:  from  3795  paid  employees 
in  June  1959  to  3456  by  the  end  of  that  year.  The  staff  fell  to  3191  six  months 
after  the  previously  auxiliary  Wallops  Station  became  an  independent  field 
installation  (on  1  January  1960).  In  the  next  three  years,  the  number  rose 
to  4007.  By  June  1966  the  Langley  staff  reached  its  all-time  high  of  4485 
employees.  This  was  nearly  1000  more  staff  members  than  Langley's  peak 
number  in  1952.  But  relative  to  the  agency  wide  growth  of  NASA  in  the 
1960s,  Langley's  expansion  was  actually  quite  moderate. 

In  1958,  Langley's  3300  employees  represented  more  than  41  percent  of 
NASA's  total  first-year  civil  service  complement  of  7966.  But  in  1964,  the 
4329  employees  of  the  Virginia  facility  amounted  to  barely  13  percent  of  the 
agency's  entire  number,  which  in  a  span  of  just  five  years  had  doubled  and 
doubled  again,  to  33,108.  In  other  words,  while  Langley  was  growing,  its  rate 
of  growth  was  slow  compared  with  NASA's.  At  this  rate  Langley  would  be 
unable  to  retain  its  traditional  position  of  dominance  in  the  agency.  NASA 
was  adding  large  new  manned  spaceflight  centers  such  as  Marshall  Space 
Flight  Center  in  Alabama,  the  Manned  Spacecraft  Center  in  Texas,  and  the 
Launch  Operations  Center  (in  November  1963,  renamed  the  Kennedy  Space 
Center)  at  Cape  Canaveral  in  Florida.  The  addition  of  Marshall  alone  had 
meant  the  mass  influx  of  over  4000  personnel  from  the  U.S.  Army  as  part  of 
the  transfer  of  the  ABMA's  Development  Operations  Division  to  NASA. 

Moreover,  NASA's  total  personnel  headcount  of  33,108  in  1964  repre- 
sented a  diminishing  fraction  of  NASA's  overall  effort.  In  the  late  1960s, 
NASA  estimated  that  Project  Apollo  employed  some  400,000  Americans 

102 


4500,- 


4250 


4000 


Number 
of  paid      3750 
employees 

3500 


3250 


3000 


Change  and  Continuity 


1952 


June 
1959 


Dec. 
1959 
Year 


June 
1960 


Dec. 
1962 


June 
1966 


Number  of  paid  employees  at  NASA  Lang  ley,  1952-1966. 


45 
40 
35 

30 

Percentage 
of  NASA         25 
total 

20 

15 

10 

5 

0 


1958    1959    1960    1961    1962    1963    1964    1965    1966   1967    1968 

Year 
Paid  employees  at  NASA  Langley — percentage  of  NASA  total,  1958-1968. 


103 


Space/light  Revolution 

in  government,  industries,  and  universities.  NASA's  civil  service  employees 
amounted  to  just  a  little  more  than  10  percent  of  the  total  NASA  work  force, 
broadly  defined.  The  other  90  percent  were  contractors. 

Langley  was  often  called  "Mother  Langley"  because  it  had  been  the 
mother  lode  for  all  NACA  facilities.  A  guiding  force  throughout  NACA 
history  and  for  the  first  years  of  NASA,  Mother  Langley  now  was  losing 
its  central  position  in  the  agency.  Although  only  a  few  concerned  research- 
oriented  people  like  Hugh  Dryden  would  have  thought  about  the  significance 
of  the  changing  personnel  numbers  at  the  time,  they  were  symptomatic  of  a 
slow  but  sure  decline  of  the  formerly  predominant  influence  of  the  research 
centers  and  the  coming  hegemony  of  the  development  centers.  The  personnel 
numbers  signified  the  ascendancy  of  organizations  devoted  primarily  not  to 
research  but  to  planning  and  conducting  actual  spaceflight  operations  and 
building  hardware. 

The  trend  did  not  go  unnoticed.  Thompson  received  a  forewarning  of  the 
siphoning  of  research  center  staff  funds  for  development  centers  from  Earl 
Hilburn,  whose  appointment  as  a  NASA  deputy  associate  administrator 
Thompson  had  summarily  discounted.  On  9  September  1963,  Thompson 
sent  a  four-page  letter  to  NASA  headquarters  regarding  Langley's  personnel 
requirements.  His  letter  underscored  what  he  called  "the  problem  of 
manpower  distribution"  among  the  NASA  centers.  "The  immediate  needs 
of  a  development  program  are  always  more  easily  recognized,"  he  began, 
"than  is  the  requirement  for  a  continuing  research  program  that  lays  the 
basic  foundation  of  technology  upon  which  the  development  program  can 
continually  depend  for  guidance  in  solving  detailed  technical  problems."28 

At  the  heart  of  Thompson's  illuminating  letter  was  his  concern  about  an 
ongoing  tug-of-war  between  the  manned  spaceflight  centers  and  the  research 
centers  over  the  apportionment  (or  reapportionment)  of  NASA's  personnel 
quotas.  The  internal  struggle,  which  the  research  centers  were  losing,  was 
the  result  of  work-load  stresses  caused  by  the  ceilings  that  were  imposed 
on  the  total  number  of  people  NASA  could  employ.  The  way  the  system 
worked,  the  agency  asked  for  the  amount  of  money  it  needed  to  pay  salaries 
based  on  the  number  of  people  it  anticipated  it  would  employ.  However,  if 
Congress  or  the  Bureau  of  the  Budget  found  reason  to  trim  the  request,  then 
NASA  had  to  cut  back  on  its  staffing  projections  accordingly,  even  though 
the  requirement  to  do  so  was  not  explicit  in  the  appropriation  act.29 

In  the  first  years  of  NASA,  this  sort  of  cutting  back  had  happened  fre- 
quently because  Congress,  the  Bureau  of  the  Budget,  President  Eisenhower, 
and  even  NASA  Administrator  Glennan  hoped  to  keep  a  rather  tight  lid  on 
civil  service  staffing.  For  Glennan  as  well  as  for  many  others,  keeping  the  lid 
on  the  personnel  total  played  directly  into  the  Republican  philosophy  that 
government  was  already  too  big.  At  a  NASA  staff  conference  in  Monterey, 
California,  in  early  March  1960,  Glennan  claimed  that  "there  was  a  need  for 
some  kind  of  arbitrary  limitation  on  NASA's  size.  By  limiting  the  number 
of  employees,  NASA  would  limit  its  in-house  capability  and  thus  be  forced 

104 


Change  and  Continuity 


The  first  woman  to  work  as  an  en- 
gineer at  NACA  Langley  was  Kitty 
O'Brien-Joyner  (left),  who  was  also 
the  first  female  to  graduate  with  an 
engineering  degree  from  the  Univer- 
sity of  Virginia. 


-64-5210 


L-59-7 

In  1959,  Langley  employed  six  women  who  were  classified  by  NASA  as  "scientists." 
During  the  Apollo  era,  women  made  up  3  to  5  percent  of  the  professional  work  force 
agencywide.  The  percentage  of  African  American  professionals  was  significantly 
smaller,  from  1.5  to  3  percent.  These  percentages  rose  slowly  for  both  groups  as  the 
decade  proceeded. 


105 


Space/light  Revolution 

to  develop  the  capabilities  of  contractors."30  This  development  would  be  far 
better  for  the  American  economy  than  hiring  larger  coteries  of  government 
workers.  Glennan  sanctioned  relatively  low  personnel  ceilings.  For  fiscal 
year  1962,  for  example,  he  approved  a  limit  of  16,802  employees,  which  was 
less  than  3  percent  above  the  total  authorized  for  the  previous  year.31  Nat- 
urally, no  NASA  center  director  facing  the  high  public  expectations  and 
enormously  expanded  work  load  of  the  early  1960s  could  be  expected  to  be 
happy  about  such  limits  on  hiring. 

The  acceleration  of  the  space  program  brought  on  by  President  Kennedy 
and  his  dynamic  new  man,  James  Webb,  jacked  the  personnel  ceiling  up  to 
new  heights.  Instead  of  the  3-percent  increase  for  fiscal  year  1962  proposed 
by  Glennan,  an  increase  of  a  whopping  43  percent  was  approved.  Between 
1961  and  1965  the  total  number  of  agency  personnel  would  double,  from 
17,471  to  35, 860. 32  Given  this  rapid  growth  in  the  size  of  the  NASA 
staff,  it  may  seem  more  than  a  bit  astonishing  to  find  a  NASA  center 
director  worried  about  the  need  for  more  personnel.  But  by  late  1963, 
that  was  the  case.  Government  controls  on  personnel  totals  even  during  the 
ensuing  Democratic  administrations  of  Kennedy  and  Lyndon  B.  Johnson 
were  such  that  the  only  way  to  take  care  of  any  unforeseen  requirements 
that  occurred  during  a  fiscal  year  was  to  transfer  manpower  and  related 
financial  resources  among  institutions.  And  when  a  transfer  was  needed, 
the  research  laboratories  invariably  lost. 

On  more  than  one  occasion,  subsequent  to  a  preliminary  formulation 
of  the  basic  data  regarding  the  agency's  manpower  requirements  at  NASA 
headquarters,  the  managements  of  Houston  and  Huntsville  would  request  a 
substantial  number  of  supplementary  personnel.  (Earl  Hilburn  was  warning 
Thompson  about  such  a  request  in  September  1963.)  To  give  the  space 
centers  several  hundred  additional  staff  positions  without  obtaining  the 
congressional  authorization  to  increase  the  agency's  overall  complement 
meant  that  NASA  headquarters  had  no  other  choice  but  to  reapportion 
the  personnel  quotas  among  the  field  centers.  In  other  words,  in  order  for 
Houston  and  Huntsville  to  get  more,  Langley  and  other  research  centers 
would  have  to  get  less.33 

In  Thompson's  mind,  this  was  a  tug-of-war  that  the  research  centers, 
given  the  priorities  of  the  space  race,  could  not  win,  but  which  the  nation 
could  not  afford  to  lose.  "Two-thirds  of  the  current  total  effort  of  Langley 
is  utilized  in  support  of  the  NASA  space  effort,"  Thompson  wrote  to  the 
NASA  administration.  "These  programs  have  been  prepared  in  cooperation 
with  and  approved  by  the  OART  and  other  cognizant  program  offices. 
They  have  been  endorsed  by  NASA  as  essential  to  continued  leadership  in 
space  exploration  and  vital  to  the  success  of  such  basic  NASA  programs 
as  Saturn,  Gemini,  and  Apollo."34  To  support  this  claim,  he  attached 
to  his  letter  (along  with  lists  and  charts  illustrating  "the  wide  range 
of  activity"  at  Langley)  a  10-page  document  listing  all  the  then-current 
Langley  investigations  relevant  to  the  program  interests  at  the  Manned 

106 


Change  and  Continuity 

Spacecraft  Center.  This  document,  prepared  by  Axel  T.  Mattson,  whom 
Thompson  had  dispatched  as  a  special  attache  to  Houston  in  the  summer  of 
1962,  demonstrated  that  Langley  was  spending  some  300  man- years  just  in 
support  of  the  Texas  center's  projects.35  If  NASA  continued  to  neglect 
Langley 's  manpower  needs  and  persisted  in  improperly  distributing  the 
quotas,  something  would  have  to  give.  Too  few  people  would  remain  at 
the  research  center  to  perform  the  total  center  mission.  Either  Langley 
would  do  all  the  support  work,  leaving  little  if  any  time  for  fundamental 
research,  or  the  support  work  would  have  to  subside,  thus  putting  the  goals 
of  the  American  space  program  at  risk. 


The  Shift  Toward  the  Periphery 

The  trend  pushing  Langley  from  the  center  of  NASA  toward  its  periphery 
is  evident  not  only  in  the  personnel  numbers  but  also  in  the  budget  figures. 
In  1959  the  direct  cost  of  Langley's  administrative  operations  in  terms  of  its 
obligations  to  pay  employees  and  honor  all  those  contracts  (not  including 
Wallops')  that  were  not  funded  by  R&D  money  was  $30.7  million.  This 
amounted  to  36  percent  of  the  NASA  total.  In  1967,  Langley  spent  $64.3 
million,  the  most  money  it  would  spend  on  operations  during  any  one  year  in 
its  entire  history;  however,  this  amount  was  less  than  10  percent  of  the  NASA 
total  for  that  year.  In  1959  the  cost  of  running  Langley  was  significantly 
higher  than  that  of  operating  any  other  NASA  facility.  But  by  1967,  Langley 
was  down  to  seventh  place  on  that  list,  while  Marshall  stood  at  the  top,  at 
$128.7  million,  or  double  what  it  cost  to  operate  Langley.  Even  the  price  of 
running  NASA  headquarters  was  nearly  up  to  the  Langley  figure.  Whereas 
$5.5  million  kept  the  offices  in  Washington  going  in  1959,  by  1967  that  figure 
had  shot  up  over  tenfold  to  a  grandiose  $57  million.36 

NASA  headquarters  was  growing  by  leaps  and  bounds  in  the  early  1960s. 
It  was  a  larger,  more  multilayered,  and  more  active  bureaucracy  than  had 
ever  been  the  case  for  the  NACA's  Washington  office.  A  host  of  headquarters 
officials  congealed  and  took  charge  of  all  the  programs  at  Langley  and  the 
other  NASA  centers.  This  meant  that  the  field  centers  had  to  work  through 
Washington  not  only  for  their  allotment  of  resources  but  also  for  many 
levels  of  program  initiation  and  administration.  Also  unlike  the  days  of 
the  NACA,  the  bureaucrats  in  Washington  were  now  directly  in  charge 
of  their  own  little  empires.  They  issued  major  contracts  to  universities 
and  industries  for  R&D  and  for  design  studies.37  Between  1960  and  1968, 
the  value  of  contracts  awarded  by  NASA  headquarters  rose  from  3  to  11 
percent  of  the  total  value  of  contracts  awarded  agencywide.  During  the 
same  period,  Langley  experienced  a  decline  from  35  to  3  percent  of  the 
total  value  of  contracts  agencywide,  and  Huntsville  and  Houston  centers 
collectively  hovered  consistently  between  50  and  60  percent  of  the  NASA 
total.38 

107 


Space/light  Revolution 

Compared  with  the  megabucks  turned  over  to  the  spaceflight  centers  for 
R&D  during  this  period,  Langley's  funding  was  also  relatively  small.  In 
1963  the  center  received  less  than  2  percent  of  the  total  money  set  aside  by 
NASA  for  R&D  programs.  On  the  other  hand,  Marshall  received  almost  30 
percent  of  the  total  NASA  R&D  budget.  The  most  Langley  ever  received 
in  R&D  funding  was  $124  million  in  1966;  the  least  that  Marshall  received 
in  the  same  period  was  10  times  that  in  1968. 39 

The  point  of  going  through  these  numbers  is  not  to  show  that  Langley 
was  being  treated  unfairly.  As  a  facility  devoted  primarily  to  applied  basic 
research  in  aviation  and  space,  Langley  simply  was  not  doing  as  much 
procurement  as  were  those  NASA  centers  responsible  for  designing,  building, 
launching,  and  operating  spacecraft.  What  the  numbers  do  show  is  a  new 
technological  order  brought  on  by  the  spaceflight  revolution.  In  examining 
the  numbers  we  hold  up  a  mirror  to  the  new  sociopolitical  context  of  research 
activities  at  the  former  NACA  aeronautics  laboratory.  The  mirror  reflects 
NASA's  determination  to  allocate  the  lion's  share  of  its  financial  resources 
to  those  arms  of  the  agency  most  directly  involved  in  what  the  country 
was  intent  on  achieving  through  its  space  program.  In  the  1960s  that  was, 
first,  getting  astronauts  into  orbit  around  the  earth;  second,  per  President 
Kennedy's  May  1961  commitment,  landing  American  astronauts  on  the 
moon;  and  third,  in  the  process,  refreshing  the  nation's  spirit,  reinvigorating 
its  economy,  and  showing  the  world  just  what  the  U.S.  system  of  democracy 
and  free  enterprise  could  do  when  the  American  people  put  their  minds  and 
energies  to  it.  In  other  words,  the  intent  was  to  win  the  space  race. 

These  figures  signify  more  specifically  the  rather  immediate  effects  that 
NASA's  broader  mission  had  on  the  lives  of  the  old  NACA  research  labora- 
tories. Unlike  the  NACA,  NASA  would  be  an  operational  organization,  not 
just  a  research  organization.  It  would  become  heavily  involved  in  projects 
with  goals  and  schedules  and  it  would  contract  out  to  American  business 
and  industry  a  great  part  of  its  work.  As  this  happened,  Langley  staff  feared 
that  administrators  in  Washington  would  no  longer  see  the  center  as  special. 
With  headquarters  now  running  many  of  its  own  shows  through  contracts 
to  industry,  a  place  like  Langley  could  come  to  be  regarded  by  many  at 
headquarters  as  just  another  contributor  to  the  program.  Langley  was  just 
one  more  possible  center  where  work  could  be  done,  if  NASA  headquarters 
chose  to  locate  it  there.  But  headquarters  might  instead  choose  the  General 
Electric  Company's  Command  Systems  Division;  BellComm,  Incorporated; 
the  Douglas  Aircraft  Company;  Thompson- Ramo-Wooldridge  (TRW);  MIT; 
or  some  other  very  capable  organization.40  Langley  was  now  for  the  first 
time  in  competition  with  "outsiders,"  the  many  laboratories  and  firms  that 
had  been  springing  up  or  growing  in  competency  in  conjunction  with  the 
burgeoning  of  the  "military-industrial  complex"  after  World  War  II. 

The  competition  was  not  inherently  harmful  to  Langley.  Given  the  ample 
budgets  brought  on  by  the  spaceflight  revolution,  NASA  had  more  money 
than  it  could  spend  on  itself  or  on  its  research  laboratories.  Langley  was 

108 


Change  and  Continuity 

simply  not  accustomed  to  the  competition,  and  it  was  not  accustomed 
to  relying  on  others.  For  more  than  four  decades  its  organization  had 
been  largely  self-sufficient.  As  an  internal  Langley  study  on  the  history 
of  contracting  at  the  NASA  center  by  Sarah  and  Steve  Corneliussen  has 
noted,  the  laboratory  staff  had  almost  always  conducted  its  own  research, 
built  its  own  models  and  instrumentation  and  wind  tunnels,  and  handled  its 
own  logistical  needs,  from  mowing  the  grass  to  operating  its  two  cafeterias. 
Only  occasionally  had  outsiders  been  brought  in  during  the  NACA  period 
to  augment  the  civil  service  staff — and  "only  temporarily  at  that,  just  to 
help  out  with  occasional  peaks  in  the  center's  housekeeping  workload."41 

Thus,  many  former  NACA  staffers  would  need  time  to  adjust  to  the  new 
environment  of  NASA  and  to  see  that  the  involvement  of  outsiders  in  the 
work  of  the  new  space  agency  would  not  take  anything  away  from  their 
historic  capabilities  or  their  tradition  of  self-sufficiency,  but  would  instead 
add  to  them.  "Contracting  out"  was  not  substituting  the  work  of  others  for 
what  the  in-house  staff  had  always  done.  It  was  augmenting  the  capabilities 
of  the  NASA  researchers  so  that  they  could  accomplish  more.  The  Langley 
organization  would  be  no  less  cohesive  nor  would  contracting  damage  its  best 
qualities;  it  would  only  enhance  them.42  That,  at  least,  was  the  argument. 

Contracting  Out 

Other  than  the  occasional  employment  of  temporary  laborers  for  odd 
jobs,  Langley  had  accomplished  almost  everything  it  had  to  do  with  its  own 
staff.  This  self-sufficiency  worked  well  during  the  NACA  period  because  the 
range  of  what  needed  to  be  done  was  usually  narrow  enough  for  the  civil 
service  work  force  to  handle  it.  If  the  work  load  increased  significantly,  as 
during  World  War  II,  then  the  solution  was  to  obtain  authorization  from 
Congress  for  additional  civil  service  staffing.  The  answer  was  not  to  hire 
contractors. 

With  the  quickening  pace  of  the  space  race  and  the  urgency  of  NASA's 
expanded  mission,  however,  the  work  load  increased  so  dramatically  that 
civil  service  staffing  authorizations  could  not  keep  up.  An  evolving  mismatch 
between  the  high  work  load  at  the  research  center  and  the  low  level  of 
congressional  authorizations  for  more  research  staff  eventually  forced  a 
reluctant  Langley  into  contracting  out  for  much  of  the  work  that  it  always 
had  done  and  would  have  preferred  to  continue  doing  itself. 

At  first  the  research  center  resisted  the  trend  toward  contracting  out 
and  was  only  willing  to  hand  over  to  outsiders  mundane  maintenance  and 
administrative  jobs,  such  as  delivering  the  mail,  operating  the  cafeterias, 
running  the  center's  credit  union,  and  maintaining  some  of  the  warehouses. 
Procurements  for  these  jobs  involved  so-called  support  service  contracts, 
that  is,  binding  legal  relationships  drawn  up  so  that  the  time  and  the  services 
of  an  outside  firm  (i.e.,  the  contractor)  could  be  secured  to  attain  a  specified 
in-house  objective.43 

109 


Space/light  Revolution 

Even  for  the  tasks  of  routine  housekeeping,  Langley  wanted  the  best 
employees.  "If  we're  going  to  hire  outsiders,"  the  procurement  officers  em- 
phasized, "then  let's  choose  a  way  of  doing  so  that  maximizes  their  contri- 
butions as  adjunct  members  of  the  team.'4  The  best  way  to  do  this,  they 
found,  was  to  use  a  "cost-plus-award  fee,"  a  special — and  for  the  government, 
novel — form  of  cost-reimbursement  contract.  In  Langley's  opinion,  this  ar- 
rangement had  the  highest  potential  for  inducing  quality  in  the  contractor's 
performance  because  the  contractor's  profit — the  award  fee — rises  or  falls 
in  direct  correspondence  to  the  customer's  (i.e.,  Langley's)  appraisal  of  the 
work.  As  with  straight  cost  reimbursement,  the  expense  to  the  government 
is  not  preset,  but  changes  over  time  with  the  changing  circumstances  of  the 
work.  This  process  differentiates  both  cost-reimbursement  and  cost-plus- 
award  fee  contracts  from  the  more  typically  used  "fixed-price"  contract,  in 
which  the  contracting  party  specifically  delineates  the  job  requested  and  the 
time  allowed  for  completing  it,  and  the  bidder  assumes  the  risk  of  match- 
ing the  forecast  of  the  demands  of  the  job  to  what  those  demands  will  in 
fact  turn  out  to  be.  However,  in  Langley's  case  of  contracting  for  ongoing 
support  services  usually  for  terms  of  several  years,  during  which  working 
circumstances  would  change  and  jobs  would  have  to  be  adjusted,  the  fixed- 
price  approach  would  not  work.45 

In  essence,  the  cost-plus-award  fee  was  an  incentives  contract;  according 
to  a  formal  NASA  definition,  it  provided  for  "a  basic  fixed  fee  for  perfor- 
mance to  a  level  deemed  acceptable,  plus  an  additional  award  fee,  not  in 
excess  of  a  stipulated  maximum,  for  accomplishment  of  better  than  the  'ac- 
ceptable' level."46  Its  downside  was  the  administrative  burden.  The  amount 
of  the  award  was  linked  to  the  contractor's  performance;  thus,  on  a  regular 
and  in  some  cases  almost  daily  basis,  responsible  Langley  employees  had 
to  inspect  and  evaluate  the  contractor's  work.  A  board  of  senior  managers 
had  to  appraise  the  contractor's  performance  at  agreed-upon  intervals  and 
decide  the  amount  of  extra  money  deserved.  A  much  larger  and  more  for- 
mal mechanism  for  handling  contractors  therefore  had  to  be  developed  at 
Langley.  One  clear  indicator  of  the  burden  of  this  added  responsibility  was 
the  growth  in  the  size  of  the  Langley  procurement  staff  itself.  Before  NASA 
replaced  the  NACA,  this  staff  comprised  25  people.  After  the  changeover, 
the  staff  quickly  expanded  to  more  than  100  before  leveling  off  at  70  to  80 
after  the  STG  left  for  Houston.47 

In  this  fashion,  Langley  did  what  it  could  to  bring  out  the  best  in 
its  contractors  and  to  make  them  feel  a  vital  part  of  the  center.  This 
method  of  contracting  was  a  way  of  bringing  outsiders  "in,"  of  making 
"them"  part  of  "us."  However,  an  inherent  and  potentially  serious  difficulty 
existed  in  carrying  out  the  philosophy  of  these  contracts.  Like  all  other 
procurements  by  the  U.S.  government,  these  contracts  for  support  services 
were  governed  by  federal  regulations.  The  regulations  clearly  allowed,  and 
the  then-current  federal  policy  indeed  encouraged,  the  direct  involvement  of 
American  businesses,  industries,  and  universities  at  government  facilities  like 

110 


Change  and  Continuity 


L-59-352 

The  Langley  division  most  assisted  by  support- service  contractors  in  the  early 
1960s  was  ACD.  By  mid- decade,  contractors  were  programming  the  computers  and 
handling  the  hardware  and  software  support  of  the  mainframe  systems,  and  by  1970, 
contractors  were  contributing  substantially  to  the  development  of  computer  programs 
for  the  guidance,  navigation,  and  control  of  aircraft  and  spacecraft. 

In  this  photo,  taken  in  1959,  engineers  are  at  work  in  Langley 's  computer 
complex.  Langley 's  electronic  analog  brain  (seen  in  photo)  with  its  plugboards  and 
vacuum  tubes  was  replaced  in  1965  by  mainframe  digital  computers.  The  conversion 
from  analog  to  digital  was  a  major  technological  development  of  the  spaceflight 
revolution.  Without  it  the  on-board  navigation  and  control  needed  to  achieve  the 
manned  lunar  landing  would  have  been  impossible. 

Langley,  but  the  same  body  of  regulations  also  insisted  that  the  contractors 
make  their  contributions  at  arm's  length  from  civil  service  management.  In 
other  words,  the  two  could  not  be  "in  bed  together."  If  civil  servants  did 
not  maintain  this  distance,  the  contractors  might  become  entrenched,  their 
expense  charges  could  get  out  of  hand,  and  they  would  essentially  have  a 
"license  to  steal."48 

Over  the  years,  despite  the  best  intentions  of  government,  Langley 
staff  would  have  trouble  adhering  always  to  the  arm's- length  requirement. 
Because  Langley  wanted  to  make  the  contractors  feel  that  they  were  part  of 
"the  family"  and  in  spirit  no  different  from  any  other  employee,  staff  could 
hardly  treat  contractors  in  the  formal,  mechanical  ways  required  by  the  rules. 
Contracting  officials  were  supposed  to  follow  a  labyrinth  of  procedures  and 
policies  to  arrive  at  the  letter  of  the  law  required  by  federal  procurement. 
But  as  civil  servants  and  contractors  worked  side  by  side,  ate  lunch  together 
in  the  NASA  cafeteria,  and  often  became  close  friends,  feelings  that  Langley 
should  keep  to  the  spirit  of  the  law,  as  opposed  to  the  letter,  prevailed.  As 
a  result,  the  position  of  the  contractors  at  Langley  slowly  grew  stronger. 


Ill 


Spaceflight  Revolution 

Starting  with  the  assignment  of  the  Scout  booster  rocket  project  to  the 
center  in  the  late  1950s,  as  Chief  Procurement  Officer  -Sherwood  Butler 
recalls,  "Langley  began  to  branch  out  and  contract  for  some  highly  technical 
services  such  as  launch  support,  support  of  research,  and  maintenance 
and  calibration  of  instrumentation."49  Several  representatives  of  the  prime 
contractor,  Ling-Temco-Vought  (LTV)  worked  on-site  on  a  daily  basis  as 
integral  members  of  the  Scout  "team."  These  contractors  included  12  LTV 
engineers  working  specifically  in  the  field  of  instrumentation.  Bringing  in 
instrumentation  experts  amounted  to  "the  first  instance  of  support  services 
contracting  in  a  technical  field  at  Langley."50  With  the  start  of  other  major 
projects  like  Fire  and  Lunar  Orbiter,  many  contract  employees  of  industrial 
firms  came  to  work  at  the  center  and  were  such  an  integral  part  of  the  team 
that  they  could  not  be  distinguished  from  the  government  workers  without 
a  glance  at  their  ID  badges. 


The  Brave  New  World  of  Projects 

In  the  brave  new  world  brought  on  by  the  spaceflight  revolution,  Langley, 
as  we  have  seen  in  its  support  of  the  STG,  for  the  first  time  became 
heavily  involved  in  project  work  and  the  formal  management  of  large- 
scale  endeavors  involving  hardware  development,  flight  operations,  and 
the  administration  of  contracts.  For  some  of  these  projects,  Langley 
personally  handled  the  reins  of  management  for  NASA  headquarters  as 
the  designated  "lead  center."  In  the  early  1960s  such  projects  included 
Scout,  which  began  in  1960  for  the  development  of  NASA's  first  launch 
vehicle,  a  dependable  and  relatively  inexpensive  solid-propellant  rocket; 
Radio  Attenuation  Measurements  (RAM),  which  came  to  life  in  1961  to 
address  the  radio  blackout  that  occurred  during  a  spacecraft's  reentry  into 
the  atmosphere;  Fire,  which  was  started  in  1962  to  study  the  effects  of 
reentry  heating  on  Apollo  spacecraft  materials;  Lunar  Orbiter,  which  was 
initiated  in  1963  to  take  photographic  surveys  of  the  moon  in  preparation  for 
the  Apollo  manned  lunar  landings;  and  the  Hypersonic  Ramjet  Experiment 
Project,  which  began  in  1964  to  explore  the  feasibility  of  a  hypersonic  ramjet 
engine. 

Other  NASA  organizations  took  the  lead  for  many  other  projects,  and 
Langley  helped  by  providing  diversified  R&D  support.  Langley  contributed 
in  this  way  to  all  the  manned  spaceflight  projects,  from  Project  Mercury 
through  Apollo.  Langley  also  participated  in  "cooperative  projects."  These 
were  projects  for  which  NASA  headquarters  assigned  the  overall  project 
management  to  another  center  but  gave  Langley  the  official  responsibility 
for  subsidiary  projects  or  for  specific  project  tasks.  The  earliest  example 
of  a  cooperative  project  involving  Langley  was  Project  Echo,  which  was 
started  in  1959  for  the  development  of  a  passive  communications  satellite. 
For  Project  Echo,  NASA  assigned  the  project  management  not  to  Langley 

112 


Change  and  Continuity 

but  to  Goddard;  however,  Langley  was  responsible  for  the  development  of 
the  Echo  balloon,  for  the  container  in  which  the  balloon  was  carried  into 
space,  and  for  the  balloon's  in-space  inflation  system.  Beyond  that,  Langley 
was  also  responsible  for  managing  two  flight  projects  in  support  of  Echo, 
Projects  Shotput  and  Big  Shot,  which  were  designed  to  test  Echo  designs 
under  suborbital  conditions  before  the  balloons  were  launched  into  orbit. 

Before  exploring  the  history  of  NASA  Langley's  early  involvement  in 
project  work  in  subsequent  chapters  of  this  book,  I  want  to  address  a  few 
basic  points  about  projects  and  about  research.  A  project  sets  out  to  do 
something  quite  specific  and  to  do  it  in  a  limited  time  frame.  For  example, 
the  goal  of  the  Manhattan  Project  during  World  War  II  was  the  design 
and  construction  of  an  atomic  bomb;  the  goal  of  Project  Sherwood  in  the 
1950s,  as  mentioned  in  the  next  chapter,  was  the  design  and  construction 
of  an  effective  fusion  reactor.  To  fulfill  these  objectives,  the  projects' 
researchers  had  to  move  ahead  quickly  and  adhere  to  strict  schedules.  They 
could  not  afford  many  detours.  The  Manhattan  Project  started  in  1941 
and  concluded  in  1945.  To  achieve  the  project  goal  in  those  four  years, 
a  vast  array  of  resources  had  to  be  effectively  mobilized,  organized,  and 
supplied.  The  enormously  complex  task  of  creating  the  first  atom  bomb 
would  not  have  been  successful  if  the  U.S.  government  and  its  wartime 
military  establishment  had  not  given  high  priority  to  completing  such  a 
"crash  effort."  With  a  far  lower  priority  and  with  more  intractable  problems 
to  solve,  Project  Sherwood  staff  never  did  achieve  the  project's  final  goal.51 

In  its  bare  essentials,  a  NASA  project  was  no  different  from  the  two 
projects  discussed  above.  According  to  a  formal  NASA  definition  in  the 
early  1960s,  a  project  was  "an  undertaking  with  a  scheduled  beginning  and 
end,"  which  involved  "the  design,  development,  and  demonstration  of  major 
advanced  hardware  items  such  as  launch  vehicles  or  space  vehicles."  The 
purpose  of  a  NASA  project  was  to  support  the  activities  of  a  program.  NASA 
defined  a  program  as  "a  related  series  of  undertakings  which  continue  over 
a  period  of  time  and  which  are  designed  to  accomplish  a  broad  scientific  or 
technical  goal  in  the  NASA  Long- Range  Plan."52  Typically,  the  time  span 
of  a  NASA  project  was  two  to  three  years.  Two  examples  of  the  agency's 
"broad  scientific  and  technical  goals"  from  the  early  1960s  were  manned 
spaceflight  (spearheaded  by  Project  Mercury)  and  the  exploration  of  the 
moon  and  the  planets  (supported  early  on  by  the  Ranger  and  Surveyor 
projects).  After  President  Kennedy's  speech  in  May  1961,  NASA's  most 
important  goal  became  a  manned  lunar  landing  that  was  achievable  by  the 
end  of  the  decade.  That  goal  was  so  primary  that  Apollo,  the  project, 
quickly  became  Apollo,  the  program.  It  so  dominated  NASA's  efforts  that 
the  moon  landing  became  virtually  coextensive  with  the  mission  of  the  entire 

CO 

space  agency.00* 

In  contrast  to  projects  with  their  definite  beginnings  and  ends  and  specific 
goals,  research  is  by  nature  more  open-ended  and  unpredictable.  To  obtain 
significant  results  from  research,  even  from  the  more  practical  engineering 

113 


Space/light  Revolution 

kind  carried  out  at  Langley  during  its  NACA  period,  risks  must  be  taken. 
Researchers  must  venture  down  long  and  winding  roads  .that  might  lead 
nowhere,  ask  questions  that  might  turn  out  to  be  unanswerable,  and  spend 
money  on  experimental  equipment  to  conduct  demonstrations  that  might 
never  work. 

In  other  words,  the  environment  for  research  has  to  be  flexible.  Needless 
to  say,  so  too  does  the  researcher  and,  perhaps  especially  so,  the  research 
manager.  For  a  technical  culture  to  be  understanding  and  supportive  of 
research,  it  must  be  forgiving  of  failure  and  the  apparent  lack  of  progress. 
On  the  other  hand,  as  a  1979  NASA  study  of  the  R&D  process  declares, 
"Projects  often  provide  the  ultimate  reality.  [They]  are  practical  demonstra- 
tions. New  equipment  must  function  well,  performance  is  measured  against 
the  previous  experience,  and  success  needs  to  be  achieved."54  Otherwise, 
the  project  is  a  total  failure.  The  situation  is  rather  black-and-white. 

In  research,  the  criteria  for  success  and  failure  are  gray;  success  needs  to 
be  achieved  only  once  in  a  while.  One  fundamental  breakthrough  that  can 
be  built  upon  for  many  years  makes  up  for  dozens  of  wrong  turns  and  dead 
ends.  A  breakthrough  may  even  be  accidental  or  the  fortuitous  consequence 
of  some  meandering.  This  is  rarely  the  case  in  a  project.  When  success 
is  a  necessity  and  the  timetable  is  short,  nothing  can  be  left  to  accident 
or  luck;  a  "fail-safe"  system  is  called  for.  Constructing  such  a  system 
requires  systematic  and  detailed  planning,  rigorous  discipline,  proof-tested 
technology,  and  extremely  prudent  management  and  overall  leadership — not 
to  mention  enough  talented  and  motivated  people  to  work  all  the  overtime 
required  to  complete  the  job  on  schedule. 

During  its  41-year-long  history  as  an  NACA  laboratory,  Langley's  "ulti- 
mate reality"  had  been  firmly  rooted  in  research,  not  in  projects.  Generally 
speaking,  Langley  valued  research  more  than  anything  else.  The  most  mer- 
itorious thing  that  a  Langley  scientist  or  engineer  could  do  was  to  write 
an  outstanding  research  paper  that  the  NACA  would  publish  as  a  formal 
technical  report.  Langley  researchers  did  not  design  or  build  airplanes;  as 
government  employees,  they  were  not  supposed  to,  or  allowed  to,  do  that. 
What  they  did  was  the  basic  testing  that  generated  the  fundamental  knowl- 
edge that  the  aircraft  industry  used  to  advance  the  state  of  the  nation's 
aeronautical  art. 

The  NACA  laboratory  was,  therefore,  not  a  place  for  pure  research;  it 
was  a  place  for  applied  basic  research  and  for  technology  development.  As 
such,  Langley  staff  understood  and  placed  great  importance  on  project  work. 
Most  NACA  research  was  neither  "basic"  nor  "scientific"  in  the  usual  sense 
of  those  words;  almost  every  investigation  at  the  center,  whether  "funda- 
mental" or  "developmental,"  was  aimed  at  a  useful  aircraft  application. 
What  Langley  researchers  did  best  was  attack  the  most  pressing  problems 
obstructing  the  immediate  progress  of  American  aviation,  particularly  those 
vexing  the  military  air  services,  and  aircraft  manufacturing  and  operating 
industries.  This  had  often  meant  "fighting  fires,"  bringing  diversified  R&D 

114 


Change  and  Continuity 

talents  to  bear  on  a  problem  of  the  moment,  and  eliminating  or  solving  that 
problem  in  as  short  a  time  as  possible.  Doing  so  was  virtually  like  carrying 
out  a  project. 

Thus,  in  the  NACA's  way  of  doing  research,  of  developing  wind  tunnels 
and  other  test  facilities,  and  of  attacking  technical  problems,  Langley 
researchers  often  followed  an  approach  akin  to  project  management.  Many 
people  at  NACA  Langley  felt  that  their  best  research  programs  were  those 
run  as  projects.  For  instance,  the  approach  the  center  adopted  to  building 
many  major  new  facilities  had  been  very  much  like  project  management. 
Frequently  during  meetings  of  employee  promotion  boards  in  the  1950s, 
a  member  of  the  senior  staff  would  ask  whether  the  candidate  was  a 
"project  engineer"  or  simply  a  "researcher."  By  project  engineer,  they 
meant  someone  who  could  take  on  all  the  responsibilities  for  carrying  out 
a  task  and  meeting  a  deadline.  To  do  this,  the  project  engineer  had  to 
deal  with  wind-tunnel  operators,  get  work  done  in  the  shops,  consult  with 
systems  engineering  and  other  technical  support  people,  and  perhaps  even 
do  a  little  bit  of  procurement,  such  as  arranging  for  the  purchase  of  supplies, 
materials,  or  some  minor  piece  of  equipment. 

This  kind  of  management  was  done  on  a  much  smaller  scale  than  would 
be  done  for  a  NASA  project,  but  NACA  Langley  researchers  did  have 
comparable  experiences.  With  the  coming  of  NASA,  they  only  had  to  learn 
to  do  it  on  a  larger  scale.  From  the  end  of  World  War  II,  PARD  had 
been  involved  with  rocket  acquisitions  and  launch  operations,  and  starting 
in  the  mid-1950s,  Langley  was  also  heavily  involved  in  the  large  Project 
WS-110A.  (The  designation  "WS"  stood  for  "Weapons  System.")  This 
was  a  top  secret  air  force  project  for  the  development  of  what  became 
the  North  American  XB-70,  an  experimental,  six-engine,  520,000-pound 
strategic  bomber  designed  for  a  speed  in  excess  of  Mach  3.  (Only  two  were 
built  before  the  project  was  canceled  in  1964. )55 

Experiences  such  as  those  in  PARD  and  with  WS-110A  made  the  man- 
agement of  a  project  easier  for  Langley  when  the  time  came.  Most  people 
who  would  be  assigned  to  many  of  the  earliest  NASA  projects  at  Langley 
would  come  from  PARD.  Although  Langley  staff  moved  into  the  project 
work  brought  on  by  the  spaceflight  revolution  and  the  changeover  to  NASA 
without  too  much  difficulty,  the  novelty  or  the  essential  differences  between 
conducting  project  work  and  doing  research  should  not  be  underestimated. 

PARD  had  more  critics  within  Langley  than  did  any  of  the  laboratory's 
other  research  divisions.  From  the  moment  of  PARD's  establishment  as  a 
separate  division  in  1946  through  its  reincarnation  as  the  Applied  Materials 
and  Physics  Division  in  1959,  researchers  in  other  divisions  were  always 
bickering  with  someone  in  PARD.  Wind-tunnel  groups  questioned  the  merits 
of  PARD's  wing-flow  and  rocket-model  transonic  testing  techniques,  arguing 
that  they  were  too  costly  and  often  took  priority  over  more  basic  tunnel 
programs.  Each  firing  of  a  PARD  rocket  model  from  Wallops  Island  required 
that  a  precious  test  model  be  sacrificed;  often  the  models  had  expensive 

115 


Space/light  Revolution 

instruments  inside.  Among  others,  John  V.  Becker,  the  influential  head 
of  the  Compressibility  Research  Division,  complained  about  the  "voracious 
appetite"  of  the  rocket-model  advocates,  suggesting  that  many  engineers  in 
PARD  were  more  interested  in  making  their  rocket  models  perform  with 
increasing  accuracy  than  in  solving  research  problems.  Becker  warned  that 
the  practice  was  causing  "a  major  slowdown"  in  the  production  of  the 
models  and  instruments  required  by  his  division  and  by  others.  In  his 
judgment,  what  PARD  was  expecting,  and  often  receiving,  from  Langley's 
model  shops  and  other  technical  support  services  was  "roughly  equivalent 
to  the  requirements  of  perhaps  10  major  wind  tunnels."56 

Although  much  of  the  criticism  was  unfair,  these  feelings  about  PARD 
and  about  its  focused,  rather  aggressive  project-like  approach  to  doing 
things  worried  many  senior  staff  members  of  the  1960s.  Becker  and  others 
thought  that  most  of  the  personnel  in  PARD  were  "blacksmiths,"  hairy- 
armed,  technical  musclemen  who  did  things  hit  or  miss,  with  hammer  and 
tongs,  and  without  much  serious  forethought.  One  of  Becker's  branch  heads, 
Macon  C.  Ellis,  Jr.,  remembers  that  feelings  against  PARD  within  the  Gas 
Dynamics  Laboratory  were  so  strong  that  "when  we  became  MPD  [the 
Magnetoplasmadynamics  Branch,  in  1960],  we  definitely  didn't  want  to  go 
into  PARD.  That  was  for  sure."57 

As  Langley  took  on  more  project  work  during  the  1960s,  people 
strictly  involved  in  research  grew  increasingly  upset.  Larry  Loftin,  Floyd 
Thompson's  technical  assistant  and  later  director  of  Group  3,  remembers 
with  some  hard  feelings  that  "anything  with  the  name  'project'  got  first 
priority  in  the  shops."  Again,  this  perturbed  those  research  groups  involved 
in  wind-tunnel  testing.  "You  couldn't  do  wind-tunnel  tests  without  mod- 
els," Loftin  recalls,  "and  you  couldn't  get  your  models  done  without  the 
shops.  All  a  person  had  to  do  was  mention  Mercury  or  some  other  project 
to  somebody  in  the  shops,  and  it  got  done.  Everybody  else  waited  their 
turn."  Hostility  was  particularly  high  regarding  Project  WS-110A.  Any 
work  connected  to  WS-110A  received  the  highest  priority  at  Langley.  Any 
test  model  needed  for  the  project  immediately  was  built  in  the  shops,  then 
was  pushed  to  the  front  of  the  line  for  wind-tunnel  testing.  This  situation 
led  a  frustrated  researcher  to  try  connecting  one  of  his  job  orders  to  Project 
WS-110A  so  that  he  could  get  some  of  his  own  work  done.58 

In  analyzing  the  impact  of  NASA  project  work  on  the  traditional 
character  of  Langley,  continuity  from  the  NACA  period  must  not  be 
exaggerated.  Researchers  like  Becker  and  Ellis  drew  a  line  dividing  the 
ways  of  NASA  projects  from  NACA  research  and  continued  to  draw  it  well 
into  the  NASA  years.  John  Stack,  the  billy-goat-gruff  of  the  Langley  senior 
staff,  never  abandoned  the  research  ideal  of  the  NACA.  In  his  opinion,  the 
most  valuable  thing  that  any  Langley  employee  ultimately  could  contribute 
was  a  published  research  paper  that  the  American  aerospace  community 
could  use.  Without  such  contributions,  a  laboratory  would  amount  to  no 
more  than  an  industrial  plant.59 

116 


Change  and  Continuity 


L-57-5231 

In  this  1957  photo,  aerodynamicists  prepare  a  scale  model  of  the  top  secret  WS-110A 
for  testing  in  Langley's  7  x  10-Foot  High-Speed  Tunnel. 


Uncharted  Territory 

No  matter  what  PARD  had  done  that  was  akin  to  project  work  during  the 
NACA  period,  large-scale  projects  for  spaceflight  were  totally  new.  Langley 
was  inexperienced  in  many  details  of  project  management,  in  procurement, 
and  in  matters  concerning  the  administration  of  the  space  agency's  expanded 
R&D  and  mission  activities. 

In  putting  together  its  diversified  operation,  NASA  faced  a  complex  task: 
it  had  to  build  an  effective  organizational  structure  involving  intraagency 
relationships;  it  had  to  devise  a  rational  complex  of  administrative  proce- 
dures that  took  care  of  both  internal  and  external  matters;  and  it  had  to 
find  the  best  ways  to  procure  supplies  and  services.  This  last  requirement, 
procurement  administration,  was  especially  problematic  for  a  technical  or- 
ganization like  Langley  because  it  involved  the  writing,  negotiating,  and 
managing  of  contracts.  This  meant  extensive  dealings,  legal  and  otherwise, 
with  corporations  and  industrial  firms  in  the  profit-motivated  private  sector 
of  the  American  economy.  Such  a  complicated  affair  had  never  been  the 
case  for  NACA  research. 

In  the  early  days  of  the  space  agency,  NASA  headquarters  realized  that 
most  of  its  executive  personnel,  especially  those  running  the  field  centers, 
were  "excellent  technical  people"  who  "lacked  experience"  in  managing  large 

117 


Space/light  Revolution 

projects.  Two  outside  studies  sponsored  by  NASA  in  mid-1960,  one  by  an 
advisory  committee  on  NASA  organization  chaired  by  University  of  Chicago 
President  Lawrence  Kimpton  and  the  other  done  under  contract  by  the 
Washington  management  consulting  firm  of  McKinsey  &  Co.,  found  that 
NASA's  executive  class  needed  beefing  up.  With  Administrator  Glennan 
enthusiastically  in  support  of  this  finding,  NASA  immediately  began  a 
formal  program  to  train  project  managers.  It  hired  a  contractor,  Harbridge 
House,  to  develop  and  lead  a  series  of  two- week  training  courses  in  project 
management.  The  first  of  these  courses  convened  in  Williamsburg,  Virginia, 
not  more  than  25  miles  from  Langley,  in  December  1960.  Employees 
from  all  the  NASA  installations  attended.  Langley  sent  several  people — 
not  all  of  them  picked  for  their  potential  as  project  managers.  Some 
general  administrative  staff  also  attended  the  seminars,  as  did  a  handful 
of  senior  managers  like  Larry  Loftin  and  Gene  Draley.  Top  NASA  officials 
and  managers  of  industry  addressed  the  participants,  while  specialists 
from  Harbridge  House  took  groups  through  case  studies  "from  actual,  but 
camouflaged,  R&D  problems"  faced  by  NASA  and  the  DOD.  Essentially, 
what  everyone  was  supposed  to  glean  from  the  training,  and  for  the  most 
part  did,  was  a  heightened  concern  for  certain  basic  management  principles 
and  theories.60 

What  NASA  hoped  to  achieve  through  this  training  course  was  "a 
measure  of  uniformity"  in  the  management  of  its  diverse  projects  agen- 
cywide.  NASA  did  not  want  more  centralized  control  over  the  projects; 
this  had  already  been  tried  to  some  extent  in  the  first  two  years  of  NASA's 
operation  and  had  resulted  in  an  impossibly  heavy  work  load  at  NASA 
headquarters.61  NASA  wanted  to  move  toward  a  more  decentralized  system 
in  which  one  field  installation  would  have  virtually  complete  management 
control  over  the  execution  of  an  entire  project;  the  need  for  interinstallation 
coordination  would  be  at  a  minimum;  and  NASA  headquarters  could  stay 
out  of  the  intraproject  coordination  and  instead  could  concentrate  on  inter- 
project  coordination,  which  included  "the  review  and  approval  of  projects 
in  the  light  of  overall  objectives,  schedules,  and  costs  of  the  entire  agency." 
All  three  points  were  underscored  in  the  October  1960  final  report  of  the 
McKinsey  &  Co.  study  of  the  NASA  organization.  In  fact,  the  firm's  advo- 
cacy of  a  training  course  in  project  management  stemmed  directly  from  the 
conclusions  of  its  specialists  about  the  advantages  of  a  decentralized  system. 
Such  a  system  could  work,  the  report  said  in  emphatic  terms,  only  if  each 
NASA  center  trained  10  or  20  people  in  this  kind  of  management.62 

NASA  would  need  three  years  to  create  the  decentralized  system  called 
for  in  the  McKinsey  report.  With  the  NASA  reorganization  of  October  1963 
asked  for  by  Administrator  Webb,  the  system  finally  was  firmly  put  into 
place.  From  that  point  on,  as  Arnold  S.  Levine  explains  in  his  1982  analysis, 
Managing  NASA  in  the  Apollo  Era,  NASA  leadership  stressed  that  "project 
management  was  the  responsibility  of  the  centers."  For  all  flight  projects, 
"there  was  to  be  one  lead  center,  regardless  of  how  many  installations 

118 


Change  and  Continuity 

actually  participated."*  To  take  the  lead,  "a  particular  center  had  to 
[have]  (or  was  assumed  to  have)  the  capacity  to  manage  large  development 
contracts,  the  skills  to  integrate  the  subsystems  of  a  project  parceled  out 
among  two  or  three  different  centers,  and  the  ability  to  draw  on  the  resources 
of  the  centers  instead  of  needlessly  duplicating  them."63  Those  in  charge  of 
a  project  at  a  lead  center  would  report  their  business,  in  a  direct  and  official 
line  of  communication,  to  the  head  of  the  appropriate  program  office  at 
NASA  headquarters,  for  example,  to  the  head  of  the  OART.  Senior  staff  in 
these  program  offices  then  supervised  and  counseled  the  work  of  the  project 
managers  in  the  field  as  they  saw  fit.64 

Ironically,  where  this  shift  in  NASA  project  management  policy  seems 
to  have  led  by  1963  was  back  to  the  NACA  concept  of  giving  the  field 
centers  the  responsibility  for  technical  decisions.  Of  course,  the  overall 
political  and  cultural  context  in  which  those  decisions  were  made  was 
far  different  from  the  one  in  which  Langley  had  operated  as  an  NACA 
aeronautics  laboratory.  The  NACA  was  not  involved  with  contractors  and 
all  the  snarly  legalities  and  procedures  that  necessarily  came  with  them.  In 
the  narrower  context  of  the  NACA,  technical  decisions  were  not  nearly  as 
visible  or  important  to  the  American  public  as  they  would  be  in  the  high- 
profile  space  program.  If  an  NACA  decision  had  been  wrong,  the  result 
might  have  been  tragic — if,  for  example,  the  aircraft  industry  or  military  air 
services  had  applied  a  mistaken  NACA  research  finding  in  a  new  airplane 
design.  But  the  overall  context  for  NACA  research  was  such  that  major 
mistakes  were  almost  impossible  to  make.  In  normal  periods,  researchers 
could  usually  take  all  the  time  necessary  to  be  scrupulously  careful  and 
certain  of  their  findings.  Even  during  the  rush  to  support  the  Allied  air 
forces  in  World  War  II,  which  involved  "cleanup"  of  existing  aircraft  designs 
as  well  as  fundamental  research  and  development,  researchers  had  time  to  be 
systematic.65  Furthermore,  the  NACA's  clients  never  applied  aerodynamic 
test  results  indiscriminately.  All  sorts  of  institutional  checks  and  balances 
would  be  exercised  to  confirm  the  veracity  of  the  government's  research  data 
before  using  it.  In  comparison,  the  context  for  NASA  projects  involved  a 
much  higher  degree  of  institutional  risk.  As  we  have  already  noted  about 
projects,  "success  needs  to  be  achieved"  and  in  a  limited  amount  of  time. 
The  successes  of  the  space  race  projects  would  eventually  cost  NASA  and 
Langley  in  ways  their  researchers  could  not  have  calculated  in  the  early 
1960s.  In  research,  success  had  always  been  broadly  defined  and  its  price 
not  so  dear,  but  Langley  would  learn  quickly  just  how  exacting  a  space 
project  could  be. 


i 

This  was  not  true  for  Apollo,  which  was  so  big  and  so  important  that  NASA  divvied  up  the  work 

among  lead  centers:  the  spacecraft  development  to  Houston,  the  launch  vehicle  development  to  Marshall, 
and  the  tracking  system  to  Goddard. 

119 


The  "Mad  Scientists"  of  MPD 


What  about  this  plasma  physics  ?  Will  it  ever  amount 
to  anything? 

-Dr.  Hugh  L.  Dryden,  NASA 
deputy  administrator,  to 
Macon  C.  Ellis,  Jr.,  head  of 
Langley's  Magnetoplasmadynamics  Branch 


While  the  Hydrodynamics  Division  sank  at  Langley,  a  few  new  research 
fields  bobbed  to  the  surface  to  become  potent  forces  in  the  intellectual 
life  of  the  laboratory.  Most  notable  of  these  was  magnetoplasmadynamics 
(MPD) — a  genuine  product  of  the  space  age  and  an  esoteric  field  of  scientific 
research  for  an  engineering-  and  applications-oriented  place  like  Langley. 
If  any  "mad  scientists"  were  working  at  Langley  in  the  1960s,  they  were 
the  plasma  physicists,  nuclear  fusion  enthusiasts,  and  space-phenomena 
researchers  found  in  the  intense  and,  for  a  while,  rather  glamourous  little 
group  investigating  MPD.  No  group  of  researchers  in  NASA  moved  farther 
away  from  classical  aerodynamics  or  from  the  NACA's  traditional  focus  on 
the  problems  of  airplanes  winging  their  way  through  the  clouds  than  those 
involved  with  MPD. 


The  ABCs  of  MPD 

The  field  of  MPD  concerned  the  effects  of  magnetic  and  electric  fields  on 
the  motions  of  plasmas.  A  plasma,  as  simply  defined  at  the  time,  consists 
of  an  ionized  high-temperature  gas.  For  those  readers  who  have  forgotten 
their  high  school  chemistry,  a  gas  consists  of  atoms  and  molecules  that  are 
virtually  unrestricted  by  intermolecular  forces,  thus  allowing  the  molecules 
to  occupy  any  space  within  an  enclosure.  In  other  words,  the  atoms  and 

121 


Spaceflight  Revolution 

molecules  are  continually  moving  around  and  colliding  with  one  another. 
When  a  sufficiently  violent  collision  between  two  atoms  occurs,  a  negatively 
charged  subatomic  particle  known  as  an  electron  is  knocked  out  of  its  orbit, 
thus  resulting  in  a  "free  electron"  (an  electron  that  is  not  bound  to  an  atom). 
Sometimes  in  the  collision,  an  ion  (a  positively  charged  particle  bound  to  the 
electron)  is  knocked  free  as  well.  At  the  instant  these  particles  are  released, 
the  gas  is  said  to  be  "ionized"  and  is  called  a  plasma. 

Considered  as  a  whole,  a  plasma  is  electrically  neutral,  composed  as  it 
is  of  an  approximately  equal  number  of  positively  and  negatively  charged 
particles  plus  a  variable  fraction  of  neutral  atoms.  A  plasma,  however,  by 
virtue  of  its  charged  particles,  is  nonetheless  a  conductor  of  electricity.  Thus, 
as  is  true  for  any  electrical  conductor,  the  motion  of  a  plasma  can  be  greatly 
influenced,  and  perhaps  even  controlled,  by  electromagnetic  forces.1 

By  the  late  1940s,  the  study  of  the  motion  of  ionized  gases  in  the 
presence  of  magnetic  fields  had  become  a  major  international  focus  for 
scientific  research.  The  new  field,  which  was  really  a  subfield  of  the 
large,  complicated,  and  still  emerging  discipline  of  "plasma  physics," 
was  known  by  many  names:  "magnetohydrodynamics,"  "hydromagnet- 
ics,"  "magneto-aerodynamics,"  "magnetogasdynamics,"  and  "fluid  electro- 
dynamics." *  Generally  speaking,  however,  the  name  "magnetohydrodynam- 
ics," or  MHD,  won  out.2 

But  the  name  did  not  prevail  at  NACA  Langley.  There,  in  the  years 
before  the  establishment  of  NASA,  a  coterie  of  aerodynamic  researchers 
involved  in  plasma  studies  conducted  in  the  center's  Gas  Dynamics  Labo- 
ratory, thought  that  the  name  magneto /lydrodynamics  was  not  appropriate. 
The  interested  researchers  were  not  concerned  with  water  but  rather  with 
hot  gases  or  plasmas,  so  they  coined  the  term  "magnetop/asmodynamics." 
Outside  of  NASA,  however,  magnetohydrodynamics  remained  the  standard 
term. 


The  Solar  Wind  Hits  Home 

Most  work  on  plasmas  before  World  War  II  pertained  to  the  dynamics 
of  upper  atmosphere  magnetic  storms  and  to  the  phenomenon  of  radiant 
auroral  displays  similar  to  the  aurora  borealis  or  "northern  lights."  These 
studies,  undertaken  most  notably  by  a  British  group  interested  in  solar 
and  terrestrial  relationships  led  by  astrophysicist  Sydney  Chapman  (1888- 
1970),  involved  questions  about  what  fueled  the  sun  and  the  stars  and  about 
how  the  ionized  gases  brought  about  by  ultraviolet  radiation  behaved  in 

>k 

Preference  for  one  name  over  the  others  depended  on  whether  the  scientists  involved  felt  that  the 
electrically  active  medium  that  they  were  studying  should  properly  be  regarded  as  a  continuum  or, 
more  accurately,  as  comprising  discrete  individual  particles.  The  astrophysicists  preferred  the  name 
"hydromagnetics" ;  the  aerodynamicists  opted  for  "magneto-aerodynamics."  ' 

122 


The  "Mad  Scientists"  of  MPD 

interstellar  space.  In  the  1920s,  Chapman  postulated  that  several  geocosmic 
phenomena  could  be  explained  by  the  "differential  action"  of  the  earth's 
magnetic  field  on  protons  and  electrons  emanating  from  the  sun.  Solar 
activity,  in  Chapman's  soon-to-be  dominant  view,  influenced  the  terrestrial 
magnetic  field,  aurorae,  the  conduct  of  atmospheric  electricity,  and  the 
earth's  weather  patterns.3 

In  1942,  Swedish  astrophysicist  Hannes  Alfven  (an  eventual  winner  of 
the  Nobel  Prize)  advanced  an  MHD  theory  of  the  so-called  solar  cycle, 
the  periodic  round  of  disturbances  in  the  sun's  behavior  as  seen  in  the 
fluctuation  in  the  number  and  the  area  of  sunspots  and  in  the  form  and 
shape  of  the  sun's  corona.  Some  10  years  later,  in  the  early  1950s,  Alfven 
proposed  an  even  more  provocative  theory.  He  postulated  that  the  planets 
had  been  formed  by  an  MHD  process  by  which  ionized  gases  became  trapped 
electromagnetically  and  pulled  inward  by  the  sun's  gravitational  force,  thus 
leaving  them  at  certain  distances  from  the  sun.  The  only  way  to  fathom  the 
process,  Alfven  argued,  was  to  work  further  with  MHD  equations.4 

Thus,  in  large  measure,  the  interest  in  MHD  began  with  the  modern 
astrophysicists.  From  the  1920s  on,  many  of  their  most  essential  questions 
concerned  MHD:  What  mechanisms  are  involved  in  galaxy  formation? 
What  is  the  nature  of  the  magnetic  fields  of  the  sun  and  the  other  stars? 
How  does  the  internal  energy  in  hot  stars  convert  into  the  kinetic  energy  of 
gaseous  clouds  in  interstellar  space?  How  do  stars  form  from  gas  clouds? 
What  is  the  origin  of  cosmic  rays,  the  Solar  System,  the  universe?  The  key 
to  understanding  the  cosmos  lay  in  the  fathoming  of  MHD  principles. 

Revolutionary  discoveries  about  the  space  environment  made  with  the 
first  space  probes  strengthened  the  belief  in  MHD's  importance.  On  1  May 
1958,  five  months  to  the  day  before  the  NACA  transition  to  NASA,  Amer- 
ican astrophysicist  James  Van  Allen  announced  his  discovery  of  a  region  of 
intense  radiation  surrounding  the  earth  at  high  altitude.  Data  from  Geiger 
counters  aboard  the  first  three  Explorer  spacecraft,  the  first  successful  Amer- 
ican satellites,  confirmed  a  theory  that  Van  Allen  had  been  working  on  for 
some  time.  This  theory  suggested  that  the  earth's  magnetic  field  trapped 
charged  subatomic  particles  within  certain  regions.  Experiments  aboard 
subsequent  exploratory  rockets  and  spacecraft  indicated  with  a  high  degree 
of  certainty  that  more  than  one  radiation  belt  in  fact  enveloped  the  earth. 
The  intensity  of  the  belts  varied  with  their  distance  from  the  earth.  The 
zone  of  the  most  intense  radiation  began  at  an  altitude  of  approximately 
1000  kilometers  (621.37  miles).5 

The  discovery  of  what  immediately  came  to  be  known  as  the  Van  Allen 
radiation  belts  inspired  a  wide  range  of  fundamental  new  investigations. 
Within  months,  scientists  around  the  world  realized  that  surrounding  the 
earth  was  a  vitally  important  magnetic  region  of  still  unknown  character, 
shape,  and  dimension  where  ionized  gases — plasmas — exerted  a  strong  force. 
They  dubbed  this  mysterious  region  "the  magnetosphere."  In  the  exciting 
but  highly  speculative  early  days  of  magnetospheric  physics,  this  region  was 

123 


Spaceflight  Revolution 


L-61-6169 


WE  EXPLORE  THE  PLANET  EARTH 


INNER  RADIATION  BUT 


CHARGED  PARTICLES 
FRO»  SOU 


MAONiTiC  FIELD 


IONOSPHERE 


ATMOSPHERE 


L-61-6163 

Langley's  MPD  researchers  used  these  schematic  drawings  to  illustrate  the  main 
features  of  earth 's  bordering  region  with  outer  space. 


124 


The  "Mad  Scientists"  of  MPD 

alternately  described  as  "a  high  region  of  the  earth's  atmosphere"  or  as  a 
"low  or  bordering  region  of  space. 

Another  important  discovery  of  the  space  age  fed  the  new  science  of 
magnetospheric  physics:  the  notion  of  "the  solar  wind."  This  theory  was 
first  expressed  by  Eugene  N.  Parker  of  the  University  of  Chicago  in  1958 
and  later  confirmed  by  measurements  taken  from  Soviet  Lunik  spacecraft 
in  1959-1960  and  from  Explorer  10  in  1961.  Parker  suggested  that  the 
sun's  corona,  or  outer  visible  envelope,  was  expanding  continuously,  causing 
streams  of  ionized  gases  to  flow  radially  outward  from  the  sun  through 
interplanetary  space.  (The  sun  is,  after  all,  a  big  ball  of  plasma.)  The 
intensity  of  these  plasma  streams  varied  greatly  relative  to  solar  activity, 
especially  solar  flares.  The  force  of  these  streams,  or  solar  wind,  impinging 
upon  the  earth's  magnetic  field  created  the  familiar  magnetic  storms.7  By 
1960  scientists  possessed  evidence  that  a  plasma  wind  did  blow  continuously 
from  the  sun,  and  the  wind  clearly  displayed  dynamic  magnetic  phenomena. 

The  field  of  study  that  the  Langley  researchers  had  come  to  call  MPD 
was  growing  quickly  in  esteem  and  importance,  not  only  in  the  United  States 
but  also  around  the  world.  Newly  conceived  experiments  with  magnetically 
compressed  plasmas  provided  scientists  with  an  opportunity  to  generate 
and  study  a  small  sample  of  the  solar  corona  in  the  laboratory.  Scientists 
gathered  basic  data  on  subatomic  behavior  at  temperatures  for  which  no 
such  information  existed  before.  A  major  and  extraordinarily  exciting  new 
age  of  modern  physics  was  dawning.  Scientists  saw  fascinating  new  research 
opportunities,  and  they  dreamed  of  fantastic  technological  applications. 
Unfortunately,  very  few  of  their  dreams  would  be  realized.  But  in  the  early 
1960s,  that  was  something  impossible  to  know. 

What  Langley  researchers,  especially  those  involved  in  gas  dynamics  and 
other  hypersonic  investigations,  did  know  in  the  late  1950s  was  that  the 
time  for  a  major  change  had  arrived.  "The  space  age  told  us  to  move  away 
[from]  classical  aerodynamics  into  more  modern  things,"  remembers  Macon 
C.  "Mike"  Ellis,  the  man  who  would  head  Langley's  formal  MPD  effort, 
"and,  as  quickly  as  we  could,  we  did."8  In  handwritten  notes  made  at 
an  internal  meeting  of  his  Gas  Dynamics  Branch  held  on  18  June  1958— 
during  the  same  period  that  plans  for  NASA's  initial  organization  were 
being  formulated  in  Washington — Ellis  wrote,  "Either  we  make  a  big  change 
now  or  [we]  try  to  make  more  significant  contributions  in  aerodynamics." 
MPD  is 

a  field  we  are  already  in  and  should  push  hard.  .  .  .  We  should  go  all  out  to  get 
qualified  physics  instructors.  .  .  .  We  should  have  seminars  on  "space-type"  and 
reentry  subjects.  .  .  .  We  must  work  and  plan  toward  ultimate  "conversion"  of  our 
work  when  aerodynamics  becomes  secondary.  .  .  .  We  must  go  big  into  the  new 
environment  of  space. 


125 


Spaceflight  Revolution 

Now  was  the  time  for  Langley  researchers  to  assume  leadership  roles  in 
the  emerging  space  disciplines  and  vigorously  seek  major  technological 
applications. 


The  MPD  Branch 

Through  the  late  1950s,  nothing  had  been  done  formally  at  Langley  to 
focus  the  efforts  of  those  involved  in  the  study  of  MPD-related  subjects. 
Many  people  at  the  laboratory,  some  of  them  senior  engineers  and  research 
managers,  did  not  know  what  MPD  was  or  did  not  understand  what  all 
the  fuss  was  about.  Furthermore,  nearly  all  of  the  people  concerned  with 
MPD  were  members  of  the  Gas  Dynamics  Laboratory,  so  they  were  already 
grouped  together  and  interacting  regularly.  Thus,  for  several  months,  even 
after  the  new  space  agency  was  established,  no  Langley  leaders  saw  a  need 
to  create  a  new  organization  just  for  the  MPD  enthusiasts. 

But  interest  in  the  new  field  kept  growing.  The  idea  that  flows  could 
get  so  hot  that  the  constituents  of  the  air  would  actually  break  down  and 
become  treatable  by  applying  magnetic  forces  was  extremely  exciting.  If  air- 
flows could  be  "treated"  electromagnetically,  they  might  even  be  controlled. 
That  was  every  aerodynamicist's  dream.  MPD  offered  a  sort  of  aerodynamic 
alchemy,  a  magical  way  of  turning  lead  into  gold,  rough  turbulent  flow  into 
smooth  laminar  flow,  dangerous  reentry  conditions  into  pacific  ones.  With 
these  glorious  possibilities,  MPD  fostered  great  technological  enthusiasm 
and  attracted  many  able  researchers  who  hoped  to  find  solutions  to  some 
fascinating  and  very  complex  problems. 

The  study  of  MPD  became  increasingly  glamourous  in  the  late  1950s,  so 
much  so  that  Langley  management  soon  understood  that  it  should  advertise 
the  progress  that  Langley  researchers  were  making  in  MPD  studies.  At  each 
of  the  former  NACA  laboratories — Lewis,  Ames,  and  Langley — research  in 
MPD  grew  in  earnest  in  the  months  just  before  the  metamorphosis  of  the 
NACA  into  NASA  and  thereafter  gained  momentum.10  At  the  first  NASA 
inspection  in  October  1959,  MPD  was  a  featured  attraction.  In  the  printed 
inspection  program,  MPD  merited  one  of  the  13  subtitled  sections.  Visitors 
on  the  inspection  tour  stopped  at  a  special  MPD  exhibit.  At  that  stop,  a 
Langley  MPD  specialist  stood  in  front  of  a  graphic  panorama  of  the  universe 
and  introduced  his  subject  by  saying  that  "the  space  environment  is  filled 
with  manifestations  of  this  new  science."11 

Above  all  other  members  of  Langley's  staff,  Floyd  Thompson,  still 
officially  the  associate  director,  became  most  enthralled  with  the  glamour  of 
MPD.  As  Mike  Ellis  remembers,  "Thompson  was  tremendously  supportive 
of  our  effort."  One  of  the  best  measures  of  Thompson's  enthusiasm  was 
his  request  that  the  MPD  staff  be  "on  tap"  as  the  special  attraction 
for  major  events.  He  "always  put  us  on  stage  at  the  NASA  inspections 
and  when  various  groups  of  scientists  came  through  the  laboratory,"  Ellis 

126 


The  "Mad  Scientists"  of  MPD 

recalls.  Thompson  appreciated  that  work  in  this  exciting  new  field  of  science 
could  enhance  the  reputation  of  his  aeronautics  laboratory.12  In  May  1960, 
the  same  month  he  took  over  officially  from  Henry  Reid  as  the  Langley 
director,  Thompson  established  a  Magnetoplasmadynamics  Branch  of  the 
Aero-Physics  Division.  Prom  its  beginning,  MPD  was  one  of  Thompson's 
pet  projects. 

The  Aero-Physics  Division  was  the  natural  home  for  Langley's  MPD 
effort.  This  division  was  led  by  hypersonics  specialist  John  V.  Becker,  an 
NAG  A  veteran  whose  employment  at  Langley  dated  back  to  1936  and  who 
by  the  mid-1950s  had  become  deeply  involved  in  work  related  to  hypersonic 
gliders  and  winged  reentry  vehicles.  A  research-minded  engineer,  Becker  was 
a  strong  and  confident  division  chief  (he  had  been  one  since  the  mid-1940s, 
passing  up  several  opportunities  to  move  up  to  posts  in  senior  management). 
He  was  comfortable  having  a  research  effort  as  esoteric  and  as  sophisticated 
as  MPD  based  in  his  division.  Scientifically,  he  was  quite  sharp  and  was  more 
than  capable  of  appreciating  the  complexities  of  this  new  field  of  research 
as  well  as  its  promise  for  making  major  contributions  to  the  space  program. 
Through  the  10-year  span  of  the  MPD  Branch  (1960-1970),  Becker  not  only 
tolerated  the  many  MPD  enthusiasts  in  his  division  but  also  almost  always 
supported  their  ideas. 

The  first  and  only  person  to  be  in  charge  of  Langley's  MPD  Branch 
was  Mike  Ellis,  an  NACA  veteran  who  was  42  years  old  when  the  branch 
was  organized.  Ellis  had  come  to  work  at  Langley  in  1939,  and  over  the 
course  of  his  career  at  the  laboratory,  he  had  been  involved  in  pioneering 
work  on  the  aerodynamics  of  jet  engines,  ramjets,  and  supersonic  inlets 
and  nozzles.  Fittingly,  Ellis  had  worked  for  Eastman  Jacobs  and  with 
Arthur  Kantrowitz  in  the  early  1940s,  and  he  had  heard  firsthand  accounts 
of  his  former  colleagues'  attempt  to  design  a  fusion  reactor  in  the  spring 
of  1938.  By  the  late  1950s,  Ellis  was  one  of  Langley's  most  outspoken 
believers  in  MPD's  promise  of  technological  benefits.  Ellis  encouraged  Floyd 
Thompson's  enthusiasm  for  MPD  and  persuaded  Langley's  senior  staff  of 
mostly  engineers  that  MPD  was  a  field  of  research  vital  to  the  future  of 
NASA.  When  the  time  came  to  pick  someone  to  head  the  new  branch,  Ellis 
was  unquestionably  the  person  for  the  job. 

Ellis  was  no  extraordinary  "scientific  brain."  As  an  aeronautical  engineer, 
his  talents  were  quite  respectable,  but  he  possessed  no  special  competency  in 
the  physics  of  fluids  beyond  his  experience  in  aerodynamics  or  gas  dynamics. 
He  was  always  the  first  to  admit  that  the  complexities  of  plasma  physics  and 
MPD  were  such  that  "there  was  no  way"  that  he  personally  could  conduct 
basic  MPD  research.  That  challenge  he  would  leave  to  minds  more  suited 
for  it.  But  Ellis  could  bring  the  MPD  researchers  together  as  a  unit,  serve  as 
their  strong  external  advocate,  shield  them  from  front-office  pressures,  and 
make  sure  that  they  received  the  support  they  needed  to  carry  out  their 
work.  "I  just  tried  to  keep  my  head  above  water,"  Ellis  explains,  "and  keep 


127 


Space/light  Revolution 


In  the  1960s,  John  V.  Becker 
(left)  headed  the  Aero-Physics  Di- 
vision, which  was  home  to  many 
of  the  center's  highest  speed,  and 
most  radical,  research  facilities. 
These  included  supersonic  and 
hypersonic  wind  tunnels,  arc-jets, 
and  shock  tubes  covering  a  speed 
range  from  Mach  1.5  to  Mach  20. 
Some  of  these  facilities,  such  as 
the  $6.5  million  Continuous- Flow 
Hypersonic  Tunnel  (below),  were 
the  forebearers  of  the  strange 
apparatuses  of  the  MPD  Branch. 


L-61-4064 


L-6 1-6268 


128 


The  "Mad  Scientists"  of  MPD 


Engineer  Macon  C.  "Mike"  Ellis  was 
an  early  believer  in  the  promise  of 
MPD. 

L-62-8043 


these  'mad  scientists'  from  going  off  on  too  many  tangents,  or  going  mad 
myself."13 

The  MPD  Branch  never  became  a  large  outfit.  By  the  end  of  1962, 
it  had  less  than  50  total  staff  members:  27  professionals,  10  mechanics, 
4  computers  (mathematicians  who  helped  to  process  and  plot  numerical 
data),  and  6  secretaries.  This  staff  was  divided  into  four  teams  or  sections. 
Plasma  Applications,  headed  by  Paul  W.  Huber,  was  the  largest  section, 
with  8  professionals.  Space  Physics,  led  by  British  physicist  David  Adamson, 
was  the  smallest  with  3.  Robert  Hess's  Plasma  Physics  Section  had  7 
professionals,  and  George  P.  Wood's  Magnetohydrodynamics  Section  had 
5.  These  sections  (and  their  section  heads)  remained  in  place  until  the 
dissolution  of  the  MPD  Branch  in  1970. 

In  addition  to  being  small,  MPD  was  self-contained.  Whereas  most  of 
the  research  done  in  the  center's  branches  regularly  spilled  over  into  other 
functioning  units,  most  MPD  work  was  done  within  the  MPD  Branch.  A 
small  amount  of  related  research  was  done  in  the  Flight  Research  Division 
and  Full-Scale  Research  Division;  however,  most  of  this  work  concerned 
the  development  of  microwave  and  spectroscopic  diagnostic  techniques.  All 
told,  the  MPD  work  conducted  outside  the  MPD  Branch  never  involved 
more  than  about  five  researchers. 

In  terms  of  organizational  genealogy,  the  MPD  Branch  grew  from  a  nar- 
row stem.  With  the  exception  of  Adamson,  and  a  trio  of  his  colleagues  from 
a  space  physics  group  in  the  Theoretical  Mechanics  Division,  all  the  original 
members  of  the  MPD  Branch  came  from  the  Gas  Dynamics  Laboratory. 
The  guru  of  MPD  studies  in  this  lab  was  Adolf  Busemann.  Throughout 

129 


Space/light  Revolution 

the  1950s,  Busemann  had  inspired  engineers  with  his  provocative  theories 
and  experimental  ideas.  At  Langley  on  22-23  September  1958,  the  Ger- 
man aerodynamicist  even  chaired  an  important  interlaboratory  meeting  on 
MHD.  Ninety-three  people  attended  the  meeting,  which  featured  6  speakers 
from  Ames,  4  from  Lewis,  and  11  from  Langley  and  was  organized  into  three 
sessions — plasma  acceleration,  arc-jets,  and  ion  beams.  Busemann  gave  a 
20-minute  talk  on  the  theory  of  alternating-current  (AC)  plasma  accelera- 
tion. This  two-day  scientific  meeting,  held  one  week  before  the  changeover 
to  NASA,  was  the  precursor  of  much  larger  conferences  on  MPD  sponsored 
by  NASA  on  almost  an  annual  basis  into  the  mid-1960s.14 

Among  the  scientists  working  in  MPD  at  Langley  were  several  Germans. 
Like  many  other  scientific  institutions  around  the  country,  Langley  had  re- 
ceived a  handful  of  German  scientists  who  were  part  of  Operation  Paperclip, 
the  U.S.  Army  intelligence  operation  that  brought  captured  German  rocket 
scientists  and  engineers  to  work  for  the  U.S.  government  at  the  end  of  World 
War  II.  Busemann  and  two  other  outstanding  researchers,  Karlheinz  Thorn 
and  Goetz  K.  H.  Oertel,  came  to  Langley  through  Paperclip.  Both  Thorn 
and  Oertel  moved  from  Gas  Dynamics  to  George  Wood's  MHD  Section  of 
the  new  MPD  Branch.  Both  men  stayed  at  Langley  for  several  years  before 
eventually  taking  posts  at  NASA  headquarters. 

At  least  10  German  scientists  came  to  Langley  as  part  of  a  postdoc- 
toral program  funded  by  NASA  but  sponsored  by  the  National  Academy  of 
Sciences.  This  program,  which  was  totally  divorced  from  the  normal  civil 
service  procurement  system,  enabled  NASA  to  obtain  talented  people  as 
Resident  Research  Associates  (RRAs)  without  going  through  the  normal 
hiring  procedures  of  the  civil  service  and  without  regard  for  NASA's  per- 
sonnel ceilings.  In  1968,  for  instance,  6  of  the  39  professionals  in  the  MPD 
Branch  were  RRAs.15 

Langley's  MPD  group  attracted  other  foreign  scientists.  These  included 
Dr.  Marc  Feix,  a  French  nuclear  scientist  who  spent  a  few  years  at  Langley 
in  the  mid-1960s  and  did  some  outstanding  theoretical  work.  Feix  was 
nominally  assigned  to  Hess's  Plasma  Physics  Section,  but  he  actually 
worked  with  various  people  throughout  the  branch,  especially  with  the 
Space  Physics  Section  under  David  Adamson.  Adamson  had  first  worked  at 
Langley  at  the  end  of  World  War  II  on  an  exchange  program  from  the  Royal 
Aircraft  Establishment  in  Farnborough  England.16  After  the  exchange, 
Adamson  went  home  to  England,  but  soon  returned  to  Langley.* 

In  the  1960s,  the  researchers  of  the  MPD  Branch  were  the  most  highly 
educated  group  of  people  at  Langley.  The  MPD  Branch  enjoyed  the 


In  1958,  in  support  of  Assistant  Director  Eugene  Draley's  initiative  to  advertise  Langley's  (then 
largely  alleged)  expertise  in  space  science,  Adamson  composed  an  excellent  paper  on  the  principles  of 
gravity.  According  to  some  experts,  this  paper,  which  NASA  published,  turned  out  to  be  "one  of  the 
best  papers  ever  written"  on  the  subject,  as  well  as  one  of  the  most  quoted.  (Ellis  audiotape,  Nov.  1991, 
author's  transcript,  p.  18.) 


130 


The  "Mad  Scientists"  of  MPD 


L-62-8122  L-65-1110 

Paul  W.  Huber  (left),  head  of  MPD 's  largest  section,  Plasma  Applications.  In  the 
mid-1960s,  French  nuclear  scientist  Dr.  Marc  Feix  (right)  floated  from  section  to 
section  within  the  MPD  Branch,  helping  researchers  solve  theoretical  problems  basic 
to  plasma  physics. 

academic  mystique  of  having  by  far  the  highest  percentage  of  advanced 
degree  holders.  At  one  point  MPD  had  eight  employees  with  earned 
doctorates,  seven  others  at  the  Ph.D.  dissertation  stage,  and  virtually  all 
of  its  younger  people  working  toward  advanced  degrees.* 

Compared  with  other  research  groups  at  Langley,  the  MPD  enthusiasts 
participated  in  more  international  scientific  conferences;  had  more  contacts 
with  consultants,  important  scientific  committees,  and  advisory  groups  out- 
side Langley;  monitored  more  research  contracts;  and  received  more  dis- 
tinguished visitors.  Senior  management  asked  MPD  researchers  to  occupy 
center  stage  during  NASA  inspections  and  to  escort  distinguished  guests  into 
their  Prankensteinian  laboratories,  which  were  filled  with  plasma  accelera- 
tors, MPD-arc  fusion  reactors,  powerful  electrical  supplies,  spectrometers, 
microwave  diagnostic  instruments,  and  other  bizarre  apparatuses.  Even  to 
other  engineers,  this  equipment  was  strange  and  unidentifiable.  Understand- 
ably, their  peers  considered  the  Ph.D.'s  and  other  'mad  scientists'  of  MPD 
a  prestigious  group.17 


*  Ironically,  neither  Wood,  head  of  the  MHD  Section,  nor  Hess,  head  of  the  Plasma  Physics  Section, 
held  a  Ph.D.  Wood  completed  all  the  course  work  toward  a  doctorate  in  the  early  1930s,  but  because  of 
the  Great  Depression  he  had  to  go  to  work  before  receiving  his  degree;  Hess  graduated  from  the  Vienna 
Institute  of  Technology  and  had  taken  graduate  courses  in  fluid  mechanics  and  thermodynamics  at  MIT 
in  the  late  1930s,  but  he  also  did  not  possess  an  advanced  degree. 


131 


Space/light  Revolution 

But  the  prestige  could  last  only  if  Langley's  MPD  work  proved  deserving; 
the  proof  lay  in  conducting  outstanding  research  programs  and  producing 
meaningful  results.  When  the  MPD  Branch  was  formed  in  1960,  Langley 
researchers  saw  three  particularly  promising  applications  for  MPD  research. 
First,  they  hoped  to  accelerate  gases  to  very  high  speeds  to  study  and 
solve  the  reentry  problems  of  intercontinental  ballistic  missiles  (ICBMs), 
spacecraft,  and  transatmospheric  or  aerospace  vehicles  such  as  the  North 
American  X-15  rocket  plane  and  the  U.S.  Air  Force  proposed  X-20  Dyna- 
Soar  boost-glider.  The  potential  for  these  applications  explains  in  part 
Langley's  commitment  to  the  small-scale  but  significant  program  of  research 
and  development  of  various  plasma  accelerators. 

Second,  the  MPD  experts  at  Langley  hoped  to  develop  prospective  ap- 
plications of  MPD  for  spacecraft  propulsion  and  power  generation  systems. 
They  were  confident  that  electric  or  ion  rockets  would  be  the  space  propul- 
sion system  of  the  future.  If  humankind  was  to  go  to  Mars  or  some  other 
planet  in  a  reasonable  travel  time,  such  radical  sorts  of  propulsion  systems 
would  be  required.  Therefore,  the  centers  for  NASA's  major  propulsion  ef- 
forts (especially  Lewis  in  Cleveland  and  Marshall  in  Huntsville)  must  begin 
studying  the  ion  and  plasma  devices  that  might  someday  offer  to  rocket 
technology  the  extraordinarily  high  specific  impulses  required  for  such  far- 
away missions.  Most  definitely,  the  design  and  operation  of  these  rockets 
would  require  the  use  of  MPD  principles.18 

Third,  Langley's  MPD  specialists  realized  that  if  controlled  thermonu- 
clear fusion  was  to  become  a  practical  source  for  the  volume  generation 
of  electricity,  much  more  about  the  subject  would  have  to  be  learned. 
Beginning  in  the  late  1950s,  the  Atomic  Energy  Commission  had  begun 
conducting  MPD  research  with  the  production  of  electric  power  in  mind. 
Branch  Head  Mike  Ellis  also  believed  that  "the  eventual  energy  source 
will  be  thermonuclear  fusion"  and  that  "the  development  of  this  energy 
source  most  likely  will  depend  upon  fundamental  discoveries  in  the  field  of 
magnetoplasmadynamics." 19 

The  promise  of  the  field  was  indeed  wonderful.  But  the  promise  of 
wonderful  or  even  revolutionary  findings  and  applications  could  sustain  the 
new  MPD  group  at  Langley  for  only  so  long.  At  some  point,  MPD  studies 
had  to  produce.  The  reality  was,  as  John  Becker  later  put  it,  "Of  all  the 
efforts  we  had,  it  was  the  most  sophisticated  and  probably  the  least  likely 
to  succeed.  We  shouldn't  have  expected  as  much  from  it  as  we  did."20 


Out  of  the  Tunnel 

Concern  for  the  problems  that  the  ICBM  encountered  during  reentry 
flight  prompted  Langley  researchers  to  begin  the  study  of  MPD  in  1958. 
The  physics  of  the  unique  conditions  of  the  hot  ionized  flow  around  the 
missile's  nose  during  reentry  demanded  special  attention.  Space  vehicles 

132 


The  "Mad  Scientists"  of  MPD 

when  reentering  the  atmosphere  quickly  became  covered  with  electrically 
charged  particles.  These  particles  formed  a  "plasma  sheath"  behind  the  bow 
shock.  Researchers  hoped  that  an  application  of  electric  and/or  magnetic 
fields  to  the  plasma  sheath  could  affect  the  airflow  in  desirable  ways;  for 
example,  it  could  reduce  the  heat  transfer  to  the  nose.  The  most  direct 
effect  of  the  plasma  sheath,  however,  was  that  radio  transmission  from  the 
vehicle  during  reentry  was  not  possible  for  obtainable  radio  frequencies.  The 
plasma  caused  a  period  of  "radio  blackout." 

To  solve  these  problems,  researchers  at  Langley  had  to  simulate  reentry 
conditions  in  the  laboratory.  This  would  require  some  new  and  unusual 
research  equipment;  conventional  wind  tunnels  would  not  do  the  job.  Small 
hypersonic  tunnels,  made  possible  by  the  development  of  high- temperature 
heat  exchangers  and  high-speed  nozzles  and  operated  on  an  intermittent 
basis  for  flow  durations  of  only  seconds  to  no  more  than  a  minute,  permitted 
studies  of  some  forces  during  reentry,  but  not  all  and  not  some  of  the  most 
important. 

Several  university,  industrial,  and  government  research  groups  had  made 
significant  advances  in  the  acceleration  of  hot  ionized  gases  by  the  late 
1950s.  Some  of  these  advances  involved  the  arc-jet,  a  novel  apparatus  for 
aerodynamic  testing  that  could  heat  a  test  gas  (usually  nitrogen,  helium, 
or  air)  to  temperatures  as  high  as  20,000°  Fahrenheit  (F).  In  essence, 
the  arc-jet  was  a  primitive  electric  rocket  engine.21  In  May  1957,  five 
months  before  Sputnik,  NACA  Langley  began  operating  a  pilot  model  of 
its  first  experimental  arc-jet.  Installed  in  Room  118  of  the  center's  Gas 
Dynamics  Laboratory,  it  was  an  "Electro-Magnetic  Hypersonic  Accelerator 
Pilot  Model  Including  Arc-Jet  Ion  Source,"  with  a  test  section  size  of  a 
minute  7x7  millimeters  and  gas  temperatures  ranging  between  10,000° 
and  12,000°F.22 

Fundamentally,  the  arc-jet  was  just  another  hot-gas  wind  tunnel,  which 
heated  the  gas  electrically  (typically  using  100,000  kilowatts)  to  high  tem- 
peratures in  a  low-velocity  settling  chamber,  and  then  expanded  it  quickly 
through  a  tiny  nozzle  to  supersonic  velocities.  No  translational  electric  or 
magnetic  forces  acted  on  the  gas  in  this  conventional  arc-jet.  The  gas  was 
simply  being  heated  by  an  electrical  discharge.  Most  of  the  charged  particles 
in  this  high-temperature  discharge  recombined  in  the  cooling  process  that 
occurred  during  expansion. 

In  1962,  Langley  tried  a  slightly  different  but  companion  arc-jet  facility 
known  as  the  hotshot  tunnel.  This  hybrid,  invented  in  the  mid-1950s  by 
engineers  at  the  U.S.  Air  Force's  Arnold  Engineering  Development  Center  in 
Tullahoma,  Tennessee,  combined  the  basic  features  of  an  arc-jet  with  those 
of  a  new  type  of  wind  tunnel  known  as  an  impulse  tunnel.  In  this  tunnel 
an  explosive  release  of  energy  created  high  pressures  and  temperatures  in 
the  test  gas.23  In  practice  Langley's  hotshot  mostly  missed  the  mark.  To 
generate  the  very  high  heat,  its  operators  had  to  resort  to  exploding  a  piece 
of  copper  through  the  tunnel  circuit,  thus  the  name  "hotshot."  The  material 

133 


Spaceflight  Revolution 


L-64-11,179 

Research  physicist  Philip  Brockman  pushes  the  button  to  start  the  MPD-arc  plasma 
accelerator  in  December  1964-  The  test  chamber  for  this  facility  was  part  of  a  larger 
high-flow,  low-vacuum  space  simulation  apparatus  housed  within  the  MPD  Branch. 

that  then  made  its  way  through  the  test  section  was  a  mixture  of  hot  air  and 
vaporized  copper,  a  very  unsatisfactory  medium  for  aerodynamic  testing. 
The  facility  remained  active  into  the  1970s,  but  the  amount  of  useful  work 
accomplished  in  it  was  quite  limited. 

Another  facility  for  reentry  testing  that  was  developed  in  the  late 
1950s  was  the  shock  tube.  Fundamentally,  this  was  an  impulse  tunnel, 
distinguished  from  a  hotshot  mainly  by  the  way  in  which  energy  was  added 
to  the  test  gas.  According  to  a  formal  definition  of  the  time,  a  shock  tube 
was  "a  relatively  long  tube  or  pipe  in  which  very  brief  high-speed  gas  flows 
are  produced  by  the  sudden  release  of  gas  at  very  high  pressure  into  a  low- 
pressure  portion  of  the  tube."  The  idea  was  to  generate  a  normal  planar 
(that  is,  lying  in  one  plane)  shock  wave  and  send  it  through  a  gas  at  a  speed 
20  to  30  times  the  speed  of  sound,  [and]  thus  heating  the  gas  behind  the 
normal  shock  to  an  extreme  temperature.24 

Langley's  first  shock  tube  began  operation  in  the  Gas  Dynamics  Lab- 
oratory in  late  1951.  By  the  end  of  the  NACA  period,  three  more  shock 
tubes  were  put  to  work  at  the  laboratory;  they  produced  temperatures  be- 
tween 10,500°  and  15,000°F,  attained  speeds  of  Mach  8  to  Mach  20,  and  had 


134 


The  "Mad  Scientists"  of  MPD 

running  times  of  0.001  to  0.002  seconds.25  Researchers  believed  that  experi- 
ments with  these  devices  would  yield  much  knowledge,  even  though  everyone 
involved  with  shock-tube  work  conceded  that  "it  was  a  very  tough  area  of 
research."  Contending  with  flows  that  lasted  for  only  a  few  thousandths  of 
a  second  and  that  required  a  considerable  amount  of  special  instrumenta- 
tion was  "a  fantastic  problem."  How  were  researchers  "to  get  answers  out  of 
something  like  that?"26  Still,  those  passionate  about  high- velocity  flows  and 
high-temperature  gases  at  Langley  put  great  faith  in  the  shock  tube.  The 
facility  was  used  for  much  basic  research  including  studies  of  shock  waves 
generated  by  atomic  bomb  blasts. 

Through  the  transition  period  of  1957  and  1958,  researchers  at  the  lab 
continued  to  seek  new  ways  to  accelerate  hot  plasmas  to  the  tremendous 
velocities  of  reentry  flight.  In  a  method  devised  by  Langley  MPD  enthusiast 
George  Wood,  a  hot  gas  was  fed  into  a  tube,  then  the  body  force  of  crossed 
electric  and  magnetic  fields  was  used  to  accelerate  the  gas  to  the  point  where 
a  mixture  of  disassociated,  high-enthalpy  flow  would  reproduce  the  very  high 
Mach  numbers  of  hypersonic  flight.  At  NASA's  First  Anniversary  Inspection 
in  1959,  Langley  engineers  demonstrated  a  crude  version  of  Wood's  crossed- 
field  plasma  accelerator.  It  produced  a  flash  of  light,  a  loud  bang,  a  startled 
audience,  and  a  belief  in  the  promise  of  major  new  scientific  findings. 

Nearly  everyone  was  excited  by  the  potential  of  plasma  accelerators. 
When  John  Stack  first  heard  about  the  facility,  he  exclaimed,  "This  is  great!" 
Stack  felt  that  Langley  should  call  the  device  something  grand;  he  proposed 
the  awe-inspiring  name,  the  "Trans- Satellite- Velocity  Wind  Tunnel."28 

Given  the  limited  performance  of  Wood's  early  version  of  the  exper- 
imental accelerator,  such  a  pretentious  name  would  have  been  a  poor 
choice.  As  part  of  a  guided  tour  for  top  officials  from  NASA  headquar- 
ters in  late  1959,  Langley  hoped  to  show  off  the  radically  new  plasma  ac- 
celeration device.  Almost  comically,  it  did  not  work.  One  embarrassed 
Langley  engineer  who  watched  the  demonstration  remembers,  "We  all  sat 
around  expectantly  while  Dr.  [Adolf]  Busemann  explained  the  system.  Then 
Busemann  went  over  and  threw  the  switch."  Unfortunately,  only  "a  little 
stream  of  red-hot  particles  sort  of  'peed  out'  the  end  of  this  tube.  It  was 
a  complete  washout.  Busemann  just  giggled  and  said,  'Well,  we  have  a 
problem.'  "29 

The  concept  behind  Wood's  crossed-field  plasma  accelerator  was  sound: 
it  was  an  application  of  a  130- year-old  theory  of  electromagnetic  force  that 
had  been  expressed  by  Ampere  in  the  1820s.  Langley  researchers  kept 
fiddling  with  the  pilot  model  until  in  1960  they  successfully  demonstrated  its 
feasibility.  Having  done  so,  they  continued  research  on  larger,  more  powerful 
versions  of  the  device.  One  version,  the  20-megawatt  plasma  accelerator,  was 
completed  in  1966  at  a  cost  of  more  than  $1  million.  With  this  facility,  the 
MPD  Branch  planned  to  achieve  more  accurate  simulation  of  the  reentry 
conditions  of  both  manned  and  unmanned  vehicles.  Shakedown  testing  in 
the  accelerator  continued  until  1969,  when  political  pressures  applied  by  the 

135 


Space/light  Revolution 


George  P.  Wood  (right),  head  of  MPD's 
Magnetohydrodynamics  Section,  developed 
Langley's  earliest  crossed- field  plasma  ac- 
celerator. The  accelerator  section  of  the 
20-megawatt  plasma  accelerator  facility  is 
shown  below.  Note  the  many  electrodes  for 
furnishing  the  high-energy  electric  field. 


L-62-8047 


L-66-1913 


136 


The  "Mad  Scientists"  of  MPD 


L-63-3361 


In  this  April  1963  photo,  MPD  lab  tech- 
nician Charlie  Diggs  regulates  the  flow 
of  a  test  gas  in  an  early  10-kilowatt  test 
version  of  Langley's  Hall-current  plasma 
accelerator  (above);  over  his  left  shoulder 
sits  a  Polaroid  camera  for  photograph- 
ing an  oscilloscope.  In  November  1965, 
an  unidentified  technician  (left)  wears 
goggles  to  protect  his  eyes  against  the  in- 
tense light  in  a  later  coaxial  version  of  a 
Hall-current  plasma  accelerator.  In  the 
test  section,  one  can  see  the  very  bright, 
high-velocity  plume  from  the  MPD  arc- 
jet  exhausting  into  a  vacuum  tank. 


L-65-8509 


137 


Space/light  Revolution 

Nixon  administration  forced  an  abrupt  halt  to  the  accelerator's  pioneering 
work.  Whether  the  machine  would  have  ever  completely  panned  out,  no  one 
can  be  sure. 

In  NASA's  report  on  the  last  tests  made  in  this  device,  published  in  1971, 
George  Wood  and  his  colleagues  pointed  out  that  an  exit  velocity  of  30,176 
feet  per  second  had  been  achieved,  which  was  a  remarkable  81  percent  of 
the  facility's  computed  capacity  of  37,064  feet  per  second.  According  to  the 
NASA  report,  the  crossed-field  accelerator  "appears  to  be  the  largest  and 
highest  velocity  nonpulsed  linear  plasma  accelerator"  to  attain  "an  operable 
status."30  An  experimental  facility  with  this  record  must  be  called  a  success. 

While  trying  to  work  out  the  kinks  in  Wood's  crossed-field  accelerator 
design,  Langley's  MPD  experts  conceived  several  other  methods  for  accel- 
erating plasmas.  One  of  these  methods,  which  was  not  pursued  very  far, 
they  called  "microwave  cavity  resonance."  The  major  alternative,  however, 
was  known  as  the  "linear  Hall-current  accelerator."  This  type  of  plasma 
accelerator  was  based  on  a  principle  of  electrical  polarization  and  current 
generation  laid  out  by  the  American  physicist  Edwin  H.  Hall  in  the  1920s 
and  1930s.  The  facility  used  a  constant  rather  than  intermittent  interaction 
of  currents  and  magnetic  fields  across  a  channel  to  accelerate  a  steady  flow 
of  plasma. 

Beginning  in  the  late  1950s,  a  small  group  of  Langley  researchers  led 
by  Robert  V.  Hess,  an  applied  physicist  from  Austria  who  had  come  to 
work  for  the  NACA  in  1945,  began  pursuing  two  major  variants  of  the 
Hall  accelerator:  the  MPD  arc  and  the  so-called  linear  Hall  accelerator. 
Throughout  the  1960s,  Hess  and  his  associates  refined  these  versions  of 
the  plasma  accelerator,  thus  making  extensive  experimental  and  theoretical 
studies  of  the  physics  and  overall  performance  of  their  devices.  Although 
they  successfully  demonstrated  the  efficiency  of  the  MPD  arc  and  linear  Hall 
accelerator  and  made  several  important  findings  relating  to  the  manner  in 
which  oscillations  and  instabilities  in  plasma  could  develop  into  turbulent 
flows,  MPD  researchers  were  never  able  to  simulate  reentry  conditions  or 
the  interaction  between  the  solar  wind  and  the  geomagnetosphere,  and  they 
would  never  realize  meaningful  applications  in  space  propulsion.  As  was 
the  case  with  the  other  MPD  experimental  facilities  mentioned,  the  linear 
Hall-current  accelerator  possessed  limitations  that  Hess  and  his  colleagues 
could  not  eradicate.  By  the  late  1960s,  Hess  and  others  in  MPD  shifted  the 
focus  of  their  work  with  these  accelerators  to  the  potential  application  of 
gas  lasers.31 

Into  the  Cyanogen  Fire 

In  the  late  1950s,  the  Langley  MPD  group  found  a  stopgap  method  of 
generating  a  plasma  in  the  laboratory.  This  method  involved  the  production 
of  a  hot  flame  fueled  by  the  combustion  of  cyanogen  gas  and  oxygen. 


138 


The  "Mad  Scientists"  of  MPD 


Robert    V.    Hess,    head   of  MPD's    Plasma 
Physics  Section. 
L-62-8120 


MPD  physicist  Bob  Hess  was  an  intense  researcher  and  bibliophile.  He 
combed  the  current  technical  and  scientific  literature  for  ideas  that  might 
prove  useful  to  his  and  his  colleagues'  work.  Proficient  in  German  and  French 
as  well  as  English,  he  was  able  to  keep  abreast  of  scientific  ideas  along  several 
fronts.  With  his  desk  piled  high  with  papers,  Hess  ferreted  out  the  best 
notions,  and  massaged  them  for  his  own  creative  uses.*  In  1957,  Hess  came 
across  a  reference  to  a  new  experimental  device  at  the  Research  Institute 
of  Temple  University  in  Philadelphia.  This  device  produced  an  extremely 
hot  flame  by  burning  oxygen  with  cyanogen,  a  colorless,  flammable,  and 
poisonous  gas,  sometimes  formed  by  heating  mercuric  cyanide.  After  reading 
about  the  cyanogen  flame  experiment,  Hess  hit  on  an  idea  for  adapting  the 
flame  to  create  a  hot  plasma  for  simulating  the  space  reentry  environment. 
By  feeding  oxygen  and  cyanogen  gas  into  a  combustion  chamber  and  igniting 
the  mix,  the  researchers  at  Temple  were  producing  a  flame  of  more  than 
8000° F.  This  was  one  of  the  hottest  flames  scientists  had  ever  produced. 
What  would  be  the  result,  Hess  mused,  if  a  potassium  vapor  that  ionized 
easily  at  that  temperature  was  added  to  the  combustion  chamber?  Would 


For  example,  in  1945  Hess  found  an  overlooked  British  translation  of  German  aerodynamicist  Dr. 
Adolf  Busemann's  seminal  1937  paper  on  sweptwing  theory.  Hess  found  it  in  the  Langley  Technical 
Library,  where  his  future  wife,  Jane,  would  someday  serve  as  the  head  librarian  and  assist  him  greatly 
with  his  search  for  references,  and  he  passed  it  on  to  colleague  Robert  T.  Jones.  This  was  just  prior 
to  Jones's  final  revision  of  a  confidential  NACA  paper  in  which  Jones  would  report  his  independent 
discovery  of  the  advantages  of  wing  sweep  for  supersonic  flight. 


139 


Space/light  Revolution 

this  create  a  jet  of  hot  gas  that  reproduced  the  extremely  ionized  plasma 
conditions  of  missile  reentry? 

On  17  June  1957,  Hess  and  his  boss  in  the  Gas  Dynamics  Laboratory, 
Macon  C.  Ellis,  Jr.,  visited  Temple  University  to  discuss  the  details  of 
producing  a  cyanogen-oxygen  flame  and  to  inquire  about  the  feasibility  of 
adding  an  easily  ionizable  alkaline  material,  potassium  or  perhaps  cesium,  to 
the  flame.  The  key  people  to  whom  they  spoke  were  Dr.  Aristid  V.  Grosse, 
director  of  the  Temple  Research  Institute,  and  Charles  S.  Stokes,  who  was  in 
charge  of  the  cyanogen  flame  program.  Grosse  and  Stokes  agreed  that  "the 
great  stability  of  the  combustion  products"  made  them  "well  suited"  for  an 
addition  of  an  ionizer  such  as  potassium;  they  told  the  Langley  visitors  that 
they  themselves  had  recognized  this  in  one  of  their  early  reports,  perhaps 
in  the  one  that  Hess  had  read.  However,  they  had  not  made  quantitative 
estimates  of  the  electron  densities  or  followed  up  on  the  idea  in  any  way. 
They  wondered  whether  the  addition  of  potassium  might  not  exert  a  cooling 
effect  that  would  somewhat  diminish  the  density  of  electrons.  Hess,  however, 
had  already  made  the  estimates  and  knew  that  the  density  of  the  electrons  in 
the  seeded  cyanogen  flame  would  be  sufficiently  high  (about  1016  per  cubic 
centimeter)  to  compensate  for  any  temperature-reducing  reactions.32 

At  Langley,  Paul  Huber  with  the  help  of  the  facilities  engineering  group 
quickly  designed  a  cyanogen  flame  apparatus,  and  the  funding  for  its 
construction  was  approved.  By  the  time  the  NACA  became  NASA,  the 
device  had  been  operating  for  several  months.  As  expected,  the  first  major 
test  program  conducted  in  Langley's  alkali-metal-seeded,  cyanogen-oxygen 
flame  explored  how  flow-field  conditions  near  an  ICBM  nose  prevented  the 
transmission  of  radio  signals  back  to  earth.  Researchers  in  the  Gas  Dynamics 
Laboratory  working  with  Joseph  Burlock  of  IRD  mounted  a  transmitting 
antenna  in  front  of  a  nozzle  that  bathed  the  antenna  in  the  hot  cyanogen 
gas  jet.  Instruments  then  measured  the  rate  at  which  the  transmitter  lost 
its  signal  power. 

The  early  MPD  test  program  demonstrated  the  feasibility  of  creating 
and  controlling  the  highly  ionized  plasmas  representative  of  the  extreme 
dynamic  conditions  of  spaceflight  and  reentry.  The  program  also  showed 
that  certain  simplified  theoretical  methods  could  be  used  to  calculate  the 
loss  of  electronic  communication  with  a  vehicle  during  reentry  of  a  vehicle 
from  space.  If  plasma  conditions  around  the  vehicle  could  be  estimated 
with  reasonable  accuracy,  researchers  then  would  be  able  to  predict  the 
expected  radio  power  loss.  This  was  critical  information  for  trips  in  and  out 
of  space  by  guided  missiles,  aerospace  planes,  and  manned  and  unmanned 
spacecraft.  Led  by  the  outstanding  theoreticians  Calvin  T.  Swift  and  John  S. 
Evans,  who  worked  in  the  Plasma  Applications  Section  under  Paul  Huber, 
MPD  researchers  at  Langley  continued  to  make  significant  contributions 
throughout  the  1960s.  On  the  problems  of  transmitting  radio  signals  to  and 
from  reentry  vehicles,  no  group  inside  or  outside  of  NASA  came  to  speak 
with  more  authority.33 

140 


The  "Mad  Scientists"  of  MPD 


L-63-2898 

Three-quarter  top  view  of  Langley's  cyanogen  burner,  which  was  located  for  safety 
reasons  in  a  remote  spot  on  the  edge  of  a  marsh  in  Langley's  West  Area.  To  the  left 
of  the  jet  is  a  microwave  "horn, "  a  device  for  electron- concentration  measurement 
and  radio-transmission  attenuation. 

The  MPD  program  was  particularly  valuable  to  the  little-known  NASA 
project  RAM.  Initiated  too  late  to  help  in  the  communications  blackout 
problems  of  the  Mercury  and  Gemini  capsules,  the  purpose  of  Project 
RAM  was  to  support  the  Apollo  program.  Many  of  the  project's  results 
proved  inconclusive,  and  most  of  the  hoped-for  technological  fixes,  for 
example,  the  use  of  higher  radio  frequencies  and  the  timed  injection  of 
small  sprays  of  water  into  the  hot  gas  envelope  surrounding  a  reentering 
spacecraft,  were  judged  too  problematic  for  use  in  Apollo.  However,  MPD 
specialists  at  Langley  did  learn  how  to  predict  the  flow-field  characteristics  of 
a  reentering  spacecraft  more  accurately,  and  their  work  led  to  viable  schemes 
for  alleviating  or  "quenching"  part  of  the  plasma  sheath  so  that  some  level 
of  effective  radio  communications  to  and  from  a  reentering  vehicle  could 
occur.34  Experience  gained  in  the  MPD  reentry  experiments  of  the  1960s 
eventually  aided  in  projecting  the  reentry  conditions  of  the  Space  Shuttle. 


141 


Spaceflight  Revolution 

The  Barium  Cloud  Experiment 

Not  all  of  Langley's  MPD  work  sought  such  direct  technological  appli- 
cations as  Project  RAM.  Some  of  the  more  fruitful  research  efforts  fell  into 
the  realm  of  basic  science  and  represented  what  MPD  Branch  Head  Ellis  de- 
scribed in  a  February  1962  briefing  to  the  Langley  senior  staff  as  "examples 
of  keeping  research  alive  on  a  reasonable  scale  without  solid,  specific  appli- 
cations or  even  the  guarantee  of  applications!"35  One  such  effort  that  made 
significant  contributions  was  a  barium  cloud  experiment  designed  for  explo- 
ration of  the  interaction  between  the  solar  wind  and  the  earth's  magnetic 
fields. 

Although  a  continuous  outpouring  of  plasma  appeared  to  emanate  from 
the  sun  (i.e.,  the  solar  wind),  this  plasma  by  virtue  of  its  high  conductivity 
did  not  seem  to  penetrate  the  earth's  strong  magnetic  field;  instead,  the 
solar  wind  flowed  around  the  earth's  field,  forming  a  huge  cavity.  Sensitive 
magnetometers  aboard  some  of  the  first  Soviet  and  American  spacecraft 
provided  useful  information  about  the  disposition  of  the  magnetic  fields 
within  this  cavity;  however,  many  questions  about  the  arrangement  of  the 
field  lines  remained  unanswered.  Conservative  estimates  of  the  volume  of  the 
cavity  placed  it  at  about  60,000  times  the  volume  of  the  earth.  Langley's 
interested  MPD  experts  knew  that  it  was  "going  to  be  a  formidable  task 
indeed  to  map  such  an  extensive  field  by  point  to  point  samplings."36  Little 
was  known  about  the  shape  of  the  cavity  on  the  nightside  of  the  earth, 
and  indeed  astrophysicists  had  suggested  that  the  cavity  was  in  fact  open 
and  that  the  earth's  magnetosphere  had  a  tail  extending  out  some  several 
"astronomical  units."* 

These  were  only  some  of  the  complications  stirring  the  "intellectual  stew" 
over  the  magnetospheric  cavity.  Other  concerns  stemmed  from  evidence  that 
the  magnetic  field  lines  of  the  earth  were  linked  at  least  partially  with  those 
of  the  interplanetary  field,  which  in  turn  were  entrained  in  the  solar  wind.  If 
so,  tangential  stresses  and  drag  forces  in  the  realm  of  space  affected  motions 
within  the  magnetosphere  in  addition  to  those  imparted  by  the  earth's  own 
rotation,  which  were  themselves  unknown.37 

At  Langley,  these  cosmological  matters  were  of  particular  interest  to 
the  small  group  of  theoretically  inclined  researchers  working  in  the  MPD 
Space  Physics  Section  under  David  Adamson.  Beginning  in  late  1963,  the 
Adamson  group  began  to  seriously  consider  a  novel  experimental  technique 
by  which  scientists  could  use  an  artificially  ionized  plasma  "cloud"  as  a  space 
probe.  As  Adamson  explained  at  the  time,  the  principle  of  the  cloud  was 
rather  simple. 


Of 

An  astronomical  unit  is  usually  defined  as  the  mean  distance  between  the  center  of  the  earth  and 
the  center  of  the  sun,  i.e.,  the  semimajor  axis  of  the  earth's  orbit,  which  is  equal  to  approximately 
92.9  X  1,000,000  miles  or  499.01  light  seconds. 


142 


The  "Mad  Scientists"  of  MPD 

If  a  charged  particle  is  projected  into  a  magnetic  field,  it  spirals  along  a  magnetic 
field  line,  remaining  tied  to  that  field  line  until  it  is  dislodged  by  colliding  with 
another  particle.  Picture  then  a  cloud  of  charged  particles,  sufficiently  dispersed  at 
a  sufficiently  great  altitude  that  collisions  can  be  ignored.  The  individual  particles 
will  be  tied  to  the  field  lines,  and  motions  of  the  cloud  perpendicular  to  the  field  lines 
will  be  inhibited.  Of  course,  the  cloud  can  and  will  diffuse  along  the  field  lines,  and 
as  it  does  so  will  serve  to  define  the  shaping  of  those  field  lines  to  which  it  is  frozen. 
Moreover,  if  the  magnetic  field  lines  are  themselves  in  motion,  this  motion,  too,  will 

oo 

be  imparted  to  the  cloud. 

Only  three  requirements  were  placed  on  the  cloud:  it  had  to  be  fully  ionized, 
the  ionized  atoms  had  to  show  resonance  lines  in  the  visible  portion  of  the 
spectrum,  and  it  had  to  be  visible  to  observers  on  earth. 

The  notion  of  an  ionized  cloud  was  not  new.  For  several  years,  research 
groups  around  the  world  had  been  experimenting  with  chemical  releases  as 
a  means  of  exploring  the  nature  of  the  upper  atmosphere.  For  the  most 
part,  the  creation  of  such  artificial  clouds  was  done  by  launching  a  sounding 
rocket  carrying  on  its  nose  a  payload  of  pyrotechnic  constituents  mixed  with 
alkali  metals.  At  the  proper  altitude  in  the  upper  atmosphere,  a  canister 
carrying  the  payload  would  be  ejected.  The  temperature  of  the  canister's 
contents  would  rise  thousands  of  degrees  and  then  escape  explosively  to  form 
a  colorful  vapor  whose  atoms  would  glow  blue- violet  in  the  sunlight.  The 
result  was  a  bright  and  rather  beautiful  space  cloud,  a  sort  of  instant  aurora, 
which  could  be  seen  quite  distinctly  by  an  observer  watching  from  the 
nightside  of  the  earth.  Highly  responsive  magnetometers  and  spectroscopes 
could  then  be  used  to  analyze  the  physics  of  what  happened  when  a  body  of 
charged  particles  exploded  in  the  outermost  realms  of  the  earth's  atmosphere 
and  at  the  fringes  of  space. 

The  world  leaders  in  developing  the  tricky  optical  cloud  technique  were 
the  West  Germans,  specifically  a  group  of  experimental  astrophysicists  in 
the  Gaerching  Laboratory  of  the  Max  Planck  Institut  in  Berlin.  The  leading 
figures  in  the  development  of  what  came  to  be  known  as  "the  barium 
bomb"  were  Dr.  Ludwig  Biermann  and  his  associate  Dr.  Riemar  Lust.  In 
1951,  Biermann  had  anticipated  Parker's  discovery  of  the  solar  wind  by 
hypothesizing  that  a  comet's  tail,  which  always  points  away  from  the  sun, 
was  being  pushed  by  streams  of  solar  particles.  He  spent  the  rest  of  the 
decade  looking  for  an  experimental  means  by  which  to  prove  his  theory. 
By  the  late  1950s,  the  Biermann  group  had  developed  a  technique  for  the 
creation  of  an  artificially  ionized  cloud  in  the  upper  atmosphere.  By  1964, 
although  the  existence  of  the  solar  wind  was  by  then  taken  for  granted,  the 
same  group  was  ready  to  use  more  powerful  rockets  to  deploy  the  first  of 
these  clouds  in  space. 

Biermann  and  Lust  used  a  payload  of  barium  inside  their  canisters.  In 
their  opinion,  a  mix  of  copper  oxide  and  barium  (a  soft,  silver-white,  metallic 
element  obtained  when  its  chloride  was  decomposed  by  an  electric  current) 


143 


Spaceflight  Revolution 

was  most  desirable  because  it  ionized  at  a  reasonable  temperature  and  even 
in  modest  concentration  could  produce  clouds  visible  to  observers  on  earth. 
In  the  early  1960s,  French  sounding  rockets  fired  from  the  Sahara  began 
carrying  West  Germany's  barium  payloads  into  the  upper  atmosphere  as 
part  of  a  research  program  sponsored  by  the  newly  founded  European  Space 
Research  Organization  (ESRO).  NASA's  space  scientists  naturally  knew 
about  the  European  program,  and  some  of  them  thought,  like  Biermann  and 
Lust,  that  the  barium  cloud  technique  could  be  adapted  for  experimental 

•JQ 

use  in  space. 

In  the  summer  of  1964,  Bob  Hess  traveled  to  Feldafing,  near  Munich, 
Germany,  to  participate  in  an  international  symposium  on  the  diffusion  of 
plasma  across  a  magnetic  field.  At  this  meeting,  Hess  spoke  with  Biermann 
about  the  barium  cloud  technique.  The  interest  of  the  West  Germans  in 
the  experiment  was  different  from  that  of  Langley's  MPD  Branch.  The 
Germans  wanted  to  release  barium  in  the  streaming  solar  wind  outside 
the  magnetosphere  in  the  hope  of  learning  more  about  the  formation  of 
comets;  the  NASA  researchers  sought  to  explore  the  magnetosphere  itself. 
Nevertheless,  the  interests  were  similar  enough  to  make  Biermann  and  Hess 
agree  that  some  measure  of  international  cooperation  would  be  useful. 

Upon  his  return  to  Langley,  Hess  wrote  a  letter  to  NASA's  Space 
Sciences  Steering  Committee.  Founded  in  May  1960  by  the  head  of  NASA's 
Office  of  Space  Sciences  and  Applications  (OSSA),  Dr.  Homer  Newell,  this 
committee  consisted  of  NASA  officials  and  leading  academic  scientists  in 
the  field.  Their  duty  was  to  advise  NASA  on  its  space  science  program 
and  evaluate  proposals  for  scientific  experiments  on  NASA  missions.  In  his 
letter,  Hess  summarized  the  observational  possibilities  of  plasma  clouds  as 
magnetospheric  probes  and  proposed  that  NASA  devise  a  cloud  experiment, 
which  perhaps  could  be  done  with  the  cooperation  of  Dr.  Biermann  and 
the  West  Germans.  Instead  of  launching  the  barium  from  inside  a  rocket, 
Hess  suggested  that  the  makings  for  the  plasma  cloud  be  released  from 
inside  the  MORL  that  NASA  was  planning,  "where  the  advantages  of  longer 
observation  of  the  plasma  cloud  and  of  a  wider  choice  of  materials  are  offered 
as  compared  with  observation  from  the  ground  through  the  atmosphere."40 

The  space  scientists  at  NASA  headquarters  were  interested  in  the  general 
idea,  but  plans  to  proceed  progressed  slowly  through  1965  and  1966.  Other 
space  science  experiments  more  directly  supportive  of  the  Apollo  lunar 
landing  program,  like  the  Surveyor  and  Lunar  Orbiter  programs,  received 
the  highest  priority.  Still,  the  MPD  Branch  in  conjunction  with  the 
appropriate  program  officers  at  NASA  headquarters,  as  well  as  with  the 
technical  support  of  the  Applied  Materials  and  Physics  Division  at  Langley, 
continued  to  plan  for  the  cloud  experiment.  From  Wallops  Island,  NASA 
would  launch  an  explosive  canister  atop  a  high-altitude  rocket.*  Early  on, 


The  type  of  rocket  was  yet  to  be  determined.    Ultimately,  several  sounding  rockets,  as  well  as 
Langley's  multipurpose  Scout  rocket,  would  be  used. 

144 


The  "Mad  Scientists"  of  MPD 

Langley  researchers  thought  that  the  canister  should  contain  a  combustible 
mixture  of  cyanogen-oxygen  with  cesium;  however,  with  input  from  Lust's 
team  in  Germany,  they  finally  chose  a  barium  payload.  At  the  appropriate 
altitude  in  space  (the  rocket  would  not  go  into  orbit),  the  canister  would 
detonate  and  out  would  float  the  ionized  particles  which  would  form  the 
space  cloud.  The  cloud  would  last  several  minutes  to  more  than  one  hour 
during  which  it  would  reflect  radio  waves  and  could  be  viewed  from  a  location 
on  earth  in  the  sun's  shadow. 

The  general  scientific  purpose  of  the  cloud  would  be  to  serve  as  "a 
ready  means  of  discerning  on  a  large  scale  the  topology  of  the  earth's 
[magnetic]  field  and  of  determining  magnetospheric  motions."  However, 
Langley's  MPD  group  felt  that  the  cloud  might  also  be  used  as  an  aid  in 
tracking  high-altitude  vertical  sounding  rockets  or  even  vehicles  (hopefully 
not  Soviet)  bound  for  the  moon.  It  could  be  used  as  a  form  of  visible  tracer, 
not  altogether  unlike  the  use  of  certain  metallic  elements  (often  barium) 
and  radioactive  isotopes  fed  into  the  stomach  or  injected  into  the  blood  as 
tracers  for  X-ray  diagnosis  of  cancer  and  other  diseases.41 

Eventually,  NASA  gave  the  go-ahead  for  the  barium  cloud-in-space 
experiment.  The  approval  was  in  part  politically  motivated;  NASA  wanted 
to  encourage  international  cooperation,  at  least  in  certain  noncritical  space 
endeavors,  and  especially  with  the  democratic  nations  of  western  Europe. 
In  June  1965,  representatives  of  the  Max  Planck  Institut  approached  NASA 
with  a  proposal  for  a  joint  barium  cloud  experiment  involving  German 
pay  loads  and  NASA  launches  from  Nike- Tomahawk  and  Javelin  rockets. 
The  following  month,  NASA  and  West  Germany's  Federal  Ministry  for 
Scientific  Research  signed  a  memorandum  of  understanding  calling  for 
cooperation  in  a  program  of  space  research  on  the  earth's  inner  radiation 
belts  and  aurora  borealis.  According  to  the  memorandum,  NASA  would 
provide  a  Scout  booster  for  the  launch  of  a  German-made  satellite  into 
polar  orbit  by  1968,  with  the  results  of  the  experiment  to  be  made  available 
to  the  world  scientific  community.  Pursuant  to  another  memorandum 
of  understanding  between  the  two  nations  (signed  in  May  1966),  the 
two  research  agencies  would  then  proceed  with  investigations  of  cometary 
phenomena,  the  earth's  magnetosphere,  and  the  interplanetary  medium 
through  studies  of  the  behavior  of  high- altitude  ionized  clouds.42 

Four  months  later,  on  24  September  1966,  in  a  joint  effort  with  the  Max 
Planck  Institut,  NASA  launched  a  four-stage  Javelin  sounding  rocket  from 
Wallops  Island  to  check  its  canister-ejection  technique,  and  on  the  next  day, 
again  from  Wallops,  launched  a  Nike- Tomahawk  rocket  which  released  a 
mixture  of  barium  and  copper  oxide.  The  second  "shot"  only  reached  160 
miles,  whereas  the  desirable  altitude  for  a  barium  cloud  release  was  3  to  5 
earth  radii.  Nonetheless,  the  experiment  was  successful.  For  hundreds  of 
miles  up  and  down  the  Atlantic  coast,  three  distinct  clouds  were  visible. 
NASA  and  West  German  scientists  photographed  the  clouds  in  an  effort  to 
track  and  measure  electric  fields  and  wind  motions  in  the  upper  atmosphere. 

145 


Space/light  Revolution 

The  results  of  both  launches  caused  quite  a  public  stir.  Some  residents  along 
the  coast  reported  sightings  of  brilliant  UFOs,  and  some  motorists  became 
so  fascinated  by  the  brightly  colored  clouds  that  they  ran  off  the  road. 

What  came  to  be  known  formally  as  the  MPI  (Max  Planck  Insti- 
tut)/NASA  Magnetospheric  Ion  Cloud  Experiment  was  the  next  step  in  the 
two  parties'  cooperative  investigation.  Proposed  formally  by  the  Germans 
in  February  1967,  the  joint  experiment  was  not  approved  by  NASA  until 
December  1968.  According  to  the  final  agreement,  the  Germans  would  pro- 
vide the  barium  payload,  two  ground  observer  stations,  and  data  analysis; 
NASA  would  furnish  the  rocket,  conduct  the  launch  from  Wallops  Island, 
and  provide  tracking  and  communications  services.43 

Despite  a  fatal  explosion  on  5  October  1967,  at  the  Downey,  California, 
plant  of  North  American  Rockwell,  which  was  caused  by  a  mishandling 
of  finely  divided  barium  mixed  with  Freon,  the  barium  cloud  experiment 
eventually  proved  a  great  success.44  On  17  March  1969,  a  barium  cloud 
1865  miles  long,  lasting  some  20  minutes,  and  visible  to  the  naked  eye, 
formed  at  an  altitude  of  43,495.9  miles  (69,999.87  kilometers).  Heos  1, 
a  "Highly  Eccentric  Orbiting  Satellite"  belonging  to  ESRO,  carried  the 
cloud-producing  canister  into  space.  Instrumented  observation  of  this 
and  subsequent  plasma  cloud- in-space  experiments  revealed  the  motions 
of  the  earth's  magnetic  field  lines,  including  those  influencing  the  aurorae; 
demonstrated  other  plasma  effects  in  space;  helped  scientists  to  correlate 
these  motions  and  effects  as  a  function  of  solar  flares;  and  generally  allowed 
world  astrophysicists  to  model  the  geomagnetosphere  more  accurately.  All 
the  barium  cloud  shots  generated  considerable  public  concern  and  interest 
and  were  widely  announced  in  advance  in  the  press. 

Aside  from  fascinating  the  public,  this  experimental  probing  of  the  near- 
earth  environment  of  space  also  led  researchers  to  explore  what  was  believed 
to  be  the  great  potential  value  of  magnetospheric  data  for  understanding  and 
perhaps  even  controlling  the  earth's  weather.  Although  the  energies  in  space 
were  recognized  to  be  small  compared  with  those  in  the  atmosphere,  those 
researchers  interpreting  the  results  of  the  barium  cloud  experiment  raised 
the  possibility  that  even  small  disturbances  of  inherently  unstable  regions 
in  space  could  trigger  significant  behavior  in  large  regions  around  the  earth. 

The  few  people  outside  Langley  who  remember  the  barium  cloud  research 
program  believed  NASA  left  most  of  the  interpretation  of  the  results  to  the 
Germans.  In  truth,  as  the  NASA  reports  on  the  program  demonstrate, 
Langley's  space  physics  group  moved  ahead  very  quickly  to  interpret  the 
data,  "scooping"  the  preeminent  Germans  by  first  reporting  and  explaining 
in  full  many  of  the  essential  findings.45 

NASA  Langley  planned  to  participate  in  at  least  one  follow-on  barium 
cloud  test  in  1974  or  1975.  The  purpose  of  this  proposed  test  was  to  shape 
the  barium  charge  along  a  magnetic  field  line,  then  time  the  discharge  to 
coincide  exactly  with  the  passage  of  an  unmanned  satellite  having  a  very 
high-frequency  (VHF)  receiver  aboard.  The  receiver  would  measure  the 

146 


The  "Mad  Scientists"  of  MPD 

cyclotron  radiation  from  the  electrons  circling  the  field  line.  The  test  did 
not  take  place,  however,  because  of  the  lack  of  support  in  the  OSSA  at 
NASA  headquarters.46 


The  Search  for  Boundless  Energy 

Astrophysics  was  not  the  only  driving  force  behind  the  explosion  of  MPD 
research  in  the  1950s.  Another  inciting  factor  was  the  quest  for  atomic 
energy.  After  World  War  II  and  the  dawn  of  the  atomic  age,  many  physicists 
had  begun  exploring  ways  to  confine  plasmas  magnetically  in  a  new  sort 
of  nuclear  reactor  based  not  on  fission  but  on  fusion.  Such  projects  were 
designed  to  explore  the  potential  of  generating  thermonuclear  power.  Many 
researchers  and  institutions  believed  this  was  the  pot  of  gold  at  the  end  of 
the  MPD  rainbow. 

In  1951  the  four-year-old  U.S.  Atomic  Energy  Commission  initiated  a 
secret  project  known  by  the  code  name  "Sherwood" ;  its  ambitious  objective 
was  the  controlled  release  of  nuclear  energy  through  stable  confinement 
of  plasmas  at  an  extremely  high  temperature.  Interestingly,  the  strategy 
behind  Project  Sherwood  was  not  to  build  a  scientific  and  technical  base  for 
advanced  fusion  experiments;  rather,  the  goal  was  to  immediately  develop  a 
working  technology.  Researchers  were  "to  invent  their  way  to  a  reactor,"  so 
to  speak,  just  as  the  scientists  and  engineers  through  a  crash  effort  had  built 
the  first  atomic  bomb.  Such  was  the  mood  of  optimism  and  enthusiasm  over 
the  human  capacity  for  solving  any  problem,  however  monumental,  in  the 
wake  of  the  successful  Manhattan  Project.47 

Many  of  the  devices  developed  during  Project  Sherwood  served  as  ad- 
vanced research  tools.  Although  highly  varied  in  their  designs,  almost  all 
the  facilities  tried,  with  only  partial  success,  to  produce  fusion  reactions 
through  some  type  of  magnetic  containment  of  a  plasma.  By  the  1950s, 
scientists  knew  that  a  thermonuclear  reactor  would  require  a  reacting  gas 
with  a  temperature  of  at  least  1,000,000,000  kelvin  (K).  Because  contain- 
ment of  such  an  extraordinarily  hot  gas  by  solid  walls  seemed  impossible, 
many  plasma  physicists  believed  that  the  only  way  to  contain  the  gas  was 
by  powerful  magnetic  forces.  Further  work  in  MPD  became  vital. 

At  Langley,  as  elsewhere,  researchers  turned  to  the  sun  (a  giant  fusion 
reactor)  to  find  the  answers.  In  Langley  laboratories,  the  MPD  group 
worked  on  designing  facilities  that  would  simulate  the  activity  of  the  solar 
corona.  George  Wood's  MHD  Section  built  several  highly  experimental 
devices  to  study  solar  physics;  however,  none  of  them  yielded  the  secret 
of  thermonuclear  power. 

Consider,  for  example,  George  Wood's  first  highly  experimental  facility 
for  the  basic  study  of  solar-coronal  physics,  the  one-megajoule  theta-pinch. 


147 


Space/light  Revolution 

"The  pinch"  used  a  powerful,  one- million- joule*  discharge  of  direct-current 
(DC)  electricity  along  a  single-turn  coil  to  generate  a  strong  longitudinal 
magnetic  field.  Wood's  section  hoped  that  an  interaction  of  this  high- 
density  current  with  its  own  magnetic  field  would  cause  a  contained  column 
of  plasma  (that  is,  a  molten  conductor)  to  self-contract  and  become  pinched 
even  tighter  and  perhaps  even  to  rupture  itself  momentarily,  thus  producing 
a  controlled  fusion  reaction. 

First  explained  in  a  theoretical  paper  by  American  physicist  Willard  H. 
Bennett  in  1934,  the  application  of  this  self-focusing  pinch  effect  had  become 
a  basic  mechanism  of  plasma  and  plasma-containment  research  worldwide 
in  the  1950s.  Langley's  MPD  enthusiasts  (notably  MHD's  Nelson  Jalufka) 
naturally  wanted  to  get  involved.  Unfortunately,  research  in  the  Langley 
pinch  facility,  as  in  all  other  reactors  of  the  time  designed  to  generate 
controlled  nuclear  fusion,  did  not  lead  to  fundamental  breakthroughs.  It 
did,  however,  make  some  solid  contributions  to  the  literature.48 

Another  device  that  perhaps  did  not  live  up  to  all  expectations  but 
nonetheless  succeeded  in  fundamental  respects  was  Langley's  Magnetic 
Compression  Experiment.  In  the  early  1960s,  Karlheinz  Thorn,  Goetz 
Oertel,  and  George  Wood  devised  an  experimental  apparatus  capable  of 
generating  a  multimillion-degree-kelvin  plasma  for  simulation  of  the  solar 
corona  and  for  studying  the  processes  that  produce  highly  ionized  atoms  in 
the  corona.  Completed  in  1965  at  a  total  cost  of  roughly  $2  million,  the 
apparatus  consisted  of  a  one-megajoule  capacitor  bank  (a  device  for  storing 
electrical  energy)  plus  a  straight  narrow  tube  that  produced  a  theta-pinch. 
Experiments  conducted  with  this  device  led  to  some  significant  results  on  the 
spectral  lines  of  highly  ionized  gases  like  deuterium  and  argon,  and  members 
of  Wood's  MPD  group  published  several  papers  on  the  experiments  into  the 
late  1960s.  Well  after  the  dissolution  of  the  MPD  branch  in  1970,  the  facility 
was  still  operating,  thanks  largely  to  the  support  of  Karlheinz  Thorn,  who 
had  moved  to  a  position  of  partronage  in  the  OSSA  at  NASA  headquarters. 
Thorn  was  able  to  keep  the  Magnetic  Compression  Experiment  alive  by 
relating  its  research  more  directly  to  astrophysics,  thereby  circumventing 
a  policy  of  the  Nixon  administration  against  basic  research  in  the  highly 
speculative  energy  field  of  thermonuclear  fusion.49 

A  third  important  fusion  research  effort  of  the  MHD  Section  involved 
the  plasma-focus  research  facility.  Although  the  stated  purpose  of  this  fa- 
cility (whose  operation  dates  to  the  mid-1960s)  was  to  simulate  and  study 
the  physics  of  solar  flares,  its  real  purpose  from  the  outset  was  to  explore 


A  joule  is  equivalent  to  one  watt-second. 


148 


The  "Mad  Scientists"  of  MPD 

the  possibilities  of  fusion.*  Essentially,  the  plasma-focus  apparatus  was  a 
coaxial  arrangement  wherein  a  sheet  of  electrical  current  was  created  by  a 
high-energy  discharge  from  a  powerful  capacitor  bank.  The  current  sheet 
traveled  down  a  ring-shaped  (annular)  channel  designed  around  a  central 
anode  (positive  electrode)  and  collapsed  by  virtue  of  its  own  self-induced 
magnetic  field  into  a  high-density  plasma. 

Several  researchers  in  Wood's  MHD  Section  became  deeply  involved  in 
experiments  with  the  plasma-focus  facility,  and  although  their  work  did  not 
produce  the  boundless  energy  of  nuclear  fusion,  it  cannot  be  called  a  failure; 
rather,  the  effort,  which  was  extensive,  turned  out  to  be  important  and 
lasting.  Between  1968  and  1985,  Langley  researchers  published  no  less  than 
81  papers  based  on  their  experiments  in  the  plasma- focus  facility;  only  7  of 
these  papers  were  written  between  1968  and  1970,  when  the  MPD  Branch 
was  still  functioning.  Clearly,  the  research  did  not  end  with  the  formal 
dissolution  of  the  branch.  In  this  collection  of  papers  authored  or  coauthored 
by  the  members  of  the  former  MPD  Branch  are  significant  offshoots  from 
the  initial  purpose  of  the  experiments.  These  offshoots  include  exploration 
of  space-based  lasers  both  for  direct  conversion  of  solar  energy  and  for 
early  "Star  Wars"  designs.  In  the  late  1970s,  the  plasma-focus  facility 
received  national  and  international  attention  and  acclaim  by  producing  more 
neutrons  per  experimental  "shot" — 1019  fusion  neutrons  from  a  deuterium 
plasma — than  had  been  produced  by  any  other  fusion  experiment  to  date 
in  the  United  States.  By  placing  enriched  uranium  at  the  end  of  the  anode, 
researchers  were  even  able  to  get  1010  fissions,  which  was  another  remarkable 
result.50 

These  achievements  signified  that  Langley's  general  fusion-related  re- 
search rated  near  the  top  of  the  American  scientific  effort  by  the  early  1980s. 
Langley's  work  was  equal  to  similar  pioneering  efforts  by  Winston  H.  Bostick 
at  the  Stevens  Institute  of  Technology  in  New  Jersey  and  G.  R.  Mather  at 
Los  Alamos  National  Laboratory  in  New  Mexico.  Of  course,  the  chronol- 
ogy for  this  work  extends  beyond  the  period  that  is  the  focus  of  this  book; 
however,  the  relevance  of  the  research  carried  on  by  the  MPD  Branch  of  the 
1960s  extended  to  these  significant  follow-on  efforts. 


5k 

A  much  earlier  piece  of  equipment  for  plasma  research  at  Langley  known  as  "the  diffusion  inhibitor" 
was  developed  to  pursue  thermonuclear  power.  In  1938,  Langley  researchers  Eastman  N.  Jacobs 
and  Arthur  Kantrowitz  tried  to  confine  a  hot  plasma  magnetically  and  thereby  achieve  a  controlled 
thermonuclear  reaction.  Although  NACA  management  quickly  stopped  the  unauthorized  research,  the 
preliminary  experiments  attempted  by  Jacobs  and  Kantrowitz  in  their  toroidal  (or  doughnut-shaped) 
chamber  represent  not  only  Langley's  first  flirtation  with  the  basic  science  later  leading  to  MPD  studies 
but  also  the  first  serious  effort  anywhere  in  the  world  .(and  three  years  before  the  Manhattan  Project) 
to  obtain  energy  from  the  atom.  For  a  complete  account  of  the  Jacobs-Kantrowitz  fusion  experiment  of 
1938,  see  James  R.  Hansen,  "Secretly  Going  Nuclear,"  in  American  Heritage  of  Invention  &  Technology 
(Spring  1992)  7:60-63. 


149 


Spaceflight  Revolution 

A  Hot  Field  Cools  Off 

Although  the  promise  of  MPD  remained  high  into  the  late  1960s,  its 
mystique  was  slowly  dissipating.  In  a  briefing  to  new  Langley  Director 
Edgar  M.  Cortright  in  1968,  Mike  Ellis  had  to  admit  that  "a  large  part  of 
the  glamour  of  moving  into  plasma  physics  that  existed  ten  years  ago  is  now 
over  and  we  feel  that  hard-headed  research  is  now  the  order  of  the  day."51 

Ten  years  had  passed,  and  the  ambitions  of  the  first  exhilarating  mo- 
ments of  the  spaceflight  revolution  had  been  moderated  by  the  mounting 
frustrations  of  trying  to  achieve  significant  research  results  in  what  was 
proving  to  be  a  much  more  illusive  area  of  research  than  anticipated.  "The 
field  was  just  so  incredibly  complicated,"  Mike  Ellis  remembers,  "that  to 
make  a  really  significant  contribution  that  would  apply  to  some  great  prob- 
lem just  became  increasingly  hard."52  The  deeper  the  Langley  researchers 
and  others  plunged  into  the  MPD  field,  the  more  they  realized  how  difficult 
contributing  to  any  applications  would  be. 

Because  they  could  not  find  clear  applications  for  most  of  their  research, 
the  sights  of  the  MPD  enthusiasts  changed  gradually  over  the  course  of  the 
1960s.  In  terms  of  simulating  the  reentry  conditions,  which  was  the  practical 
application  driving  so  much  of  the  MPD  effort  in  its  early  years,  neither 
Langley's  arc-jets,  nor  its  plasma  accelerators,  nor  any  other  new  facility 
ever  succeeded  in  generating  on  the  ground  a  flow  of  high-temperature 
air  that  corresponded  to  actual  flight  conditions.  And,  by  the  late  1960s, 
NASA  knew  that  a  spaceflight  program  could  do  well  without  having  that 
capability.  As  John  Becker  of  Aero-Physics  explains, 

We  learned  everything  that  we  could  in  an  airstream  that  was  way  too  cool,  and  then 
we  corrected  wherever  we  needed  to  for  the  effects  of  the  temperature,  by  calculation, 
by  studying  the  effects  of  temperature  in  adequate  facilities,  and  then  adding  that 
to  what  we  already  knew.  It  was  a  partial  simulation,  but  the  corrections  were  good 
enough  to  design  successful  hardware. 

In  other  words,  much  of  what  MPD  researchers  had  been  trying  to  do  just 
proved  unnecessary. 

The  primary  motivation  for  many  who  had  joined  the  MPD  field  had 
been  the  hope  of  controlled  thermonuclear  fusion.  Anybody  and  everybody 
in  the  scientific  community  who  was  connected  to  plasma  physics  had  the 
dream  of  inventing  the  final  device  that  would  allow  controlled  fusion,  or 
at  least  they  hoped  to  contribute  in  some  direct  way  to  its  eventual  design. 
But  by  the  late  1960s,  the  lack  of  progress  in  the  field  clearly  indicated  that 
any  practical  technology  based  on  fusion  (other  than  an  atomic  bomb  or 
nuclear  warhead)  was  still  a  long  way  off.54 

At  the  dawn  of  the  space  age,  NASA's  MPD  enthusiasts  at  Langley  and 
elsewhere  had  also  believed  that  nuclear-powered  rockets,  ion  rockets,  and 
other  advanced  space  propulsion  systems  might  be  just  around  the  corner 

150 


The  "Mad  Scientists"  of  MPD 


This  May  1965  photograph  shows  a 
Langley  concept  for  a  Mars  landing  ve- 
hicle. By  the  end  of  the  decade,  all 
thoughts  of  making  a  quick  trip  to  Mars 
ended. 


L-65-3960 


and  that  with  them  astronauts  would  soon  be  shooting  off  for  Mars  and  other 
faraway  places.  As  the  decade  passed,  the  idea  of  the  nuclear  rocket  fell  by 
the  wayside  and  was  for  all  practical  purposes  killed  when  NASA  planning 
for  a  manned  Mars  mission  was  put  to  an  abrupt  halt  in  1970  by  President 
Nixon.  The  value  of  exploring  the  potential  of  electric  propulsion  systems 
also  diminished.  Mike  Ellis  remembers  the  impact  of  the  presidential  policy 
on  his  own  work:  "In  early  1970,  I  was  told  to  cease  working  immediately 
on  a  paper  I  was  preparing  for  formal  presentation  on  the  proposed  manned 
mission  to  Mars.  The  paper,  on  which  I  was  working  with  Walter  B.  Olstad 
and  E.  Brian  Pritchard  in  the  Aero-Physics  Division,  was  all  ready  for 
rehearsal.  But  then  word  came  down  from  Washington,  and  I  was  told 
not  even  to  breathe  the  notion  of  a  manned  Mars  mission."55 

As  their  lofty  aspirations  were  forced  down  to  earth,  the  MPD  enthusiasts 
shifted  their  focus  and  began  to  look  for  other  objectives.  A  group  in  the 
Plasma  Physics  Section,  for  example,  started  to  explore  the  potential  of  gas 
lasers.  Under  the  direction  of  Bob  Hess  and  his  associate  Frank  Allario, 
this  new  area  of  interest  grew  into  a  sustained  field  of  intense  research  at 
NASA  Langley.  By  the  early  1970s,  this  effort  provided  some  information 
basic  to  the  eventual  development  of  the  plasma  cutting  torches  and  plasma 
metal-definition  apparatuses  that  have  since  come  to  dominate  the  metals 
field.56 


151 


Spaceflight  Revolution 

In  1970,  Edgar  Cortright  as  part  of  his  major  reorganization  of  the  center 
dissolved  the  MPD  Branch  and  put  most  of  its  people  and  many  of  their 
facilities  under  a  new  Space  Sciences  Division  headed  by  William  H.  Michael. 
Aware  of  MPD's  practical  limitations,  Mike  Ellis  did  not  complain  about 
his  branch's  dissolution,  nor  did  any  other  member  of  his  staff.*  Cortright 
was  somewhat  familiar  with  the  MPD  field  from  his  days  as  a  researcher 
at  Lewis  laboratory  and  from  his  management  experience  in  the  OSSA  at 
NASA  headquarters,  so  he  did  not  criticize  MPD's  work  or  refer  to  it  in 
any  way  as  a  failure.  John  Becker,  who  had  supported  his  MPD  Branch  for 
nearly  a  decade,  best  sums  up  Langley's  view  of  MPD:  It  was  "a  field  that 
we  had  to  explore  in  detail  because  of  the  great  promise.  The  fact  that  it 
didn't  yield  any  earth-shaking  new  things  is  not  our  fault.  It's  just  the  way 
nature  turned  out  to  be."57 

Never  before  in  the  history  of  applied  basic  research  at  Langley  had  a 
field  of  study  promised  so  much,  yet  delivered  so  little.  But  the  "mad  scien- 
tists" of  MPD  were  not  mad  in  their  pursuit;  they  were  just  different  from 
the  "normal"  body  of  researchers  at  Langley,  who  searched  for  practical 
solutions  and  did  not  stray  into  matters  of  fundamental  cosmological  im- 
portance. The  MPD  group's  commitment  to  basic  scientific  research  was  in 
fact  quite  sensible.  At  a  time  when  NASA  had  an  increasingly  strong  politi- 
cal mandate  for  research  that  was  "relevant"  to  the  technological  objectives 
of  space  projects,  the  "mad  scientists"  of  MPD  maintained  a  broader  and 
more  fundamental  interpretation  of  relevant  research. 

Mike  Ellis  would  always  feel  that  MPD's  interpretation  was  the  proper 
one  and  that  the  urgency  of  project  work  had  deteriorated  the  status  of 
basic  research  at  Langley.  Project  work  so  dominated  the  agency  in  the 
late  1960s  that  all  work,  even  basic  research  such  as  that  conducted  by  the 
MPD  Branch,  was  judged  by  the  black-and-white  criteria  for  project  success. 
Results  must  be  quickly  achieved  and  immediately  applicable.  The  results 
of  Langley's  MPD  work  were  neither.  Mike  Ellis  puts  the  experience  in 
perspective:  "It  is  certainly  true  that  we  didn't  produce  any  earth-shaking 
results  or  great  breakthroughs.  Not  many  efforts  do."58 


ill 

Ellis  himself,  however,  did  not  move  into  the  new  Space  Sciences  Division;  instead,  he  became  one 
of  the  assistant  chiefs  (and  later  associate  chief)  of  John  Becker's  Aero-Physics  Division.  Paul  Huber, 
head  of  the  Plasma  Applications  Section,  became  head  of  Aero-Physic's  Propulsion  Research  Branch, 
which  worked  on  hypersonic  scramjets. 

152 


6 


The  Odyssey  of  Project  Echo 


The  vitality  of  thought  is  an  adventure.  Ideas  won't 
keep.  Something  must  be  done  about  them.  When  the 
idea  is  new,  its  custodians  have  a  fervor.  They  live 
for  it. 

—Dialogues  of  Alfred  North  Whitehead 

For  the  things  we  have  to  learn  before  we  can  do  them, 
we  learn  only  by  doing  them. 

— Aristotle 


In  the  early  hours  of  28  October  1959,  five  days  after  the  close  of  the 
first  NASA  inspection,  people  up  and  down  the  Atlantic  coast  witnessed  a 
brilliant  show  of  little  lights  flashing  in  the  sky.  This  strange  display,  not 
unlike  that  of  distant  fireworks,  lasted  for  about  10  minutes.  From  New 
England  to  South  Carolina,  reports  of  extraordinary  sightings  came  pouring 
into  police  and  fire  departments,  newspaper  offices,  and  television  and  radio 
stations.  What  were  those  mysterious  specks  of  light  flashing  overhead? 
Was  it  a  meteor  shower?  More  Sputniks?  UFOs?  Something  NASA  finally 
managed  to  launch  into  space? 

Several  hours  later,  the  press  was  still  trying  to  solve  the  mystery.  At 
about  three  o'clock  in  the  morning,  a  night  watchman  roused  NASA  Langley 
rocket  engineer  Norman  L.  Crabill  from  a  sound  sleep  in  a  dormitory  near 
the  launchpads  on  Wallops  Island.  The  watchman  told  Crabill  that  a  long- 
distance telephone  call  was  waiting  for  him  in  the  main  office.  A  reporter 
for  a  New  York  City  newspaper  wanted  a  statement  about,  as  he  put  it, 
"the  lights  that  you  guys  had  put  up."  Crabill,  an  irascible  young  member 
of  Langley 's  PARD,  had  not  been  able  to  celebrate  his  thirty-third  birthday 
properly  the  night  before  because  of  what  had  happened,  and  now  he  had 
gotten  out  of  a  warm  bed,  put  on  his  pants,  and  taken  a  walk  in  the  cool  night 

153 


Spaceflight  Revolution 

air  just  to  explain  the  situation  to  some  newspaper  guy.  "My  statement  is, 
'It's  three  o'clock  in  the  morning,'  "  growled  Crabill,  slamming  the  receiver 
down.  As  he  would  later  remember,  "It  was  the  only  time  I,  a  government 
employee,  ever  told  off  the  press  and  got  away  with  it."1 

Given  the  events  of  that  evening,  Crabill  s  anger  was  understandable. 
Although  the  disaster  that  had  occurred  was  minor,  it  was  big  enough  to 
potentially  damage  Crabill's  NASA  career.  The  initial  test  of  a  110- foot- 
diameter  inflatable  sphere  for  the  Echo  1  Passive  Communication  Satellite 
Project  had  ended  abruptly  with  the  sphere  blowing  up  as  it  inflated. 
Floating  back  into  the  atmosphere,  the  thousands  of  fragments  of  the 
aluminum-covered  balloon  had  reflected  the  light  of  the  setting  sun,  thus 
creating  the  sensational  flashing  lights. 

The  inflatable  sphere  had  been  launched  from  Wallops  Island  at  5:40 
p.m.  For  the  first  few  minutes,  everything  went  well.  The  weather  was 
fine  for  the  launch,  and  the  winds  were  not  too  high.  PARD  engineers 
were  worried  about  the  booster  called  "Shotput,"  an  experimental  two-stage 
Sergeant  X248  rocket,  because  the  performance  of  the  rocket's  second-stage 
Delta  was  to  be  the  initial  test  of  the  U.S.  Thor-Delta  satellite  launching 
system.  However,  in  the  early  moments  of  its  test  flight,  Shotput  1  had 
performed  flawlessly.  The  rocket  took  the  26-inch-diameter,  spherical, 
190-pound  payload  canister — inside  of  which  the  uninflated  130-pound 
aluminum-coated  Mylar-plastic  satellite  had  been  neatly  folded — to  second- 
stage  burnout  at  about  60  miles  above  the  ocean.  There,  the  payload 
separated  successfully  from  the  booster,  the  canister  opened,  and  the  balloon 
started  to  inflate.  The  first  step  in  Project  Echo  had  been  taken  with 
apparent  success. 

Then,  unexpectedly,  the  inflating  balloon  exploded.  The  payload  engi- 
neers had  left  residual  air  inside  the  folds  of  the  balloon  by  design  as  an 
inflation  agent.  The  air  expanded  so  rapidly,  because  of  the  zero  pressure 
outside,  that  it  ruptured  the  balloon's  thin  metallized  plastic  skin,  ripping 
the  balloon  to  shreds.  Shotput  1  was  history;  the  use  of  residual  air  to  help 
blow  up  the  balloon  had  been,  in  Crabill's  words,  a  "bad  mistake."2 

After  spending  a  depressing  night  reviewing  why  the  test  went  wrong, 
the  only  thing  for  Crabill  to  do  the  next  morning  was  to  get  to  work  solving 
the  problem.  After  all,  this  was  project  work — the  ultimate  reality — not 
general  research.  No  time  to  cry  over  spilled  milk — or  burst  balloons. 

At  the  NASA  press  briefing  at  Wallops,  held  about  one  hour  after  the 
explosion,  Crabill  and  others  had  given  their  usual  matter-of-fact  postlaunch 
systems  report.  In  the  midst  of  taking  a  quick  look  at  the  telemetry  records 
to  make  sense  of  the  balloon  failure,  a  NASA  official  sensitive  to  public  affairs 
approached  Crabill  and  told  him,  "Just  tell  them  everything  worked  all 
right."3  Sure,  Crabill  thought,  no  problem.  No  data  pointed  to  the  contrary. 
All  the  visual  evidence  on  the  Shotput  launch  vehicle,  which  was  Crabill's 
responsibility,  suggested  that  Shotput  had  worked  as  planned.  Moreover,  the 


154 


The  Odyssey  of  Project  Echo 


L-93-8337  L-93-8339  L-93-8333 

During  a  test  of  the  Echo  deployment  in  1962,  which  was  three  years  after  Shotput's 
first  failed  deployment  of  the  Echo  satelloon,  a  structural  load  problem  caused  the 
balloon  once  again  to  explode.  A  camera  aboard  the  launcher  captured  these  images. 
The  earlier  Shotput  failure  would  have  looked  very  much  the  same. 

purpose  of  the  Shotput  phase  of  Project  Echo*  was  to  determine  whether  the 
mechanism  designed  to  deploy  an  inflatable  passive  communications  satellite 
of  that  size  and  weight  would  work,  and  it  had;  in  that  sense,  Shotput  1  was 
indeed  successful.  To  tell  the  whole  truth  about  that  scintillating  collection 
of  little  moving  lights  tumbling  through  the  upper  atmosphere  before  all  the 
records  were  examined  and  understood  was  premature — and  the  complete 
story  would  be  too  complicated  for  the  press  to  understand.  This  was  a  time 
when  launching  any  object  into  space  was  big  news  for  the  American  people. 
Why  let  an  otherwise  uplifting  moment  be  turned  into  another  letdown? 

Thus,  for  the  initial  newspaper  stories  about  the  launch  of  Shotput  1,  the 
press  would  not  be  told  enough  even  to  hint  at  the  possibility  of  a  failure.  For 
example,  in  a  front-page  article  appearing  in  the  next  morning's  Newport 
News  Daily  Press,  the  headline  for  military  editor  Howard  Gibbons'  article 
about  the  launch  was:  "Eart tilings  Stirred  by  NASA  Balloon,  Awesome 
Sight  in  the  Sky."  According  to  Gibbons,  NASA  had  launched  "the  largest 
object  ever  dispatched  into  space  by  man,  stirring  the  curiosity  and  awe  of 
thousands  of  Americans  residing  on  the  Eastern  Seaboard."  The  inflated 
sphere  "rode  for  13  minutes  in  the  sun's  rays  . . .  before  it  fell  again  into 
the  atmosphere  and  dropped  into  the  Atlantic  about  500  miles  east  of 
Wallops."  Gibbons  made  no  mention  of  the  rupture.  The  balloon  "was 
probably  deflated  on  the  way  down  into  the  atmosphere,  NASA  reported." 
Not  a  word  appeared  about  a  mistake  involving  the  use  of  residual  air  as  an 
inflation  agent.  According  to  Gibbons,  "NASA's  assessment  of  the  operation 
was  that  'it  did  what  we  wanted  it  to  do.'  "4 

Also  on  the  front  page  of  the  Daily  Press  that  morning,  next  to  the 
article  on  Shotput  1,  was  an  Associated  Press  wire  story  from  Washington, 
D.C.,  announcing  the  start  of  extensive  congressional  hearings.  The  House 


The  word  "echo"  was  already  in  use  by  the  late  1950s  to  describe  a  pulse  of  reflected  radio-frequency 
energy. 

155 


Spaceflight  Revolution 

Space  Committee  was  investigating  why  the  United  States  continued  "to 
play  second  fiddle"  to  the  Soviets  in  the  exploration  of  space;  the  headline 
of  this  second  article  read:  "Why  U.S.  Lagging  In  Space  Explorations  To 
Be  Probed."  The  last  thing  NASA  needed  at  the  moment  was  to  explain  a 
burst  balloon. 

Norm  Crabill  knew  that  NASA  was  not  telling  the  press  the  truth,  but 
he  and  the  rest  of  the  Langley  crew  responsible  for  the  shot  understood 
and  accepted  the  subterfuge.  This  was  project  work,  and  it  had  to  succeed. 
For  public  consumption,  both  failure  and  the  inability  to  achieve  complete 
success  need  not  be  admitted,  at  least  not  immediately.  Sometimes  mistakes 
could  not  be  concealed,  such  as  a  missile  blowing  up  on  the  launchpad  before 
hundreds  of  cameras,  as  so  many  had  been  doing.  But  a  balloon  bursting 
in  space,  especially  one  producing  such  a  sensational  show  of  flashing 
lights,  could  be  presented  as  a  total  success.  This  age  of  the  spaceflight 
revolution  was  a  new  epoch.  Research  activities  were  now  exposed  to  the 
nontechnical  general  public,  and  so  many  old  rules  and  definitions  no  longer 
were  applicable.  Some  discretion  in  the  discussion  of  results  seemed  justified. 


The  International  Geophysical  Year  and  the  V-2  Panel 

As  with  so  many  early  NASA  projects  and  programs,  Project  Echo 
originated  in  NAG  A  work.  In  fact,  the  idea  predated  the  Sputnik  crisis 
by  several  months  and  at  first  had  nothing  to  do  with  proving  the  feasibility 
of  a  global  telecommunications  system  based  on  the  deployment  of  artificial 
satellites.  Rather,  the  original  purpose  of  Echo  was  to  measure  the  density 
of  the  air  in  the  upper  atmosphere  and  thereby  provide  aerodynamic 
information  helpful  in  the  design  of  future  aircraft,  missiles,  and  spacecraft. 
Like  so  many  other  matters  affected  by  the  spaceflight  revolution,  the 
concept  that  led  to  Project  Echo  had  modest  and  circumscribed  beginnings 
that  ballooned  into  sensational  results. 

The  father  of  the  Echo  balloon  was  Langley  aeronautical  engineer 
William  J.  O'Sullivan,  who  was  a  1937  graduate  of  the  University  of  Notre 
Dame  (and  Langley  employee  since  1938)  and  a  former  staff  member  of 
PARD.  The  idea  for  the  air-density  experiment  first  came  to  O'Sullivan  on 
26  January  1956,  nearly  two  years  before  the  launch  of  Sputnik  1.  All  that 
raw  winter  day,  the  40-year-old  O'Sullivan  sat  in  a  meeting  of  the  Upper 
Atmosphere  Rocket  Research  Panel,  which  was  being  held  at  the  Univer- 
sity of  Michigan  in  Ann  Arbor.  Originally  known  as  the  "V-2  Panel,"  this 
body  had  been  formed  in  February  1946  to  help  the  army  select  the  most 
worthwhile  experiments  to  be  carried  aboard  the  captured  and  rebuilt  Ger- 
man V-2  rockets.*  After  World  War  II,  scientists  from  around  the  country 


The  V-2  rockets  were  originally  known  as  A-4s.  To  avoid  association  with  the  German  "Vengeance 
Weapons"  that  had  terrorized  England,  the  U.S.  military  often  referred  to  them  by  their  original  name. 

156 


The  Odyssey  of  Project  Echo 


L-61-7786 

William  J.  O  'Sullivan,  the  father  of  the  Echo  balloon,  was  also  the  father  of  five 
children.  They,  too,  were  caught  up  in  the  enthusiasms  of  the  spaceflight  revolution. 
Notice  the  homemade  NASA  emblems  on  the  blazers  worn  by  the  two  teenage  sons. 
The  NASA  public  affairs  office  distributed  copies  of  this  family  portrait  to  the  news 
media  along  with  stories  about  O 'Sullivan's  ingenious  invention  of  the  Echo  balloon. 

had  flooded  the  army  with  requests  for  an  allotment  of  space  aboard  the 
V-2s,  and  the  army  had  handled  the  awkward  situation  rather  adroitly  by 
instructing  the  scientists  to  form  a  panel  of  their  own  to  decide  which  exper- 
iments should  go  on  the  rockets.  Thus,  the  V-2  Panel  came  to  life  as  a  free 
and  independent  body,  with  no  authority  to  enforce  its  decisions,  but  with 
a  voice  that  carried  the  weight  of  the  scientific  community  behind  it.  By 
the  early  1950s,  the  name  of  the  panel  changed  to  the  Upper  Atmosphere 
Rocket  Research  Panel,  signifying  both  a  wider  agenda  of  research  concerns 
and  the  use  of  rockets  other  than  the  V-2s  as  flight  vehicles.5 

The  purpose  of  the  Ann  Arbor  meeting  was  to  choose  the  space  ex- 
periments for  the  forthcoming  International  Geophysical  Year  (IGY).  This 
event,  to  be  celebrated  by  scientists  around  the  world  beginning  1  July  1957, 
stimulated  many  proposals  for  experiments,  including  the  stated  ambition  of 
both  the  American  and  Soviet  governments  to  place  the  first  artificial  satel- 
lites in  orbit  about  the  earth.  The  panel's  job  was  to  sort  these  proposals  into 
two  groups:  those  that  could  most  satisfactorily  be  conducted  with  sounding 
rockets  and  those  that  could  be  performed  aboard  "Vanguard,"  the  proposed 

157 


Space/light  Revolution 

National  Academy  of  Sciences/U.S.  Navy  earth  satellite.  Then,  after  hearing 
20-minute  oral  presentations  in  support  of  each  proposal,  the  panel,  chaired 
by  University  of  Iowa  physicist  James  Van  Allen,  was  to  choose  the  most 
deserving  experiments. 

As  the  NACA  representative  to  this  panel,  O'Sullivan  sat  through  the 
day-long  meeting  and  grew  increasingly  frustrated  with  what  he  was  hearing. 
He  was  particularly  disappointed  by  the  methods  proposed  to  measure 
the  density  of  the  upper  atmosphere.  As  an  aeronautical  engineer,  he 
understood  that  information  about  air  density  might  prove  vital  to  the 
design  of  satellites,  ICBMs,  and  every  aerospace  vehicle  to  fly  in  and  around 
the  fringes  of  the  earth's  atmosphere.  In  1952  and  1953,  O'Sullivan  had 
belonged  to  a  three-man  study  group  supported  by  Langley  management  for 
the  purpose  of  exploring  concepts  for  high-altitude  hypersonic  flight.  With 
fellow  Langley  researchers  Clinton  E.  Brown  and  Charles  H.  Zimmerman, 
O'Sullivan  had  educated  himself  in  the  science  of  hypersonics  and  helped 
the  group  to  conceptualize  a  manned  research  airplane  that  could  fly  to 
the  limits  of  the  atmosphere,  be  boosted  by  rockets  into  space,  and  return 
to  earth  under  aerodynamic  control.  In  essence,  the  Brown- Zimmerman- 
O' Sullivan  study  group  had  envisioned  a  "space  plane"  very  similar  to  the 
future  North  American  X-15  and  its  related  heir,  the  Space  Shuttle.7 

Given  his  enthusiasm  for  spaceflight,  O'Sullivan  was  disappointed  to  hear 
respected  scientists  offering  such  defective  plans  for  obtaining  the  critical 
air-density  data.  One  proposal,  developed  by  a  group  at  the  University  of 
Michigan,  involved  the  use  of  a  special  omnidirectional  accelerometer  whose 
sensitivity,  according  to  O'Sullivan,  would  have  to  be  "improved  by  between 
100  and  1000  times  before  the  experiment  would  work."  Another  proposal, 
one  of  two  submitted  by  Princeton  University  physicist  Lyman  Spitzer,  Jr., 
called  for  the  measurement  of  drag  forces  on  a  satellite  spiraling  its  way  back 
to  earth.  In  O'Sullivan's  view,  the  principle  behind  Spitzer's  proposal  was 
sound,  but  not  practical.  The  experiment  would  work,  he  predicted,  "only  at 
altitudes  much  below  that  at  which  practical  satellites  of  the  future  would 
have  to  fly  in  order  to  stay  in  orbit  long  enough  to  be  worth  launching, 
probably  at  least  five  years."8  Several  of  the  day's  proposals,  including  the 
ones  just  mentioned,  were  based  on  the  presumption  that  somebody  could 
build  and  launch  a  lightweight  structure  strong  enough  to  remain  intact 
during  its  turbulent  ballistic  shot  into  space.  But  as  far  as  O'Sullivan  knew, 
nobody  had  yet  discovered  how  to  do  this.  At  that  moment,  still  more  than 
a  year  and  a  half  before  Sputnik,  no  one  had  yet  succeeded  in  launching 
even  a  simple  grapefruit-sized  object  into  space,  let  alone  objects  as  big 
and  complicated  as  those  being  suggested  by  the  scientists  that  day  in  Ann 
Arbor. 

In  his  hotel  room  that  evening,  O'Sullivan  could  not  let  go  of  the  problem. 
None  of  the  proposals  he  had  heard  were  satisfactory.  But  were  there  better 
alternatives?  Even  if  he  could  think  of  one  himself,  technically,  as  a  panel 
member,  he  was  supposed  to  judge  the  suggestions  formally  submitted  by 

158 


The  Odyssey  of  Project  Echo 

others,  not  make  any  of  his  own.  With  a  pad  of  paper  from  the  hotel  desk  in 
hand,  however,  he  could  not  resist  making  some  rough  calculations.  Over  the 
next  several  hours,  O'Sullivan  was  engaged  in  a  process  of  creative  problem 
solving,  which  he  would  later  outline  in  seven  major  points  of  analysis.9 


O'Sullivan's  Design 

(1)  Aero  theory.  O'Sullivan  naturally  started  his  analysis  from  the  point 
of  view  of  aerodynamic  theory.   He  knew  from  theory  that  "the  drag  force 
experienced  by  a  satellite  in  the  outer  extremities  of  the  earth's  atmosphere 
was  directly  proportional  to  the  density  of  the  atmosphere."    This  meant 
that  "if  the  drag  could  be  measured,  the  air  density  could  be  found."  Thus, 
in  the  first  few  minutes  of  his  analysis,  O'Sullivan  had  reduced  the  entire 
problem  to  measuring  the  satellite  drag.10 

(2)  Shape  and  size.    How  big  should  the  satellite  be  and  what  shape? 
O'Sullivan  chose  a  sphere.    Such  a  fixed  shape  eliminated  the  problem  of 
the  satellite's  frontal  area  relative  to  the  direction  of  motion,  and  it  also 
simplified  the  question  of  the  satellite's  size. 

(3)  Drag  forces.    O'Sullivan  turned  next  to  a  consideration  of  celestial 
mechanics.  On  his  pad  of  paper,  he  began  to  play  with  the  classical  equations 
for  the  drag  forces  on  a  body  as  it  moves  on  a  ballistic  trajectory  through 
the  atmosphere  and  into  orbit  around  the  earth.     After  several  minutes 
of  mathematical  work,  during  which  he  constantly  reminded  himself  that 
the  success  of  the  experiment  depended  on  making  the  satellite  extremely 
sensitive  to  aerodynamic  drag,  O'Sullivan  realized  that  he  must  devise  a 
satellite  of  exceedingly  low  mass  relative  to  its  frontal  area.    The  satellite 
could  be  a  large  sphere,  but  the  sphere's  material  could  not  be  so  dense  as 
to  make  the  satellite  insensitive  to  the  very  air  drag  it  was  to  detect.  Only 
an  object  with  a  low  mass-to-frontal  area  ratio  could  be  pushed  around  by 
an  infinitesimal  amount  of  air. 

(4)  Design  considerations.  No  researcher  who  worked  at  such  a  diversified 
place  of  technical   competence   as   Langley,    which  was  the  only  NACA 
aerodynamics  laboratory  with  a  Structures  Research  Division,  could  long 
proceed  with  the  analysis  of  a  flight-vehicle  design  without  considering 
matters  of  weight,  loads,  elasticity,  and  overall  structural  integrity.    The 
structural  problems  of  O'Sullivan's  sphere  might  prove  serious,   for  the 
lighter  the  weight  (or  lesser  the  mass),  the  weaker  the  structure.     With 
this  conventional  knowledge  about  structural  strength  in  mind,  O'Sullivan 
contemplated  the  magnitude  of  the  loads  that  his  satellite  structure  would 
have  to  withstand.   Calculations  showed  that  the  loads  on  his  sphere,  once 
in  space,  would  be  quite  small,  amounting  to  perhaps  only  one-hundredth  to 
one-thousandth  the  weight  that  the  sphere  would  encounter  at  rest  on  the 
surface  of  the  earth.  From  this,  he  concluded  that  the  satellite  need  only  be 
a  thin  shell,  as  thin  perhaps  as  ordinary  aluminum  foil. 

159 


Space/light  Revolution 

But  herein  was  the  dilemma.  In  orbit,  the  sphere  would  encounter 
negligible  loads  and  stresses  on  its  structure,  but  to  reach  space,  it  would 
have  to  survive  a  thunderous  blast-off  and  lightning-like  acceleration  through 
dense,  rough  air.  O'Sullivan  knew  that  he  could  not  design  a  satellite  for  the 
space  environment  alone;  rather,  a  structure  must  be  designed  to  "withstand 
the  greatest  loads  it  will  be  exposed  to  throughout  its  useful  life."  The 
satellite  would  have  to  withstand  an  acceleration  possibly  as  high  as  10  Gs, 
which  was  1000  to  10,000  times  the  load  the  structure  would  be  exposed  to 
in  orbit.  To  survive,  the  satellite  could  not  consist  merely  of  a  thin  shell; 
it  would  have  to  be  so  strong  and  have  such  a  high  mass-to-area  ratio  that 
it  would  be  insensitive  to  minute  air  drag  and  thereby  "defeat  the  very 
objective  of  its  existence." J 

Midnight  was  approaching,  and  O'Sullivan,  the  scientific  wizard  of 
Langley's  PARD,  still  sat  at  the  hotel  desk,  perched  on  the  horns  of  this 
dilemma.  Finally,  in  the  early  hours  of  the  morning,  he  arrived  at  a 
possible  solution:  why  not  build  the  sphere  out  of  a  thin  material  that 
could  be  folded  into  a  small  nose  cone?  If  the  sphere  could  be  packed 
snugly  into  a  strong  container,  it  could  easily  withstand  the  acceleration 
loads  of  takeoff  and  come  through  the  extreme  heating  unscathed.  After 
the  pay  load  container  reached  orbit,  the  folded  satellite  could  be  unfolded 
and  inflated  pneumatically  into  shape.  Finding  a  means  of  inflation  should 
not  be  difficult.  Either  a  small  tank  of  compressed  gas  such  as  nitrogen,  or 
a  liquid  that  would  readily  evaporate  into  a  gas,  or  even  some  solid  material 
that  would  evaporate  to  form  a  gas  (such  as  the  material  used  to  make 
mothballs)  could  be  used  to  accomplish  the  inflation.  (He  apparently  had 
not  yet  thought  of  using  residual  air  as  the  inflation  agent,  as  in  Shotput  1.) 
Almost  no  air  pressure  existed  at  orbiting  altitude,  so  a  small  amount  of  gas 
would  do  the  job.  "Clearly  then,"  O'Sullivan  concluded,  "that  is  how  the 
satellite  had  to  work." 12 

(5)  Construction  materials.  Other  critical  questions  still  needed  answers. 
Surely,  if  he  presented  his  notion  of  an  inflatable  satellite  to  the  prestigious 
scientific  panel  the  next  day,  someone  would  ask  him  to  specify  its  con- 
struction material.  The  material  had  to  be  flexible  enough  to  be  folded, 
strong  enough  to  withstand  being  unfolded  and  inflated  to  shape,  and  stiff 
enough  to  keep  its  shape  even  if  punctured  by  micrometeoroids.  O'Sullivan 
reviewed  the  properties  of  the  materials  with  which  he  was  familiar  and 
quickly  realized  that  "no  one  of  them  satisfied  all  the  requirements."  Next 
he  tried  combining  materials.  The  forming  of  thin  sheet  metal  into  certain 
desired  shapes  was  a  standard  procedure  in  many  manufacturing  industries, 
but  sheet  metal  thin  enough  for  the  skin  of  his  satellite  would  tear  easily 
during  the  folding  and  unfolding.  Perhaps,  thought  O'Sullivan,  some  tough 
but  flexible  material,  something  like  a  plastic  film,  could  be  bonded  to  the 
metal  foil.13 

Here  was  another  critical  part  of  the  answer  to  O'Sullivan's  satellite 
design  problem:  a  sandwich  or  laminate  material  of  metal  foil  and  plastic 

160 


The  Odyssey  of  Project  Echo 

film.  "I  could  compactly  fold  a  satellite  made  of  such  a  material  so  that 
it  could  easily  withstand  being  transported  into  orbit,  and  once  in  orbit,  I 
could  easily  inflate  it  tautly,  stretching  the  wrinkles  out  of  it  and  forming  it 
into  a  sphere  whose  skin  would  be  stiff  enough  so  that  it  would  stay  spherical 
under  the  minute  aerodynamic  and  solar  pressure  loads  without  having  to 
retain  its  internal  gas  pressure."14  Such  a  thin-skin  satellite  would  be  so 
aerodynamically  sensitive  that  even  a  minute  amount  of  drag  would  cause 
a  noticeable  alteration  in  its  orbit.  Researchers  on  the  ground  could  track 
the  sphere,  measuring  where  and  when  it  was  being  pushed  even  slightly 
off  course,  and  thereby  compute  the  density  of  the  air  in  that  part  of  the 
atmosphere. 

(6)  Temperature  constraints.  Would  a  satellite  made  out  of  such  material 
grow  so  cold  while  in  the  earth's  shadow  that  the  plastic  film  would  embrittle 
and  break  apart?    O' Sullivan  reckoned  that  this  would  not  be  a  problem 
as  he  knew  of  several  plastic  films  that  could  withstand  extremely  low 
temperatures.     The  real  concern  was  heat.     Exposure  to  direct  sunlight 
might  melt  or  otherwise  injure  the  outer  film.    But  this,  too,  seemed  to 
have  a  remedy.  Rough  calculations  showed  that  high  temperatures  could  be 
controlled  by  doping  the  outside  of  the  satellite  with  a  heat-reflecting  paint. 
Some  heat-reflecting  metals  might  even  do  the  job  without  paint,  if  they 
could  be  made  into  a  metal  foil. 

(7)  Satellite  tracking.  One  problem  remained:  the  means  of  tracking  the 
satellite.    As  a  member  of  the  Upper  Atmosphere  Rocket  Research  Panel, 
O' Sullivan  was  familiar  with  current  tracking  techniques.  These  included  the 
radio  method  built  into  the  navy's  Minitrack  network  in  which  the  object 
to  be  tracked  carried  a  small  transmitter  or  radio  beacon.     This  system 
would  be  adopted  for  the  Vanguard  satellite  project.    Unfortunately,  the 
radio-tracking  method  would  not  work  for  O'Sullivan's  satellite  concept.  If 
a  radio  beacon  was  attached  to  his  sphere,  it  would  add  significantly  to  the 
structural  mass,  thereby  reducing  the  sphere's  sensitivity  to  air  drag.   The 
only  way  to  track  the  sphere,  it  seemed,  was  optically,  with  special  cameras 
or  telescopes. 

Tracking  a  satellite  optically  when  the  satellite  would  be  made  out  of 
something  as  bright  and  highly  reflective  as  polished  sheet  metal  would  not 
be  difficult;  however,  optical  tracking  limited  satellite  observation  to  the 
twilight  hours.  At  all  other  times,  reflections  from  the  satellite  would  not  be 
practical.  At  night  the  satellite  would  be  in  the  earth's  shadow  and  hence 
would  receive  no  sunlight  to  reflect  to  tracking  instruments  or  observers  on 
earth;  in  daylight,  the  satellite  would  reflect  light,  but  that  light  would  be 
obliterated  from  view  by  the  scattering  of  the  sun's  rays  in  the  earth's  dense 
lower  atmosphere. 

O' Sullivan  knew  that  tracking  a  satellite  at  all  times  of  the  day  and  at 
night  was  possible  only  by  radar.  As  a  member  of  Langley's  PARD,  he 
was  intimately  familiar  with  the  radar  tracking  of  rocket  models;  at  Wallops 
Island,  it  had  been  a  routine  and  daily  procedure  for  several  years.  As 

161 


Space/light  Revolution 

O'Sullivan  suspected,  the  problem  was  that  radars  powerful  enough  to  do 
the  job,  even  if  the  satellite  were  as  big  as  a  house,  would  not  be  available 
for  several  years  because  they  were  still  in  development. 


Extraterrestrial  Relays 

These  thoughts  about  radar  tracking  led  O'Sullivan  to  a  much  higher 
level  of  technological  speculation.  "As  I  thought  about  this  ability  to  reflect 
radio  and  radar  waves,  there  came  a  stream  of  thoughts  about  the  future 
possibilities  of  such  a  satellite  when  there  would  exist  . . .  rockets  capable 
of  launching  big  enough  satellites,  and  when  powerful  enough  radars  and 
radios  [would  exist]  to  be  able  to  use  such  satellites  for  radio  and  television 
communication  around  the  curvature  of  the  earth,  and  as  navigational  aids 
that  could  be  seen  by  ship  and  airplane  radars  night  and  day,  clear  weather 
or  cloudy:  satellites  that  some  day  might  take  the  place  of  the  stars  and  sun 
upon  which  navigators  have  depended  for  so  many  generations." 15 

On  many  occasions  in  the  history  of  modern  technology,  science  fiction 
has  blazed  the  way  to  an  understanding  of  real  possibilities  and  has  moti- 
vated scientists  and  engineers  to  seek  practical  results;  perhaps  O'Sullivan's 
concept  for  the  inflatable  satellite  was  one  of  those  occasions.  Several 
years  earlier,  in  October  1945,  the  British  science-fiction  writer  Arthur  C. 
Clarke  had  published  a  visionary  article  in  the  popular  British  radio  journal 
Wireless  World  suggestively  entitled  "Extraterrestrial  Relays."  In  the  arti- 
cle, Clarke  predicted  the  development  of  an  elaborate  telecommunications 
system  based  on  artificial  satellites  orbiting  the  earth.16 

The  key  to  such  a  system,  according  to  Clarke,  would  be  the  "geosyn- 
chronous" satellite.  Such  a  satellite,  launched  to  a  distance  of  roughly  22,000 
miles  high  in  an  equatorial  orbit,  would,  according  to  the  laws  of  celestial 
mechanics,  take  exactly  24  hours  to  complete  one  orbit,  thereby  staying 
fixed  indefinitely  over  the  same  spot  on  the  earth.  The  satellite  would  act 
as  an  invisible  television  tower,  which  could  maintain  line-of-sight  contact 
with  one-third  of  the  earth's  surface.  If  three  satellites  were  put  into  geosyn- 
chronous orbit  above  the  equator  and  made  to  communicate  with  one  an- 
other through  long-distance  "extraterrestrial  relays,"  as  Clarke  called  them, 
electronic  signals,  be  they  radio,  television,  or  telephone,  could  be  passed 
from  satellite  to  satellite  until  those  signals  made  their  way  around  the 
globe.  For  the  first  time  in  history,  people  all  over  the  world  would  be  able 
to  communicate  instantaneously. 

The  impact  of  a  global  communications  system  would  be  revolutionary, 
Clarke  was  sure.  "In  a  few  years  every  large  nation  will  be  able  to  establish 
(or  rent)  its  own  space-borne  radio  and  TV  transmitters,  able  to  broadcast 
really  high-quality  programs  to  the  entire  planet."  This  would  mean  "the 
end  of  all  distance  barriers  to  sound  and  vision  alike.  New  Yorkers  or 
Londoners  will  be  able  to  tune  in  to  Moscow  or  Peking  as  easily  as  to  their 

162 


The  Odyssey  of  Project  Echo 

local  station."  The  new  communications  technology  might  even  lead  to  a 
new  order  of  world  cooperation  and  peace.  "The  great  highway  of  the  ether 
will  be  thrown  open  to  the  whole  world,  and  all  men  will  become  neighbors 
whether  they  like  it  or  not."  Inevitably,  peoples  of  all  nations  will  become 
"citizens  of  the  world."17 

We  do  not  know  if  Langley's  William  J.  O'Sullivan  (who  died  from  cancer 
in  1971  at  age  56)  had  read  any  of  Arthur  Clarke's  writings  or  was  in  any 
other  way  acquainted  with  Clarke's  ideas  about  communications  satellites 
at  the  time  of  the  Ann  Arbor  meeting  in  1956.  Most  likely  O'Sullivan  knew 
something  of  them,  given  the  intellectual  proclivities  of  the  flight-minded 
community  in  which  he  worked  and  the  extent  to  which  some  of  the  bolder 
ideas  about  space  exploration  were  making  their  way  into  the  mainstream 
of  American  culture  during  the  early  1950s  through  books,  magazines,  and 
movies.  Some  of  the  ambitious  ideas  about  space,  including  the  scheme  for 
a  global  system  of  communications  satellites,  were  beginning  to  appear  in 
the  serious  technical  literature. 

Take  the  relevant  case  of  John  R.  Pierce,  the  visionary  American  elec- 
trical engineer  working  at  Bell  Telephone  Laboratories.  In  1952,  Pierce 
began  to  develop  his  own  ideas  for  a  communications  satellite  system  but, 
fearing  the  ridicule  of  his  colleagues,  decided  to  publish  his  ideas  under  a 
pseudonym.  They  appeared  in  a  popular  magazine,  Amazing  Science  Fic- 
tion. In  the  next  few  years,  however,  the  climate  of  opinion  changed;  the 
electronics  revolution,  as  well  as  the  notion  of  integrating  rockets,  transis- 
tors, computers,  and  solar  cells,  progressed  far  enough  to  make  a  serious 
technical  discussion  of  the  possibility  of  "comsats"  (communication  satel- 
lites) professionally  acceptable.  Pierce  then  came  out  of  the  closet;  in  April 
1955,  he  published  "Orbital  Radio  Relays"  under  his  own  name  in  the  re- 
spected trade  journal  Jet  Propulsion.1^ 

In  the  same  month,  another  provocative  article  by  Pierce  appeared  in  the 
Journal  of  the  American  Rocket  Society;  in  it,  the  Bell  engineer  specified  the 
use  of  large  spherical  reflector  satellites,  much  like  the  one  being  designed 
by  O'Sullivan,  for  long-range  telecommunications.  Such  satellites  would 
be  "passive"  rather  than  "active."  A  passive  satellite  served  simply  as  an 
electronic  mirror,  retransmitting  back  to  earth  only  those  signals  that  were 
intercepted.*  The  chief  advantage  of  a  passive  system,  Pierce  indicated,  was 
that  a  passive  satellite  was  less  complicated  electronically  than  an  active 
satellite.  Unlike  a  passive  satellite,  an  active  one  could  receive  and  amplify 
signals  before  retransmitting  them  to  the  ground,  but,  in  order  to  do  so,  it 
had  to  carry  its  own  power  supply  or  possess  the  means  of  deriving  power 
from  external  sources.19 

A  global  communications  network  based  on  a  series  of  geosynchronous 
satellites  like  those  suggested  by  Pierce  and  Clarke  interested  O'Sullivan,  but 


The  navy  was  soon  to  use  this  concept  as  the  basis  for  an  experimental  system  called  "Communi- 
cation by  Moon  Relay,"  in  which  the  moon  was  used  as  a  passive  reflector  for  radar  waves. 

163 


Spaceflight  Revolution 

in  January  1956,  as  a  member  of  the  Upper  Atmosphere  Rocket  Research 
Panel,  his  sights  were  set  on  a  much  more  limited  and  immediate  goal, 
the  upcoming  IGY.  So  just  as  quickly  as  he  began  to  speculate  about  the 
potential  of  communications  satellites  in  earth  orbit,  the  Langley  engineer 
once  again  narrowed  his  focus  and  concentrated  on  the  requirements  of  the 
air-density  experiment  at  hand.  These  other  applications  "were  things  for  a 
few  years  in  the  future,"  he  said  to  himself,  "not  for  the  year  1956."  20  Little 
did  he  know  how  quickly  those  wildly  ambitious  applications  would  be 
realized  once  the  spaceflight  revolution  began. 


Finessing  the  Proposal 

Having  pondered  the  problems  of  designing  an  air-density  flight  experi- 
ment into  the  wee  hours  of  the  morning,  O'Sullivan  finally  went  to  bed.  But 
he  could  not  sleep.  He  tossed  and  turned,  worrying  that  when  he  disclosed 
his  idea  to  the  Upper  Atmosphere  Rocket  Research  Panel  the  next  day  he 
would  find  that  he  had  "overlooked  some  factor  that  would  invalidate  the 
whole  idea."  At  one  point,  he  sat  up  in  bed,  laughed,  and  said  aloud,  "It 
will  probably  go  over  like  a  lead  balloon!"21  His  plastic-covered,  inflatable 
metal-foil  sphere  was  about  as  close  to  a  lead  balloon  as  any  professional 
engineer  would  ever  want  to  get. 

The  next  day,  January  27,  after  hearing  several  members  of  the  panel 
express  their  disappointment  in  the  proposed  methods  of  measuring  satellite 
drag  and  air  density,  O'Sullivan  mustered  enough  courage  to  tell  a  few  of 
the  panel  members  about  his  design.  He  talked  to  Fred  L.  Whipple,  then 
of  the  Harvard  College  Observatory  (and  soon  to  be  named  director  of  the 
Smithsonian  Astrophysical  Laboratory),  as  well  as  to  Raymond  Minzer  of 
the  U.S.  Air  Force  Cambridge  Research  Center.  Their  principal  concern 
was  that  the  payload  space  in  the  Vanguard  satellite  was  almost  completely 
taken  up  by  other  experiments.  All  that  was  left  for  O'Sullivan's  inflatable 
balloon  was  a  tiny  space  the  size  of  a  doughnut.  Could  the  balloon  be  made 
to  fit?  And  could  it  be  made  to  weigh  no  more  than  seven-tenths  of  a  pound? 
O'Sullivan  was  not  sure  about  meeting  either  requirement,  but,  puffing  on 
a  cigarette,  said  he  would  try  to  work  it  out.  That  was  enough  of  an  answer 
for  Whipple  and  Minzer.  Both  men  urged  O'Sullivan  to  put  his  proposal 
in  writing  and  submit  it  to  the  U.S.  National  Committee/International 
Geophysical  Year  (USNC/IGY)  Technical  Panel  on  the  Earth  Satellite 
Projects,  which  was  being  formed  in  Ann  Arbor  that  afternoon.22 

For  O'Sullivan,  the  proposal  posed  a  problem.  Most  technical  panels 
of  the  USNC/IGY  had  already  been  formed,  and  he  had  just  accepted  an 
appointment,  with  the  NACA's  permission,  on  the  Technical  Panel  on  Rock- 
etry. Not  only  was  he  a  member  of  this  panel,  but  he  also  was  responsible  for 
coordinating  the  NACA's  development  of  two  sounding,  rockets:  the  Nike- 
Deacon  (DAN)  and  an  improved  version  of  it,  the  Nike-Cajun  (CAN).  Both 

164 


The  Odyssey  of  Project  Echo 

were  to  become  mainstays  of  the  USNC/IGY  sounding  rocket  program.  Fur- 
thermore, O'Sullivan  knew  that  a  rather  strict  USNC/IGY  policy  required 
that  a  "principal  experimenter"  accept  complete  responsibility  for  carrying 
out  his  experiment  from  beginning  to  end.  The  USNC/IGY  would  deal 
only  with  him,  not  with  any  organization  with  which  he  was  associated,  in 
all  matters  pertaining  to  his  experiment,  including  funding.  The  purpose  of 
this  policy  was  to  ensure  that  every  scientist,  regardless  of  institutional  affil- 
iation or  backing,  would  enjoy  an  equal  opportunity  to  propose  experiments 
and  to  obtain  the  necessary  funding  from  the  USNC/IGY  if  the  experiment 
was  accepted.  Taking  on  the  heavy  duties  of  a  principal  experimenter  would 
be  a  full-time  job  that  O'Sullivan  could  not  possibly  do  and  still  hold  his 
civil  service  position  with  the  NACA.  The  options  appeared  to  be  either 
to  resign  his  18-year  position  with  the  NACA  and  obtain  funding  to  do  the 
experiment  from  the  USNC/IGY  or  to  forget  about  the  inflatable  satellite.23 

O'Sullivan  found  another  option,  which  was  to  share  the  satellite  project 
with  someone  else.  He  talked  again  with  Raymond  Minzer,  this  time  about 
joining  him  as  a  "co-experimenter."  Minzer  agreed,  and  within  a  few  weeks, 
the  two  men  submitted  a  successful  proposal  to  Dr.  Richard  W.  Porter, 
the  General  Electric  engineer  in  charge  of  the  V-2  test  program  at  White 
Sands  and  chairman  of  the  Technical  Panel  on  the  Earth  Satellite  Projects. 
Unfortunately,  after  the  proposal  was  accepted,  the  air  force  withdrew 
its  support  of  the  experiment,  and  Minzer  had  to  bow  out.  Once  again, 
O'Sullivan  was  left  alone  with  his  lead  balloon,  and  by  that  time,  O'Sullivan 
explains,  the  USNC/IGY  Technical  Panel  on  the  Earth  Satellite  Projects 
was  "hounding  [him]  to  get  the  experiment  under  way."24 

Upon  returning  to  Langley,  the  only  option  left  open  to  O'Sullivan, 
besides  dropping  the  experiment,  was  to  secure  full  support  from  his 
employer.  As  the  NACA's  representative  on  the  Upper  Atmosphere  Rocket 
Research  Panel,  O'Sullivan  had  reported  all  of  his  activities  in  travel  reports 
and  memoranda  that  were  routed  to  the  office  of  the  Langley  director 
(still  Henry  Reid),  with  copies  forwarded  to  NACA  headquarters.  In  mid- 
June  1956,  Associate  Director  Floyd  Thompson  and  Bob  Gilruth,  then  the 
assistant  director  responsible  for  PARD,  had  heard  all  about  the  concept 
for  an  inflatable  satellite.  They  advised  O'Sullivan  to  prepare  a  formal 
memorandum  giving  the  complete  theory  of  the  experiment  and  requesting 
that  the  NACA  sponsor  the  experiment  for  development  at  Langley  as 
another  NACA  contribution  to  the  IGY. 

Immediately,  O'Sullivan  wrote  the  proposal,  dated  29  June  1956.  In  it,  he 
explained  why  the  NACA,  an  organization  hitherto  devoted  to  the  progress 
of  aircraft,  "not  only  should,  but  must"  become  engaged  in  development  of 
earth  satellites.  With  the  recent  advances  in  rocket  propulsion  and  guidance 
systems,  O'Sullivan  argued,  "earth  satellites  can  and  will  be  developed 
and  used  for  numerous  defense  and  commercial  purposes.  .  .  .  Not 
least  among  the  foreseeable  benefits  is  the  lessening  of  world  tension  by 
bringing  closer  together  the  various  nations  through  interest  in  a  common 

165 


Spaceflight  Revolution 

beneficial  development."  The  development  of  earth  satellites  was  therefore 
"inevitable."  For  the  NACA  not  to  be  involved  with  satellites  would  be 
a  serious  mistake.  "In  every  industry,  failure  to  undergo  evolution  in  pace 
with  technological  development  inevitably  leads  to  extinction.  In  the  field  of 
research,  by  virtue  of  it  being  the  technological  frontier,  no  time  lag  between 
recognition  of  an  important  problem  and  initiation  of  work  upon  it  can 
exist  without  loss  of  ground."  To  begin,  the  NACA  should  perform  research 
"particularly  in  the  field  of  air  drag  measurement,  employing  lightweight 
inflatable  spheres,"  with  a  special  task  group  established  at  Langley  to 
perform  the  necessary  technical  work.  Given  the  low  estimated  cost  of 
the  experiment,  which  O' Sullivan  placed  very  conservatively  at  just  over 
$20,000,  he  was  hopeful  that  NACA  management  would  accept  his  proposal, 
even  though  he  knew  that  his  employer  would  have  to  bear  all  the  expenses 
of  developing  the  project  because  federal  law  prevented  the  NACA  from 
accepting  any  funds  from  the  USNC/IGY.25 

Very  quickly  the  NACA  accepted  the  proposal.  Hugh  Dry  den,  the 
director  of  research  for  the  NACA  in  Washington,  liked  the  concept  and 
in  September  1956  authorized  John  W.  Crowley,  his  associate  director,  to 
report  to  the  USNC/IGY  Technical  Panel  on  the  Earth  Satellite  Projects 
with  news  of  the  NACA's  willingness  to  develop  the  satellite.  But  the 
advocacy  was  not  over.  Not  everyone  on  Dr.  Richard  Porter's  newly 
constituted  technical  panel  had  heard  about  O 'Sullivan's  idea,  and  many  of 
them  needed  to  be  convinced.  O' Sullivan  remembers,  "I  had  to  describe  [the 
experiment]  in  minute  detail  and  defend  it  against  all  scientific  and  technical 
objections  the  [panelists]  could  think  of."  This  was  "the  acid  test,"  for 
most  members  of  this  panel  came  from  academe  and  not  from  government; 
everyone  on  the  panel  had  a  Ph.D.,  and  O'Sullivan  did  not.  After  a  careful 
presentation  of  his  proposal,  however,  O'Sullivan  managed  to  clear  the 
hurdle  and  persuade  the  panel  to  accept  the  NACA  project.  At  a  meeting 
on  9  October  1956,  Porter's  committee  put  it  on  the  official  list  of  approved 
experiments  and  designated  O'Sullivan  as  the  principal  experimenter.  The 
committee  was  convinced  that  no  other  means  for  measuring  satellite  drag 
and  thus  deducing  air  density  in  the  upper  atmosphere  approached  the 
sensitivity  of  O'Sullivan's  little  inflatable  balloon.26 


The  "Sub-Satellite" 

The  panel's  approval  gave  O'Sullivan's  air-density  experiment  only  the 
right  to  compete  for  what  little  space  remained  on  the  Vanguard  launching 
rocket;  it  did  not  guarantee  that  the  experiment  would  ever  be  flown.  The 
sole  allotment  of  space  remaining  in  the  payload  amounted  to  a  few  cubic 
inches  of  space  in  an  annular  or  ring-shaped  area  between  the  head  end  of  the 
third  stage  of  the  rocket  motor  and  the  placement  of  an  IGY  magnetometer 
satellite  developed  by  the  NRL.  Into  these  cramped  quarters,  O'Sullivan 

166 


The  Odyssey  of  Project  Echo 

and  his  helpers  at  Langley  would  have  to  squeeze  their  satellite,  along  with 
its  inflation  mechanism  and  surrounding  container.  All  of  it  together  could 
be  no  more  than  20  inches  in  diameter  or  a  mere  seven-tenths  of  a  pound. 
Because  it  was  so  little  and  was  to  be  carried  into  orbit  underneath  the 
magnetometer  satellite,  O'Sullivan  named  the  small  inflatable  vehicle  the 
"Sub-Satellite."* 

Although  the  USNC/IGY  technical  panel  designated  O'Sullivan  as  the 
principal  experimenter  for  the  balloon  project,  too  many  problems  had  to  be 
solved  in  too  short  a  time  for  one  man  to  do  all  the  work  alone.  Therefore,  in 
the  fall  of  1956,  Floyd  Thompson  authorized  the  formation  of  a  small  team 
of  engineers  and  technicians,  mostly  from  PARD,  to  assist  O'Sullivan  in  the 
preparation  of  the  satellite  project.  Administratively,  Thompson  facilitated 
this  in  late  December  1956  by  appointing  O'Sullivan  as  head  of  a  new 
Space  Vehicle  Group  placed  inside  PARD.  The  group  would  report  directly 
to  the  division  office,  which  was  headed  by  Joseph  A.  Shortal.  Jesse  L. 
Mitchell  of  the  Aircraft  Configurations  Branch  and  Walter  E.  Bressette  from 
the  Performance  Aerodynamics  Branch  of  PARD  would  assist  O'Sullivan. 
Significantly,  this  small  Space  Vehicle  Group  was  the  first  organizational 
unit  at  Langley  to  have  the  word  "space"  in  its  title.27 

First,  the  Space  Vehicle  Group  tested  dozens  of  plastic  and  metal  foils 
(even  gold)  in  search  of  the  right  combination  to  withstand  the  extreme 
range  of  temperatures  that  the  little  satellite  would  encounter:  from  300°F 
in  direct  sunlight  to  -80°F  when  in  the  shadow  of  the  earth.  The  group 
found  half  of  the  answer  to  the  problem  in  a  new  plastic  material  called 
"Mylar."  Made  by  E.  I.  du  Pont  de  Nemours  &  Co.,  Mylar  was  being 
used  for  recording  tape  and  for  frozen-food  bags  that  could  be  put  directly 
into  hot  water.  When  manufactured  in  very  thin  sheets,  perhaps  only  half 
as  thick  as  the  cellophane  wrapper  on  a  pack  of  cigarettes,  Mylar  plastic 
proved  enormously  tough.  It  showed  a  tensile  strength  of  18,000  pounds  per 
square  inch,  which  was  two-thirds  that  of  mild  (low-carbon  content)  steel. 

The  second  half  of  the  answer,  that  is,  an  effective  metal  covering  for 
the  plastic  that  could  protect  the  satellite  from  radiation  and  make  it 
visible  to  radar  scanners,  proved  a  little  more  difficult  to  find.  For  more 
than  a  month,  the  O'Sullivan  group  "tested  metal  after  metal,  looking  for 
ways  to  paint  them  on  Mylar  in  layers  far  thinner  than  airmail  onionskin 
paper."28  Then,  one  man  in  the  Space  Vehicle  Group  heard  about  a 
technique  for  vaporizing  aluminum  on  plastic  that  the  Reynolds  Metals 


>k 

An  interesting  and  seemingly  appropriate  name,  it  nonetheless  turned  out  to  be  problematic 
technically,  because  of  confusion  with  the  term  "subsatellite  point,"  a  term  from  orbital  mechanics 
that  defined  the  point  of  intersection  where  a  straight  line  (known  as  the  "local  vertical" )  drawn  from  a 
satellite  to  the  center  of  the  body  being  orbited  (in  this  case,  the  earth)  cuts  through  the  surface  of  that 
body.  The  confusion  would  grow  worse  in  the  late  1960s  when  the  term  "subsatellite"  came  to  be  used  to 
describe  small  artificial  satellites  ejected  from  other  satellites  or  spacecraft,  such  as  those  released  from 
Apollo  15  in  1971  and  Apollo  16  in  1972  for  the  purpose  of  carrying  out  certain  scientific  experiments. 

167 


Space/light  Revolution 


To  determine  the  capacity  of  the  30- 
inch  "Sub-Satellite"  (right)  to  withstand 
the  high  temperature  of  direct  sunlight  in 
space,  Langley  researchers  subjected  it  to 
a  450 °F  heat  test  (below).  Results  indi- 
cated that  the  aluminum- covered  Mylar 
plastic  would  effectively  reflect  the  danger- 
ous heat. 


L-57-3113 


L-58-4293 


168 


The  Odyssey  of  Project  Echo 

Company  of  nearby  Richmond,  Virginia,  was  experimenting  with  for  the 
development  of  everyday  aluminum  foil.  This  new  and  unique  material 
was  acquired  and  successfully  tested.  The  fabrication  problem  was  solved 
by  cutting  the  material  into  gores,  that  is,  into  three-cornered  or  wedge- 
shaped  pieces,  and  gluing  them  together  along  overlapping  seams.  Using 
this  technique  with  this  material,  the  Langley  researchers  built  the  outer 
skin  of  their  first  20-inch  domes  for  inflation  tests. 

Almost  everyone  involved  was  excited  by  the  prospect  of  sending  the 
experiment  into  space,  and  several  individuals  worked  nights  and  weekends 
through  the  last  months  of  1956  to  get  the  Sub-Satellite  ready.  The  right 
blend  of  materials  had  been  found  for  the  inflatable  sphere;  now,  two  major 
problems  remained:  how  to  fold  the  sphere  so  that  it  could  expand  quickly 
without  a  single  one  of  its  folds  locking  up  and  causing  a  tear,  and  how  to 
inflate  the  balloon.  As  for  the  means  of  inflation,  the  Langley  researchers 
tried  dozens  of  strange  chemicals  before  discovering  that  a  small  bottle  of 
nitrogen  would  inflate  the  little  balloon,  at  the  proper  rate,  so  that  the 
sphere  would  not  blow  apart.  Learning  how  to  fold  the  satellite  was  also 
a  purely  empirical  process;  no  theory  existed  to  guide  the  group.  As  one 
observer  remembers,  "Harassed  by  O 'Sullivan,  men  who  couldn't  fold  a  road 
map  properly  found  a  way  to  fold  his  aluminum  balloon."29 

However,  this  characterization  gives  O' Sullivan  too  much  credit.  Walter 
Bressette  and  Edwin  C.  Kilgore  were  the  engineers  who  actually  worked 
out  not  only  the  folding  pattern  but  also  the  ejection  method  and  inflation 
bottle  pressure  for  the  Sub-Satellite.  Bressette,  an  airplane  pilot  in  World 
War  II  and  a  1948  graduate  in  mechanical  engineering  from  Rhode  Island 
College  (now  the  University  of  Rhode  Island),  had  spent  the  last  10  years 
working  on  ramjet  propulsion  systems,  jet  effects  on  airplane  stability,  and 
reentry  problems  using  both  rocket-propelled  vehicles  and  a  supersonic 
blowdown  tunnel  at  Wallops  Island.  Kilgore,  a  1944  engineering  graduate 
from  Virginia  Polytechnic  Institute  and  State  University,  had  proved  to  be 
one  of  Langley's  top  machine  designers.  The  two  men  conducted  many 
trials  in  a  small  vacuum  chamber  in  the  PARD  shop  before  solving  the  Sub- 
Satellite's  problems.  In  addition,  Bressette  made  frequent  trips  to  the  NRL 
in  Washington  to  put  the  Sub-Satellite  package  through  what  he  called  "the 
shake,  rattle,  and  roll"  of  vehicle  environmental  tests.30 

By  January  1957,  the  Sub-Satellite  was  almost  ready.  A  front-page  article 
appearing  in  the  Langley  Air  Scoop  on  5  January  touted  the  little  satellite 
for  having  originated  at  Langley,  credited  O' Sullivan  with  "having  conceived 
the  novel  manner  of  construction,"  and  told  employees  to  look  forward  to 
its  impending  launch.  Next  to  the  article  was  a  photograph  of  O'Sullivan 
holding  in  his  right  hand  the  shiny  inflated  Sub- Satellite,  the  emblem  of 
the  NACA  wing  on  its  side,  and  in  his  left  hand,  a  folded  uninflated  Sub- 
Satellite.31 

However,  just  when  everything  was  proceeding  on  schedule,  a  compli- 
cation developed.  The  Baker-Nunn  precision  optical  tracking  cameras  at 

169 


Spaceflight  Revolution 

White  Sands  would  not  be  able  to  follow  and  photograph  such  a  small 
sphere;  Fred  Whipple  urged  O 'Sullivan  and  the  NAG  A  to  increase  the  size 
of  the  Sub-Satellite  from  20  to  30  inches  in  diameter.  The  Sub-Satellite 
could  not  weigh  more  or  take  up  more  space  on  the  Vanguard;  it  just  had  to 
be  10  inches  bigger.  On  7  February  1957,  Whipple's  panel  made  this  request 
official,  reconfirming  assignment  of  the  NACA  experiment  on  Vanguard  if 
the  change  was  made.  O'Sullivan  believes  that  the  new  size  requirement 
"would  have  been  a  deathblow  to  the  Sub-Satellite  had  it  occurred  at  the 
start";  however,  with  the  experience  gained  at  Langley  through  actually 
building  the  sphere,  the  increase  to  a  30-inch  diameter  was  accomplished 
over  the  next  few  months  without  too  much  trouble.32 

After  all  Langley's  work,  the  Sub-Satellite  was  finally  on  the  launchpad 
at  Cape  Canaveral  on  13  April  1959.  Seconds  after  takeoff,  the  second  stage 
of  the  Vanguard  SLV-5  vehicle  experienced  a  failure  that  sent  the  rocket 
and  the  Sub-Satellite  crashing  ignominiously  into  the  depths  of  the  Atlantic 
Ocean.  With  this  launch  failure,  the  attempt  to  determine  air  density  with 
the  Sub-Satellite  came  to  an  end.  (This  was  Vanguard's  third  attempted 
launch.)  However,  other  models  of  the  30-inch  sphere  were  used  for  a  short 
time  both  before  and  after  the  SLV-5  misfire  as  a  calibration  target  for  a  new 
long-range  radar  being  developed  at  MIT's  Lincoln  Laboratory  at  Millstone 
Hill  Radar  Observatory  in  Westford,  Massachusetts.33 


Something  the  Whole  World  Could  See 

Even  before  the  Sub-Satellite's  fatal  plunge  into  the  ocean  in  April  1959, 
O'Sullivan  had  started  to  contemplate  the  benefits  of  a  larger  reflector 
satellite  that  could  be  the  sole  payload  on  a  Vanguard.  In  November  1957, 
Fred  Whipple  presided  over  a  space  science  symposium  in  San  Diego,  which 
was  sponsored  jointly  by  the  air  force  and  Convair  Astronautics.  At  this 
symposium,  O'Sullivan  proposed  that  a  large  inflatable  launched  by  a  rocket 
more  powerful  than  Vanguard  could  be  used  as  a  lunar  probe.  "It  could  be 
seen  and  photographed  through  existing  astronomical  telescopes,  not  only 
giving  conclusive  proof  to  everyone  that  such  a  probe  had  reached  the  moon, 
but  its  location  as  it  orbited  the  moon  or  impacted  on  the  moon  would  be 
known."  Before  sending  a  balloon  to  the  moon,  however,  O'Sullivan  felt 
that  something  must  be  put  into  earth  orbit,  perhaps  a  12- foot-diameter 
satellite,  which  "the  whole  world  could  see."34 

For  a  professional  engineer,  O'Sullivan  was  something  of  a  universalist. 
He  worked  on  airplanes,  missiles,  and  satellites;  he  knew  aerodynamics,  and 
he  knew  space.  But  he  was  not  gracious  about  sharing  credit.  The  idea  for 
the  12-foot  satellite  was  not  O'Sullivan's  but  Jesse  Mitchell's.  An  analysis 
performed  by  Mitchell  had  indicated  that  a  30-inch-diameter  sphere  would 
not  make  a  suitable  optical  device  for  a  lunar  probe;  the  sphere  would  have  to 
be  several  times  larger.  So  in  the  summer  of  1957  while  O'Sullivan  was  away 

170 


The  Odyssey  of  Project  Echo 


Space  Vehicle  Group  engineer  Jesse 
Mitchell  examines  the  Sub-Satellite 
package  in  early  October  1958,  only 
days  after  NASA 's  establishment. 
Mitchell,  who  would  later  become  head 
of  the  Geophysics  and  Astronomy  Di- 
vision at  NASA  headquarters,  directed 
Langley's  program  development  plan 
for  Echo  1 . 
L-58-393a 


from  Langley,  Mitchell  and  Bressette  had  consulted  the  model  shop  about 
building  a  larger  sphere.  The  size  of  the  sphere  became  12  feet  because  of 
the  ceiling  height  in  the  model  shop,  not  because  O' Sullivan  had  determined 
it  to  be  the  perfect  size.35 

For  millions  of  people,  the  spaceflight  revolution  began  the  first  time 
they  looked  up  in  wonder  at  the  bright  twinkling  movement  of  an  artificial 
satellite.  O' Sullivan  was  aware  of  this  when  he  proposed  his  12-foot 
inflatable.  With  the  appearance  of  Sputnik  1  a  month  earlier  in  October 
1957,  people  around  the  world,  especially  Americans,  developed  a  heightened 
if  not  exaggerated  interest  in  searching  the  sky  for  UFOs.  A  widespread 
interest  in  UFOs  had  existed  before  the  ominous  overflights  of  the  Russian 
satellites.  As  historian  Walter  McDougall  explains  in  his  analysis  of  the 
onset  of  the  space  age,  10  years  prior  to  the  first  Sputnik,  "beginning  in  the 
midsummer  of  1947  the  American  people  began  to  see  unidentified  flying 
objects,  kicking  off  a  flying  saucer  'epidemic'  of  such  proportions  that  the 
air  force  launched  a  special  investigation  and  began  compiling  thousands  of 
case  studies  that,  in  the  end,  satisfied  no  one."36  The  cause  of  the  epidemic 
was  the  new  need  of  Americans  to  externalize  their  postwar  fears  about 
technology,  about  the  atomic  bomb,  and  about  nuclear  war  destroying  the 
world. 


171 


Space/light  Revolution 

Into  the  blackness  of  that  anxiety-ridden  mass  psychology  came  the 
specter  of  Sputnik.  Across  the  United  States,  people  went  outside  with 
binoculars  and  telescopes,  straining  to  see  the  faint  blinking  reflection  of 
the  tiny  yet  ominous  metal  globe  tumbling  end  over  end.  For  instance,  in 
San  Francisco  on  Friday  night,  4  October  1957,  volunteer  crews  of  amateur 
astronomers  with  special  "moon-watch"  telescopes  maintained  a  vigil  atop 
the  Morrison  Planetarium  in  Golden  Gate  Park  in  hopes  of  sighting  the 
Russian  satellite.  Crew  members  took  up  their  prearranged  stations  as  soon 
as  reports  of  the  satellite's  launching  were  received.  Six  tireless  individuals 
continued  the  lonely  vigil  until  morning,  when  conditions  for  viewing  were 
allegedly  at  their  best.  How  many  people  actually  spotted  the  satellite  that 
night  and  over  the  next  several  months  as  it  moved  in  its  north-to-south  orbit 
is  unknown,  but  certainly  far  fewer  saw  Sputnik  1  than  said  they  did.37 

On  that  same  evening  in  early  October  on  a  large  ranch  in  Texas,  Senate 
Majority  Leader  Lyndon  B.  Johnson  was  having  a  few  guests  in  for  dinner 
when  the  news  of  Sputnik  1  came  over  the  radio  and  television.  After  eating, 
the  party  went  outside  rather  nervously  for  what  was  supposed  to  be  a 
calming  stroll  in  the  dark  along  the  road  to  the  Padernales  River.  But  the 
walk  only  unnerved  them.  As  one  of  the  guests,  Gerald  Siegel,  a  lawyer  with 
the  Senate  Democratic  Policy  Committee,  remembers  thinking  at  the  time: 
"In  the  Open  West  you  learn  to  live  closely  with  the  sky.  It  is  a  part  of  your 
life.  But  now,  somehow,  in  some  new  way,  the  sky  seemed  almost  alien.  I 
also  remember  the  profound  shock  of  realizing  that  it  might  be  possible  for 
another  nation  to  achieve  technological  superiority  over  this  great  country 
of  ours."38 

On  the  Atlantic  coast,  among  the  millions  of  people  all  over  the  country 
and  the  world  who  were  looking  up  in  the  sky  that  night  to  see  Sputnik 
were  O' Sullivan  and  his  colleagues  at  NACA  Langley.  Bob  Gilruth  recalls 
seeing  the  satellite  from  his  bayside  home  in  Seaford,  Virginia.  In  Gilruth's 
words,  the  sighting  "put  a  new  sense  of  value  and  urgency"  on  everything  he 
and  his  co-workers  were  doing  at  Langley.  Charles  Donlan  also  remembers 
sighting  the  little  satellite:  "I  was  running  around  my  yard  in  Hampton 
one  evening,  when  I  looked  up  and  saw  Sputnik  go  right  over  my  house.  I 
remember  stopping  and  staring  at  it.  What  I  remember  thinking  was  how 
much  better  it  would  be  if  the  thing  belonged  to  America."39 

Everyone  involved  with  decisions  regarding  U.S.  satellites,  including  the 
State  Department  and  the  Central  Intelligence  Agency  (CIA),  felt  the  same 
way.  In  the  wake  of  the  Sputniks,  virtually  all  government  officials  concerned 
expressed  a  desire  to  orbit  a  satellite  that  would  be  visible  over  Russia  as 
well  as  the  United  States.  O'Sullivan's  12-foot  inflatable  sphere  seemed  to 
fit  the  bill.  Because  of  shocking  world  events,  what  had  started  out  as  a 
simple  air-density  experiment  was  becoming  an  instrument  of  propaganda 
in  the  cold  war. 

The  idea  for  an  inflatable  sphere  big  enough  for  everyone  to  see  received 
high  priority.  Whipple  and  other  members  of  the  USNC/IGY  Technical 

172 


The  Odyssey  of  Project  Echo 

Panel  on  the  Earth  Satellites  expressed  serious  interest  in  O' Sullivan's 
proposal  for  a  bigger  inflatable,  but  they  had  to  wait  a  few  months  to 
see  whether  a  booster  more  powerful  than  the  Vanguard  rocket  could  be 
obtained.  Finally,  in  the  spring  of  1958,  the  USNC/IGY  informed  the 
NACA  that  some  space  was  available  inside  the  nose  cone  of  a  Jupiter  C,  an 
intermediate-range  ballistic  missile  developed  by  the  ABMA  that  was  more 
powerful  than  the  Vanguard  rocket.  If  the  Jupiter  failed,  and  of  course  none 
of  these  boosters  had  yet  proved  reliable,  a  Juno  II,  a  new  vehicle  similar  to 
the  Jupiter  C,  might  be  available  as  the  backup.  A  Juno  II  would  launch 
America's  first  successful  lunar  flyby,  Pioneer  4,  on  3  March  1959. 40 

The  NACA  agreed  to  the  project,  and  the  Space  Vehicle  Group  continued 
to  construct  and  test  its  12-foot  inflatable.41  Because  it  was  to  orbit  at  300 
to  400  miles  above  the  earth  and  thus  would  appear  as  bright  as  the  north 
star,  Polaris,  the  satellite  eventually  came  to  be  called  "Beacon."  Beacon 
would  be  easy  to  see  with  the  naked  eye  and  so  could  be  tracked  optically 
and  photographically  without  difficulty.  Big  new  radars,  such  as  the  one 
being  developed  by  MIT  at  Millstone  Hill,*  were  just  coming  on  line  and 
would  be  able  to  track  day  or  night,  regardless  of  the  weather.42 

On  25  June  1958,  the  USNC/IGY  officially  assigned  the  12-foot  Beacon 
satellite  as  a  pay  load  for  the  launch  of  Jupiter  C  No.  49.  To  obtain 
the  difficult  orbit  that  O' Sullivan  insisted  on — it  was  circular  rather  than 
elliptical — the  Jupiter  C  had  to  have  a  small  "high-kick"  rocket  motor 
that  gave  an  extra  boost  to  help  the  satellite  reach  the  desired  orbit. 
Unfortunately,  on  23  October  1958,  the  "high-kick"  did  not  get  a  chance  to 
"kick  in,"  because  the  "low  kicks"  kept  failing.  As  was  the  case  with  its  30- 
inch  ancestor,  the  Beacon  was  not  launched  into  orbit  from  Cape  Canaveral 
because  the  booster  failed.  Fourteen  months  later,  Juno  II  No.  19  was  ready 
to  carry  a  second  12-foot  satellite  into  orbit  but  failed  to  do  so  when  the 
rocket's  fuel  supply  emptied  prematurely.43 

With  three  failures  in  a  row,  O'Sullivan  and  the  Space  Vehicle  Group 
might  have  given  up  on  the  balloon  if  not  for  the  spectacular  successes 
of  Explorer  1  on  31  January  and  Vanguard  1  on  17  March  1958.  These 
American  satellites  proved  not  only  that  Americans  could  put  an  object  in 
orbit  but  also  that  those  objects,  tiny  as  they  were,  could  disclose  great 
scientific  discoveries  such  as  the  Van  Allen  radiation  belts.  Beyond  that, 
satellites  could  be  of  tremendous  economic  and  social  benefit.  They  could 
make  continuous  worldwide  observation  of  the  weather  possible,  and  the 
existence  and  the  likely  paths  of  hurricanes  and  other  destructive  storms 
would  be  accurately  predicted.  By  studying  the  development  of  the  world's 
weather  patterns  from  space,  humans  might  someday  control  the  climate. 
In  summary,  satellites  offered  too  many  far-reaching  benefits  for  researchers 
to  allow  a  few  launch  vehicle  failures  to  discourage  them.  Rockets  were  still 


On  3  June  1959,  Millstone  Hill  would  transmit  a  voice  message  from  President  Eisenhower  and 
reflect  it  off  the  moon  to  Prince  Albert  in  Saskatchewan,  Canada. 

173 


Space/light  Revolution 


In  July  1959,  William  J. 
O 'Sullivan  (right,  standing)  and 
an  unidentified  engineer  exam- 
ine the  capsule  containing  the 
tightly  folded  and  packed  12- 
foot-diameter  Beacon  satellite 
(below). 


L-59-2836 


L-59-4973 


in  their  infancy,  in  the  "Model-T"  stage  of  technical  evolution.  Failures  were 
to  be  expected,  Langley's  team  consoled  itself.  All  the  problems  had  been 
with  the  boosters,  not  with  their  own  satellites.44 

According  to  O'Sullivan,  he  set  an  example  of  grit  and  determination  for 
the  rest  of  his  people,  many  of  whom  were  still  quite  young.  As  he  told  a 
magazine  writer  at  the  time,  he  was  "mindful  of  his  research  associates  who 
had  labored  so  hard"  to  produce  the  experiments  and  who  "looked  to  me  as 
their  leader."  This  was  driven  home  to  him,  he  told  the  writer,  during  the 
unsuccessful  launch  of  the  12- foot  satellite.  Watching  the  Doppler  velocity 
drop  off  rather  than  climb,  he  knew  instantly  the  launching  rocket  had  failed. 
Turning  to  his  associates,  he  said,  "The  launching  is  a  failure."  Standing 
dumbfounded,  staring  at  O'Sullivan,  one  of  them  asked,  "What  do  we  do 
now?"  O'Sullivan  immediately  answered,  "We  pack  up  our  instruments  and 
equipment  as  quickly  as  we  can.  We  haven't  a  moment  to  lose.  We  have  to 
get  back  to  the  Laboratory  and  get  the  next  satellite  ready  for  launching." 
When  his  men  started  moving  in  a  hurry,  O'Sullivan  informed  the  writer, 
he  knew  for  sure  that  he  "must  never  waiver  or  hesitate  no  matter  how 
stunning  the  blow."45 

According  to  other  key  individuals  involved  with  the  project,  however, 
O'Sullivan  was  not  the  leader  he  claimed  to  be.  Walter  Bressette  remembers 
that  "O'Sullivan  never  went  to  the  satellite  launch  areas."  In  fact,  he  gave 


174 


The  Odyssey  of  Project  Echo 


L-58-88a  L-58-1063a 

The  genius  of  William  J.  O  'Sullivan  (left)  rested  in  the  theory  stage  of  an  engineer- 
ing development;  other  Langley  researchers  took  over  the  main  responsibility  during 
the  design  and  deployment  phases.  Walter  Bressette  (right)  played  a  major  role  in 
the  Echo  program  from  start  to  finish.  Here  he  examines  a  scaled  prototype  of  the 
satelloon  in  December  1958. 


up  direction  of  the  12-foot  satellite  mission  immediately  after  the  Juno  II 
failure,  handing  it  over  to  Claude  W.  Coffee,  Jr.,  and  Bressette,  who  then 
made  the  12-foot  Scout  proposals.  O'Sullivan  abandoned  his  project,  leaving 
it  to  others  to  carry  on.  Those  who  did  continue  the  work  view  O'Sullivan's 
self-publicized  heroic  role  in  the  eventual  success  of  the  effort  as  egotistical 
and  inaccurate.46 


Big  Ideas  Before  Congress 

Up  to  the  point  of  the  Juno  II  failure,  Langley's  interest  in  inflatable 
satellites  had  been  limited  to  air-density  experiments  in  the  upper  atmo- 
sphere and  to  orbiting  an  object  large  enough  to  be  seen  by  the  naked  eye; 
the  notion  of  deploying  satellites  for  a  worldwide  telecommunications  net- 
work like  the  one  suggested  by  John  Pierce  and  Arthur  Clarke  had  not  yet 
taken  hold  as  an  immediate  possibility. 

But  the  flight  of  the  Sputniks  emboldened  conservative  researchers.  In 
the  spring  of  1958,  as  plans  for  NASA  were  being  formulated  in  Washington, 
communications  satellites  or  "comsats"  became  a  moderately  hot  topic.  Not 
surprisingly,  even  the  NAG  A  began  to  take  a  healthy  interest  in  them.  At 

175 


Space/light  Revolution 

Langley,  an  advance  planning  committee  recommended  that  the  center  begin 
a  comprehensive  study  of  radio- wave  propagation  and  channel  requirements, 
as  well  as  the  requirements  for  active  relays.  In  a  decision  that  would  later 
come  to  haunt  them,  the  planning  committee  resolved  that  the  first  flight 
experiment  should  involve  only  a  simple  passive  reflector,  one  in  which  the 
satellite  acted  merely  as  a  mirror  and  retransmitted  only  those  signals  it 
received.  That  sort  of  simple  experimental  communications  satellite  could 
be  placed  in  orbit  very  soon,  perhaps  as  early  as  fiscal  year  1959,  the  Langley 
planners  stated.  The  greater  difficulties  of  building  an  active  system  were 
being  tackled  elsewhere.  A  passive  flight  experiment  would  demonstrate  the 
feasibility  of  a  space-based  system,  and  the  new  NASA  could  accomplish  the 
task  largely  on  its  own,  without  extensive  help  from  industrial  contractors, 
notably  Radio  Corporation  of  America  (RCA),  American  Telephone  and 
Telegraph  (AT&T),  and  General  Electric  (G.E.),  who  at  that  time  were 
petitioning  the  federal  government  to  invest  in  their  own  special  comsat 
projects.47 

Throughout  the  spring  and  summer  of  1958,  Congress  listened  to  argu- 
ments about  the  potential  of  space  exploration  and  what  should  be  done  to 
ensure  that  the  country's  nascent  "into  space"  enterprises  would  continue 
far  beyond  the  end  of  the  ICY.  This  testimony,  in  part,  was  the  genesis 
of  NASA.  The  NACA's  director  of  research,  Hugh  Dryden,  testified  more 
than  once  on  Capitol  Hill.  Before  the  House  Select  Committee  on  Science 
and  Astronautics  on  22  April,  Dryden  explained,  among  many  other  things, 
how  large  aluminized  balloons  could  be  inflated  in  orbit  and  used  for  com- 
munication tests.  Accompanying  him  on  this  occasion  was  O' Sullivan,  who 
took  a  full-size  Beacon  satellite  into  the  Capitol  and  inflated  it  there  "to 
demonstrate  the  structural,  optical,  and  electronic  principles  involved."  In 
his  testimony,  O'Sullivan  delighted  the  congressmen  by  saying,  quite  em- 
phatically, that  Langley  had  been  studying  the  problem  of  communications 
satellites  for  several  months  and  that  its  staff  was  absolutely  convinced  that 
a  very  large  inflatable  reflecting  sphere,  at  least  10  stories  high,  could  be 
built  quickly  and  launched  into  space.  This  big  balloon  "would  reflect  ra- 
dio signals  around  the  curvature  of  the  earth  using  frequencies  not  other- 
wise usable  for  long  range  transmission,  thus  mostly  increasing  the  range 
of  frequencies  for  worldwide  radio  communications  and,  eventually,  for  tele- 
vision, thus  creating  vast  new  fields  into  which  the  communications  and 
electronics  industries  could  expand  to  the  economic  and  sociological  benefit 
of  mankind."48 

The  ideas  of  Pierce  and  Clarke  were  finding  a  home  at,  of  all  places,  a 
government  aeronautics  laboratory.  On  31  March  1958,  some  three  weeks 
before  Dryden  and  O'Sullivan  testified  in  Washington,  John  W.  "Gus" 
Crowley,  Dryden's  associate  director,  had  visited  Langley  and  told  Floyd 
Thompson,  O'Sullivan,  Joseph  Shortal,  and  others  that  Dryden  had  been 
having  conversations  with  Dr.  Pierce  of  Bell  Telephone  Labs  and  with 
members  of  President  Eisenhower's  Science  Advisory  Committee  about 

176 


The  Odyssey  of  Project  Echo 

the  potential  of  a  global  telecommunications  system  based  on  satellites. 
What  NACA  headquarters  now  wanted  to  know,  Crowley  said,  was  whether 
Langley  was  interested  in  constructing  a  larger  100- foot  inflatable  sphere,  on 
a  tight  schedule,  to  be  used  as  an  orbital  relay  satellite  like  that  envisioned 
by  Pierce.49 

O' Sullivan  assured  Crowley  a  few  days  later  that  his  Space  Vehicle  Group 
was  "not  only  interested  but  enthusiastic  about  the  possibility  of  placing 
such  a  satellite  in  orbit,  and  that  the  schedule  could  be  met."  On  3  April 
1958,  a  follow-up  meeting  took  place  at  Shortal's  PARD  office  to  consider 
designs  for  the  big  balloon.  At  this  meeting,  O'Sullivan,  adopting  Jesse 
Mitchell's  scheme,  suggested  using  the  100-foot  sphere  as  a  lunar  probe.  On 
18  April,  Langley  submitted  to  NACA  headquarters  a  proposed  research 
authorization  entitled,  "A  Large  Inflatable  Object  for  Use  as  an  Earth 
Satellite  or  Lunar  Probe."  The  NACA  did  not  formally  approve  the  proposal 
until  8  May,  but  work  on  the  big  sphere  had  actually  started  at  Langley  on 
a  high-priority  basis  even  before  Crowley's  visit.50 

In  early  February  1959,  Project  Echo,  as  O'Sullivan  had  begun  to  call 
it,  cleared  another  major  hurdle  when  NASA  headquarters  assured  Lang- 
ley  that  an  allotment  of  space  would  be  devoted  to  the  large  inflatable  in 
a  forthcoming  "space  shot."  Following  this  authorization,  on  19  February, 
Langley  Assistant  Director  Draley  approved  the  creation  of  a  large  interdis- 
ciplinary "task  group"  of  approximately  200  people,  assigned  on  a  temporary 
basis  without  change  of  organization  and  initially  under  O' Sullivan's  leader- 
ship. The  Space  Vehicle  Group  alone  could  not  handle  the  entire  work  load, 
which  at  this  point  still  involved  the  30-inch  Sub-Satellite  and  the  12-foot 
inflatable.  Significantly,  as  befitting  a  project  that  had  to  succeed,  Draley 
announced  that  the  move  was  necessary  to  meet  an  "emergency."  He  in- 
formed the  directorate  that  "for  the  duration  of  this  emergency  condition," 
O 'Sullivan's  Space  Vehicle  Group  and  Clarence  L.  Gillis's  Aircraft  Config- 
uration Branch,  both  of  PARD,  "will  merge  and  work  as  one  unit"  with 
O'Sullivan  as  head  and  Gillis  as  his  deputy.  To  make  room  for  the  work 
load  in  this  merged  group,  "it  may  be  necessary  to  postpone,  or  transfer 
to  other  units,  some  of  the  work  now  in  progress."  In  other  words,  Project 
Echo  took  priority  over  business-as-usual,  and  everyone  at  Langley  would 
just  have  to  adjust.51 

Assigning  Responsibilities 

The  first  planning  meetings  for  Project  Echo  convened  at  NASA  head- 
quarters in  the  summer  of  1959,  not  long  before  the  first  NASA  inspection. 
At  the  second  of  these  meetings,  on  13  October  1959,  Leonard  Jaffe,  chief  of 
NASA's  fledgling  communications  satellite  program  and  director  of  one  of 
the  program  offices  in  the  Office  of  Space  Sciences  at  NASA  headquarters, 
surprised  Langley  representatives  by  announcing  that  the  "primary  respon- 
sibility" for  managing  Echo  was  being  given,  not  to  Langley,  but  to  Goddard, 

177 


Spaceflight  Revolution 

which  was  still  under  construction  in  Greenbelt,  Maryland.  Various  parties 
would  contribute  to  the  project  through  expanded  in-house  activities  and 
some  extensive  contracting,  Jaffe  explained.  The  Douglas  Aircraft  Com- 
pany plant  in  Tulsa  (a  converted  B-24  factory)  would  provide  the  assembled 
booster,  a  three-stage  Thor-Delta  (later  it  would  be  called  just  a  Delta);  Bell 
Telephone  Laboratories,  where  comsat  pioneer  John  Pierce  worked  as  direc- 
tor of  electronics  research,  would  make  available  at  Holmdel,  New  Jersey,  a 
20  x  20-foot  horn-fed  parabolic  receiver,  a  60-foot  antenna,  as  well  as  ampli- 
fiers, demodulators,  and  other  electronic  and  radar  equipment;  RCA  would 
provide  the  radar  beacon  antenna  for  incorporation  upon  the  Echo  spheres; 
NRL  would  use  its  large  60-foot  dish  antenna  at  Stump  Neck,  Maryland, 
to  receive  the  reflected  signals  from  Echo;  and  JPL  would  employ  its  two 
85-foot  low-noise  antennae  at  the  Goldstone  (California)  Receiving  Site  to 
track  the  satellite.52 

Naturally,  Langley  was  quite  disturbed  over  the  assignment  of  the  overall 
responsibility  to  Goddard.  As  one  senior  Langley  researcher  remembers, 
"Echo  was  considered  to  be  but  the  first  in  a  long  series  of  large  satellite 
experiments  under  the  jurisdiction  of  Langley."  If  Langley  lost  Echo  to 
Goddard,  all  the  other  large  satellite  experiments  would  probably  go  to 
Goddard  as  well.  Whatever  proved  to  be  the  case,  however,  Langley  felt  that 
Jaffe's  instructions  need  not  have  any  immediate  effect  on  Echo.  Langley, 
both  through  in-house  work  and  the  monitoring  of  contracts,  would  keep 
the  responsibilities  for  the  key  research  and  development  tasks.  These  tasks 
were  not  spelled  out  precisely  by  Jaffe  at  the  planning  meeting,  and  more 
than  a  year  would  pass  before  a  working  agreement  satisfactory  both  to 
Goddard  and  Langley  was  finalized.  Even  after  the  agreement  was  reached  in 
January  1961,  relations  between  the  two  NASA  centers  were  stressful.  As  we 
have  seen,  tensions  already  existed  between  them.  Goddard  staff  wanted  to 
exercise  management  authority  over  a  project  they  felt  was  rightfully  theirs; 
Goddard  was  the  center  for  all  NASA  space  projects.  As  the  originators  of 
the  Echo  concept,  O'Sullivan  and  his  associates  saw  Goddard  as  an  intruder. 
Langley  researchers,  therefore,  planned  to  ignore  Goddard  and  continue 
working  as  before  the  reassignment.53 

Pending  the  final  agreement  over  the  division  of  responsibilities,  Langley's 
Project  Echo  Task  Group  continued  to  do  whatever  it  felt  needed  to  be  done 
to  assure  the  success  of  the  "satelloon."*  This  included  doing  virtually  all 


••a 

Langley  could  proceed  independently  of  Goddard  in  part  because  of  the  manner  in  which  NASA 
managed  Echo  and  provided  funding  to  Langley  for  the  project.  With  its  establishment  as  an 
official  NASA  spaceflight  project,  responsibility  for  managing  Echo  went  to  the  Office  of  Space  Flight 
Development  under  Abe  Silverstein,  who  then  assigned  the  project  to  the  Office  of  Space  Sciences, 
wherein  Leonard  Jaffe,  the  chief  of  communications  satellites,  took  over  the  regular  responsibilities. 
Funding  for  Echo  came  from  Silverstein's  bailiwick,  through  Jaffe's  office,  and  then  made  its  way  to 
Langley  via  transfers  from  Bob  Gilruth's  STG.  For  a  time,  O'Sullivan's  entire.  Space  Vehicle  Group  was 
carried  on  the  personnel  rolls  of  the  STG.  In  effect,  this  convoluted  but  cozy  arrangement  meant  that 

178 


The  Odyssey  of  Project  Echo 

of  the  preliminary  design  for  the  payload,  including  the  satellite  itself;  the 
satellite  container  with  all  its  associated  circuitry,  hardware,  and  pyrotech- 
nics; the  container-separation  or  deployment  mechanism;  and  the  inflation 
system.  The  Langley  group  developed  the  techniques  for  fabricating,  fold- 
ing, packing,  and  inflating  the  rigidized  sphere,  and  it  carried  out  the  sys- 
tematic ground  tests  to  make  sure  that  everything  worked  properly.  After 
completing  the  ground  tests,  Langley  also  assisted  in  all  launches  and  test 
flights. 

Nonetheless,  as  the  Langley  engineers  involved  would  soon  discover,  the 
assignment  of  Project  Echo  to  the  Goddard  Space  Flight  Center  was  the 
initial  step  in  the  demise  of  the  development  of  any  passive  communications 
satellite  system.  The  Goddard  director  had  already  heavily  committed  his 
resources  to  the  development  of  an  active  system;  his  organization  was  thus 
reluctant  to  take  on  the  added  burden  of  the  passive  system,  which  many 
Goddard  engineers,  and  probably  Goddard  Director  Goett,  believed  would 
prove  inferior. 

Shotput 

One  of  the  responsibilities  taken  on  by  Langley  in  early  1959  was  the 
management  of  a  project  essential  to  Echo's  success:  Shotput.  The  purpose 
of  Shotput  was  "to  ensure  proper  operation  of  the  payload  package  at 
simulated  orbital  insertion" — in  other  words,  to  do  thorough  developmental 
testing  of  the  techniques  by  which  the  folded  Echo  balloon  would  be  ejected 
from  its  canister  and  inflated  in  space.  The  techniques  conceived  and 
refined  for  the  Sub-Satellite  and  the  12-foot  Beacon  satellite  were  almost 
totally  inapplicable  to  the  giant  Echo  balloon,  so  new  schemes  had  to  be 
perfected.  Only  some  of  the  critical  tests  could  be  made  on  the  ground 
because  a  vacuum  chamber  large  enough  to  simulate  the  complete  dynamics 
of  the  balloon  inflating  in  space  was  impractical  to  build.  The  only  option 
was  to  do  the  testing  in  the  actual  environment  of  space,  and  that  meant 
developmental  flight  tests.54 

The  importance  of  Shotput  to  Project  Echo's  ultimate  success  bears  wit- 
ness to  the  need  for  thorough  developmental  testing  prior  to  any  spaceflight 
program.  Before  NASA  researchers  risked  an  expensive  launch  of  a  precious 
piece  of  space  hardware,  they  made  sure  that  the  project  would  work  from 
start  to  finish.  Langley's  plan  was  to  flight-test  suborbital  Shotput  vehi- 
cles from  Wallops  Island,  then  conduct  as  many  orbital  launches  from  Cape 
Canaveral  as  needed  to  put  an  Echo  satellite  in  orbit  successfully.  For  the 
most  part,  that  plan  was  followed. 


the  part  of  Langley  working  on  Echo  was  really  working  for  the  Office  of  Space  Flight  Development  under 
Silverstein.  But  it  also  meant  that  the  Langley  Project  Echo  Task  Group  relied  not  on  Goddard,  but 
on  Langley's  Procurement  Division  for  its  funding.  See  Joseph  A.  Shortal,  A  New  Dimension:  Wallops 
Flight  Test  Range,  the  First  Fifteen  Years,  NASA  RP-1028  (Washington,  1978),  p.  688. 

179 


Spaceflight  Revolution 

One  of  the  most  difficult  technical  tasks  facing  Langley  researchers 
working  on  Project  Echo  was  designing  a  container  that  would  open  safely 
and  effectively  release  the  satelloon.  After  several  weeks  of  examining 
potential  solutions  to  this  problem,  the  Langley  engineers  narrowed  the 
field  of  ideas  to  five.  They  then  built  working  models  of  these  five  container 
designs,  and  12-foot-diameter  models  of  the  satellite  for  simulation  studies. 
With  help  from  Langley's  Engineering  Service  and  Mechanical  Service 
divisions,  the  Echo  group  built  a  special  41-foot-diameter  spherical  vacuum 
chamber  equipped  with  pressure-proof  windows.  There  the  dynamics  of 
opening  the  container  and  inflating  the  satelloon  could  be  studied  as  the 
satelloon  fell  to  the  bottom  of  the  tank.  To  observe  and  photograph 
the  explosive  opening  and  inflation  within  the  dark  chamber,  a  special 
lighting  rig  had  to  be  devised.  Employing  heavy  bulbs  enclosed  in  protective 
housings,  the  rig  ensured  that  in  the  short  time  the  test  required,  the  bulbs 
would  not  overheat  or  be  shattered  by  a  shock  wave.55 

The  container-opening  mechanism  that  eventually  resulted  from  these 
vacuum  tests  was  surely  one  of  the  oddest  explosive  devices  ever  contrived. 
The  container  was  a  sphere  that  opened  at  its  equator  into  top  and  bottom 
hemispheres.  The  top  half  fit  on  the  bottom  half  much  like  a  lid  fits  snugly 
atop  a  kitchen  pot.  The  joint  between  the  two  hemispheres,  therefore, 
formed  a  sliding  valve.  The  halves  had  to  move  apart  an  inch  or  two  before 
the  canister  was  actually  open.  It  was  in  this  joint  between  the  hemispheres 
that  the  charge  was  placed. 

The  charge  was  incased  in  a  soft  metal  tube  that  encircled  the  canister; 
in  cross  section,  the  tube  had  the  shape  of  a  sideways  V.  This  shape 
concentrated  the  blast  into  a  thin  jet  that  shot  out  the  mouth  of  the  V.  When 
the  charge  had  been  placed,  Langley  technicians  fastened  the  hemispheres 
of  the  container  together.  Because  even  minimal  pressure  remaining  inside 
the  canister  would  be  greater  than  that  in  space,  the  team  had  to  take  steps 
to  prevent  the  canister  from  blowing  apart  too  soon.  The  solution  was  to 
lace  fishing  line  through  eyelet  holes  in  the  hemispheres.  When  the  explosive 
charge  fired  out,  the  resulting  jet  cut  the  lacing  so  that  the  container  halves 
were  free  to  separate.  At  the  same  time,  pressure  from  the  charge  drove  the 
hemispheres  apart,  releasing  the  balloon. 

This  ingenious  arrangement  proved  successful  despite  its  inelegance.  So 
pleased  were  the  Langley  researchers  with  their  invention  that  they  were 
"somewhat  taken  aback"  when  visiting  scientists  and  engineers,  hearing 
descriptions  of  a  container-opening  mechanism  involving  such  crazy  things 
as  a  pot-lid  sliding  valve  and  a  lacing  made  of  fishing  line,  "thought  we  were 
joking."56 

As  challenging  as  the  opening  of  the  satelloon  container  was,  the  problem 
of  inflating  the  large  satelloon  without  bursting  it  was  even  more  vexing. 
O'Sullivan  once  explained  the  crux  of  the  matter:  "When  the  satelloon 
container  is  opened  to  release  the  satelloon  in  the  hard  vacuum  of  space, 
any  air  inside  the  folded  satelloon  or  outside  of  the  satelloon  between  its 

180 


The  Odyssey  of  Project  Echo 


A  technician  assigned  to  the  Project 
Echo  Task  Group  separates  the  two 
hemispheres  of  the  Echo  1  container 
for  inspection.  The  charge  that  freed 
the  balloon  was  placed  inside  of  a  ring 
encircling  the  canister  at  its  equator. 


L-64-6971 


folds  tends  to  expand  with  explosive  rapidity  and  rip  the  satelloon  to  pieces. 
But  this  understanding  of  the  problem  was  not  easily  acquired,  for  there  is 
no  vacuum  chamber  on  earth  big  enough  and  capable  of  attaining  the  hard 
vacuum  of  space,  in  which  the  ejection  and  complete  inflation  of  the  satelloon 
could  be  performed  and  the  process  photographed  with  high  speed  cameras 
to  detect  malfunctionings  of  the  process.  7 

Before  risking  the  launch  of  a  balloon  into  space,  the  Project  Echo  Task 
Group  determined  that  it  should  first  conduct  a  static  inflation  test  on 
the  ground  to  see  whether  the  100- foot-diameter  satelloon  would  assume 
a  spherical  shape  with  surface  conditions  sufficient  to  serve  as  a  passive 
communications  relay  satellite  between  two  distant  stations  on  the  surface 
of  the  earth.  To  make  the  static  inflation  tests,  Jesse  Mitchell  took  a  team 
of  engineers  to  nearby  Weeksville,  North  Carolina,  off  the  north  shore  of 
the  Albemarle  Sound,  where  a  cavernous  navy  blimp  hangar  big  enough  to 
inflate  the  Echo  balloon  to  full  size  stood  empty.  The  inflation  process  was 
slow,  taking  more  than  12  hours,  and  thus  did  not  offer  a  dynamic  simulation 
of  the  explosive  inflation  that  would  take  place  in  space;  however,  the  results 
did  reassure  everyone  that  the  balloon  would  work  as  a  communication  relay. 
As  Norm  Crabill,  present  at  the  Weeksville  tests,  explains,  "It  was  another 
one  of  the  tests  we  had  to  go  through  before  we  could  trust  the  design."58 

These  tests  also  demonstrated  that  the  original  balloon,  manufactured  by 
General  Mills,  was  seriously  defective.  When  the  balloon  was  inflated  in  the 
hangar,  the  triangular  panels,  or  gores,  began  coming  apart  at  the  seams. 


181 


Spaceflight  Revolution 


L-58-3598 


L-B 


L-F  L-58-3600 

Testing  Echo  1  's  inflation  (above)  in  the  navy  hangar  at  Weeksville  took  half  the 
day  but  proved  worth  the  trouble. 

182 


The  Odyssey  of  Project  Echo 


L-58-3603 

Langley  engineers  Edwin  Kilgore  (center),  Norman  Crabill  (right),  and  an  uniden- 
tified man  take  a  peek  inside  the  vast  balloon  during  inflation  tests. 


L-6 1-4603 

The  Echo  1  team  stand  in  front  of  their  balloon.    William  J.  O  'Sullivan  is  the  tall 
man  at  center;  Walter  Bressette  is  to  his  left. 


183 


Space/light  Revolution 

^Pt  \       m 


L-60-490  L-60-485 

Langley  technicians  Will  Taub  and  James  Miller  (left)  prepare  to  spin-balance  the 
final  stage  of  the  Shotput  launch  vehicle.  The  ABL  X248  motor  sits  on  the  spin 
table;  the  balloon- containing  canister  is  at  the  top.  When  the  Shotput  was  fully 
prepared  for  launch  (right),  a  pencil-shaped  shroud  was  fitted  over  the  payload. 

Another  manufacturer,  the  G.  T.  Schjeldahl  Company  of  rural  Northfield, 
Minnesota,  had  a  glue  perfect  for  sealing  the  seams,  so  General  Mills  hired 
the  company  to  construct  a  second  sphere.  The  proud  Schjeldahl  Company 
provided  all  subsequent  inflatable  spheres  for  NASA.59 

Although  the  ground  testing  proved  critical,  the  only  sure  way  to  test 
the  inflation  process  was  to  launch  the  sphere  in  its  container  up  to  satellite 
altitude.  To  do  this,  members  of  the  Project  Echo  Task  Group  designed  the 
special  two-stage  test  rocket  called  "Shotput."  This,  they  thought,  was  the 
perfect  nickname  for  a  vehicle  that  would  essentially  hurl  a  big  ball  out  of 
the  atmosphere. 

Shotput's  first  stage  was  the  Sergeant  XM-33;  its  second  stage  was  the 
ABL  (Allegheny  Ballistics  Laboratory)  X248.  The  latter  also  served  as  the 
third  stage  of  the  Douglas  Thor-Delta,  soon  to  be  one  of  the  United  States' 
primary  satellite  launchers.  Although  the  test  program's  main  purpose  was 
to  check  out  the  Echo  satelloon,  testing  this  part  of  the  Thor-Delta  became 
a  critical  secondary  task.  The  ABL  X248  stage  included  a  solid-propellant 
rocket  motor  designed  to  achieve  proper  satellite  velocity  and  altitude.  The 
motor  was  spin-stabilized,  so  after  it  had  burned  out  and  the  motor-satellite 
complex  had  entered  orbit,  the  whole  ensemble  had  to.  be  de-spun  before 
the  satellite  could  be  separated.  To  accomplish  that,  engineers  fashioned  a 


184 


The  Odyssey  of  Project  Echo 

weighted  mechanism  known  as  a  "yo-yo,"  which  stopped  the  spinning  and 
allowed  the  container  to  separate  safely.60 

Solving  the  problems  of  the  launch  vehicle  was  as  difficult  as  solving  the 
problems  of  the  balloon.  Norm  Crabill  traveled  back  and  forth  to  Tulsa 
several  times  to  understand  the  detailed  design  of  the  Delta  third  stage. 
(O'Sullivan  once  tried  to  remove  Crabill  from  the  project  because  he  thought 
the  young  Langley  engineer  did  not  know  enough  to  be  in  charge  of  the 
development  of  the  Shotput  test  vehicle.)  Crabill  and  his  assistant  Robert 
James  intensely  studied  the  forces  and  moments  (i.e.,  the  aerodynamic 
tendency  to  cause  rotation  about  a  point  or  axis)  on  the  Shotput  vehicle 
as  it  shot  up  and  out  of  the  atmosphere,  spun  its  way  to  altitude,  and  de- 
spun  for  payload  separation.  The  researchers  had  to  assimilate  in  just  a  few 
months  what  amounted  to  an  advanced  course  in  aerodynamics  and  missile 
dynamics,  but  finally,  after  numerous  analytical  studies  and  simulations, 
Crabill  and  his  helpers,  one  by  one,  solved  the  problems  of  launching 
Shotput  1. 


A  Burst  Balloon 

By  the  second  Project  Echo  planning  meeting,  Langley  had  established 
a  schedule  for  four  Shotput  tests.  (Five  Shotput  launches  would  in  fact 
occur;  the  last  would  take  place  on  31  May  1960.)  Everyone  inside  NASA, 
including  the  interested  parties  at  Goddard,  agreed  that  the  responsibility 
for  managing  Shotput  and  launching  the  vehicles  from  Wallops  Island  should 
remain  in  Langley's  hands. 

Unfortunately,  keeping  their  brainchild  at  home  did  not  assure  total 
success.  As  described  in  this  chapter's  opening,  the  launch  of  Shotput  1  on 
28  October  1959  started  off  well,  but  far  above  the  "sensible"  atmosphere, 
upon  inflation,  the  big  balloon  blew  up.  Instead  of  a  respectable  scientific 
experiment,  Echo  looked  more  like  a  Fourth  of  July  skyrocket.61 

Despite  the  initial  subterfuge  of  calling  the  test  a  success  and  omitting 
any  mention  of  the  balloon's  explosion,  the  group's  spokesmen  finally  con- 
fessed under  pressure  from  the  media  and  with  great  embarrassment,  "that  it 
was  not  supposed  to  work  that  way."  For  several  weeks  thereafter,  everyone 
at  Langley  became  an  authority  on  inflatable  satellites,  telling  O' Sullivan's 
associates  (not  daring  to  tell  O'Sullivan  himself,  as  he  was  known  to  have 
little  charity  for  opinions  contrary  to  his  own)  what  had  caused  the  explo- 
sion and  how  to  fix  it.  Many  of  these  "self-appointed  experts"  demanded 
to  be  heard.  The  Project  Echo  Task  Group  accommodated  most  of  them, 
trying  to  keep  in  mind  that  "all  of  these  people  meant  well  and  were  trying 
to  help."62  Thereafter,  NASA  headquarters  also  announced  the  Shotput 
tests  well  ahead  of  time,  so  that  everyone  on  the  East  Coast  could  watch 
and  enjoy  them.  However,  if  everything  went  right  with  the  balloon,  the 
spectacular  fireworks  would  not  occur. 

185 


Space/light  Revolution 

A  500-inch  focal-length  photographic  camera  set  up  on  the  beach  at 
Wallops  Island  had  taken  pictures  of  Shotput  1  as  the  balloon  inflated  and 
blew  up,  but  even  with  these  data  a  team  from  the  Project  Echo  Task  Group 
spent  several  weeks  trying  to  confirm  why  the  balloon  had  burst  apart.  Some 
researchers  believed  that  the  water  used  to  help  inflate  the  balloon  had  been 
the  culprit.  Like  other  volatile  liquids,  water  will  boil  explosively  in  the  zero 
pressure  of  space.  It  was  "entirely  conceivable  that  the  elastic  containers 
in  which  the  water  was  carried  inside  the  satellite  might  have  leaked  or 
ruptured  during  launch,  and  thus  did  not  release  the  water  at  a  slow  and 
controlled  rate  as  planned,  to  give  a  slow  and  gentle  inflation."63  Leaked 
water  could  easily  have  produced  an  explosion. 

To  ensure  that  the  water  inflation  system  would  not  malfunction  in  the 
future,  the  team,  led  by  Walter  Bressette,  switched  to  benzoic  acid,  a  solid 
material  that  underwent  sublimation — that  is,  transformation  from  a  solid 
state  directly  to  a  vapor.  With  such  a  material,  conversion  to  a  gas  would  be 
limited  by  the  rate  at  which  it  would  absorb  heat  from  the  sun.  In  essence, 
it  would  "gas  off"  slowly,  not  instantaneously. 

Researchers  worried  that  another  contributor  to  the  explosion  may  have 
been  residual  air,  which  the  payload  engineers  had  intentionally  left  inside 
the  folds  of  the  balloon  as  an  inflation  agent.  Langley's  O'Sullivan  once 
explained:  "When  the  satelloon  container  is  opened  to  release  the  satelloon 
in  the  hard  vacuum  of  space,  any  air  inside  the  folded  satelloon  or  outside 
the  satelloon  between  its  folds  tends  to  expand  with  explosive  rapidity 
and  rip  the  satelloon  to  pieces."64  To  remove  all  residual  air  from  future 
deployments,  the  engineers  made  over  300  little  holes  in  the  balloon  to  allow 
the  air  to  escape  after  the  balloon  was  folded.  Once  the  balloon  was  packed, 
the  canister  was  placed,  slightly  open,  in  a  vacuum  tank.  When  its  internal 
pressure  had  been  reduced  to  near  zero,  the  canister  was  closed,  and  an 
O-ring  maintained  the  internal  vacuum. 

Finally,  to  better  identify  deployment  problems,  the  engineers  put  a  red 
fluorescent  powder  into  the  folded-up  balloon.  If  the  balloon  ruptured  during 
ejection  or  inflation  in  subsequent  tests,  the  powder  would  blow  out  and  leave 
a  trail  that  could  be  instantly  seen  around  the  satellite  even  from  the  earth. 

Four  Shotputs  were  launched  before  the  Langley  researchers  were  satis- 
fied that  Echo  would  work.  The  second  shot,  on  16  January  1960,  failed 
because  of  a  problem  with  Crabill's  beloved  launch  vehicle.  The  yo-yo  de- 
spin  system  of  the  Shotput  second  stage  did  not  deploy  properly,  and  the 
payload  separated  from  the  burned-out  second  stage  still  spinning  at  250 
rpm.  When  the  red  dye  appeared  in  the  sky,  it  was  clear  that  the  de-spin 
failure  had  caused  the  balloon  to  tear  while  inflating.  Following  this  test, 
no  more  serious  problems  with  the  launch  vehicle  occurred;  there  were  only 
problems  with  the  test  balloon.  On  the  third  shot  five  weeks  later,  on  27 
February,  the  balloon  tore  and  developed  a  hole,  although  not  before  Bell 
Labs  was  able  to  use  the  sphere  to  transmit  voice  signals  from  its  headquar- 
ters in  Holmdel,  New  Jersey,  to  Lincoln  Labs  in  Round  Hill,  Massachusetts. 

186 


The  Odyssey  of  Project  Echo 

A  successful  shot  took  place  on  1  April,  but  the  tests  were  still  incomplete 
as  the  satellite  did  not  yet  carry  any  of  the  tracking  beacons  that  the  final 
version  would  have.65  (Because  Echo's  orbit  would  not  be  geostationary — 
hovering  over  the  same  spot  on  earth  24  hours  a  day — such  devices  were 
required  to  enable  ground  crews  to  track  the  balloon.) 

The  Project  Echo  Task  Group,  however,  believed  that  "they  were  over 
the  hump"  and  that  the  next  step  was  to  move  beyond  Shotput,  put  the 
completely  equipped  100-foot  passive  reflector  balloon  on  the  Thor-Delta, 
and  attempt  a  launch.  The  scheduled  launch  date  of  "TD  No.  1"  from  Cape 
Canaveral  was  13  May  1960,  just  over  a  month  away.  Unlike  the  Shotput 
tests,  whose  ABL  X248  carried  the  test  balloons  only  to  200  to  250  miles 
above  the  surface,  the  much  more  powerful  92-foot-high  Thor-Delta  would 
ultimately  take  the  balloon  to  an  orbit  1000  miles  above  the  earth.  From 
there,  the  enormous  Echo  would  be  visible  to  people  all  around  the  world.66 


"Anything's  Possible!" 

The  Echo  balloon  was  perhaps  the  most  beautiful  object  ever  to  be  put 
into  space.  The  big  and  brilliant  sphere  had  a  31,416-square-foot  surface  of 
Mylar  plastic  covered  smoothly  with  a  mere  4  pounds  of  vapor-deposited 
aluminum.  All  told,  counting  30  pounds  of  inflating  chemicals  and  two  11- 
ounce,  3/8-inch-thick  radio-tracking  beacons  (packed  with  70  solar  cells  and 
5  storage  batteries),  the  sphere  weighed  only  132  pounds. 

For  those  enamored  with  its  aesthetics,  folding  the  beautiful  balloon  into 
its  small  container  for  packing  into  the  nose  cone  of  a  Thor-Delta  rocket 
was  somewhat  like  folding  a  large  Rembrandt  canvas  into  a  tiny  square  and 
taking  it  home  from  an  art  sale  in  one's  wallet.  However,  the  folding  of  the 
balloon  posed  more  than  aesthetic  problems.  The  structure  not  only  had  to 
fit  inside  the  spherical  canister  but  also  had  to  unfold  properly  for  inflation. 

The  technique  for  folding  the  100- foot  inflatable  balloons  evolved  from 
a  classic  "Eureka"  moment.  One  morning  in  1960,  Ed  Kilgore,  the  man 
in  the  Engineering  Service  Division  responsible  for  the  Shotput  test  setups, 
received  a  call  from  Schjeldahl,  the  manufacturer  of  the  Echo  balloons.  The 
company's  technicians  were  having  a  terrible  time:  not  only  were  they  unable 
to  fit  the  balloon  into  its  canister,  they  couldn't  even  squeeze  it  into  a  small 
room. 

Kilgore  mulled  over  the  problem  all  day  and  part  of  the  night,  but  it 
wasn't  until  the  next  morning  that  he  happened  upon  a  possible  solution. 
"It  was  raining,"  he  recalls,  "and  as  I  started  to  leave  for  work,  my  wife 
Ann  arrived  at  the  door  to  go  out  as  I  did.  She  had  her  plastic  rain  hat 
in  her  hand.  It  was  folded  in  a  long  narrow  strip  and  unfolded  to  a  perfect 
hemisphere  to  fit  the  head."  Recognizing  the  importance  of  his  accidental 
discovery,  Kilgore  told  his  wife  that  she  "would  have  to  use  an  umbrella  or 
get  wet  because  I  needed  that  rain  hat."67 

187 


Spaceflight  Revolution 

At  Langley,  Kilgore  gave  the  hat  to  Austin  McHatton,  a  talented 
technician  in  the  East  Model  Shop,  who  had  full-size  models  of  its  fold 
patterns  constructed.  Kilgore  remembers  that  a  "remarkable  improvement 
in  folding  resulted."  The  Project  Echo  Task  Group  got  workmen  to 
construct  a  makeshift  "clean"  room  from  two-by-four  wood  frames  covered 
with  plastic  sheeting.  In  this  room,  which  was  150  feet  long  and  located 
in  the  large  airplane  hangar  in  the  West  Area,  a  small  group  of  Langley 
technicians  practiced  folding  the  balloons  for  hundreds  of  hours  until  they 
discovered  just  the  right  sequence  of  steps  by  which  to  neatly  fold  and  pack 
the  balloon.  For  the  big  Echo  balloons,  this  method  was  proof-tested  in  the 
Langley  60-foot  vacuum  tank  as  well  as  in  the  Shotput  flights.68 

Whether  the  packed  balloon  would  have  deployed  properly  on  13  May 
1961,  no  one  will  ever  know  because  once  again  the  launch  vehicle  failed.  The 
second  stage  of  the  Delta  refused  to  fire,  and  the  whole  rocket  dropped  into 
the  Atlantic.  The  vehicle's  manufacturer,  Douglas,  blamed  a  malfunctioning 
accelerometer.69 

By  this  point,  the  program  had  experienced  a  total  of  seven  failures 
including  those  of  the  two  small  pre-Echo  test  satelloons.  For  a  test 
conducted  on  31  May,  the  team  returned  to  using  the  Shotput  launcher. 
With  tracking  beacons  aboard,  the  balloon  deployed  successfully,  which 
helped  the  NASA  engineers  rally  from  their  recent  setback. 

Still,  critics  continued  to  doubt  the  overall  Echo  concept.  Some  swore 
that  even  if  the  satelloon  ever  got  up  into  space  and  inflated  properly, 
micrometeorites  would  puncture  its  skin,  thus  destroying  the  balloon  within 
hours.  Not  so,  the  Langley  engineers  countered.  The  idea  was  to  pressurize 
the  balloon  just  enough  to  overstress  the  material  slightly,  thus  causing  it 
to  take  on  a  permanent  set.  Even  after  its  internal  pressure  had  dwindled 
to  nothing,  the  balloon  would  retain  its  shape.  Because  the  outer  skin  was 
not  extremely  rigid — it  was  in  engineering  slang  "dead-soft" — it  could  be 
punctured  by  a  small  meteorite  and  still  not  shatter.  Finally,  a  study  by 
Bressette  showed  that  micrometeorites  would  erode  less  than  one-millionth 
of  the  surface  area  a  day.  If  only  a  launching  and  deployment  would  go  right, 
the  satelloon 's  sublimating  solid-pressurization  system  would  work  long 
enough  to  enable  engineers  to  conduct  their  communications  experiment. 

The  next  time  around,  the  launch  finally  did  go  right.  At  5:39  a.m. 
on  12  August  1960,  Thor-Delta  No.  2  blasted  into  the  sky  from  launchpad 
17  at  Cape  Canaveral,  taking  its  balloon  into  orbit.  A  few  minutes  later, 
the  balloon  inflated  perfectly.  At  7:41  a.m.,  still  on  its  first  orbit,  Echo  1 
relayed  its  first  message,  reflecting  a  radio  signal  shot  aloft  from  California 
to  Bell  Labs  in  New  Jersey.  "This  is  President  Eisenhower  speaking,"  the 
voice  from  space  said.  "This  is  one  more  significant  step  in  the  United 
States'  program  of  space  research  and  exploration  being  carried  forward 
for  peaceful  purposes.  The  satellite  balloon,  which  has  reflected  these 
words,  may  be  used  freely  by  any  nation  for  similar .  experiments  in  its 
own  interest."71  After  the  presidential  message,  NASA  used  the  balloon  to 

188 


The  Odyssey  of  Project  Echo 

transmit  two-way  telephone  conversations  between  the  east  and  west  coasts. 
Then  a  signal  was  transmitted  from  the  United  States  to  France  and  another 
was  sent  in  the  opposite  direction.  During  the  first  two  weeks,  the  strength 
of  the  signal  bounced  off  Echo  1  remained  within  one  decibel  of  Langley's 
theoretical  calculations. 

The  newspapers  sounded  the  trumpets  of  success:  "U.S.  Takes  Big  Jump 
in  Space  Race";  "U.S.  Orbits  World's  First  Communications  Satellite:  Could 
Lead  to  New  Marvels  of  Radio  and  TV  Projection" ;  "Bright  Satellite  Shines 
Tonight."  So  eager  was  the  American  public  to  get  a  glimpse  of  the  balloon 
that  NASA  released  daily  schedules  telling  when  and  where  the  sphere  could 
be  seen  overhead.72 

For  the  engineers  from  Langley  who  were  lucky  enough  to  be  at  Cape 
Canaveral  for  the  launch,  this  was  a  heady  time.  Norm  Crabill  remembers 
hearing  the  report  that  "Australia's  got  the  beacon,"  meaning  that  the 
tracking  station  on  that  far-off  continent  had  picked  up  the  satellite's  beacon 
signal.  To  this  day,  Crabill  "gets  goose  bumps  just  thinking  about  that 
moment."  He  remembers  thinking,  "Anything's  possible!"73  After  all,  the 
space  age  had  arrived,  and  in  a  sense,  anything  was. 


Reflections 

Out  of  the  seven  failures,  including  the  scintillating  bits  of  Shotput  1, 
NASA  built  a  successful  communications  satellite  program,  which  entranced 
the  public.  After  a  fully  operational  Echo  balloon  was  launched  into  orbit 
on  12  August  1960,  the  big  silver  satelloon  continued  to  orbit  for  eight 
years,  not  falling  back  to  earth  until  May  1968.  For  that  entire  period,  the 
satelloon  served  as  a  significant  propaganda  weapon  for  the  United  States. 
It  was  a  popular  symbol  of  the  peaceful  and  practical  uses  of  space  research, 
especially  in  the  early  1960s  when  the  country  still  seemed  so  far  behind  the 
Soviets. 

During  its  long  sojourn  in  space,  Echo  1  proved  to  be  an  exceptionally 
useful  tool.  First  and  foremost,  by  enabling  numerous  radio  transmissions 
to  be  made  between  distant  ground  stations,  it  demonstrated  the  feasibility 
of  a  global  communications  system  based  on  satellites.  The  rapid  and 
successful  development  of  worldwide  communications  in  the  1960s  depended 
upon  this  demonstration.  Echo  1  also  proved  wrong  the  experts  who 
said  that  the  satelloon,  after  losing  internal  pressure  because  of  meteoroid 
punctures,  would  collapse  from  external  pressure.  Echo  actually  retained  its 
sphericity  far  longer  than  expected,  the  external  pressures  (including  solar 
radiation)  doing  more  to  change  the  orbit  of  the  satelloon  than  to  collapse 
it.74  In  addition,  NASA  researchers  studied  the  long-term  durability  of  the 
unique  metallized  plastic  of  the  Echo  balloons  (an  Echo  2  was  launched 
in  1964)  in  order  to  evaluate  similar  materials  proposed  for  components 
of  other  spacecraft,  including  early  versions  of  a  manned  space  station. 

189 


Space/light  Revolution 

Finally,  Echo  permitted  scientists  to  demonstrate  a  triangulation  technique 
for  determining  the  distance  between  various  points  on  the  earth's  surface, 
thus  improving  mapping  precision.  The  satelloon  also  served  as  a  test  target 
for  the  alignment  and  calibration  of  a  number  of  new  radars. 

However,  the  Echo  satelloon  demonstrated  some  critical  limitations.  As 
it  turned  out,  the  balloon's  shape  was  a  poor  passive  reflector.  When  hit 
with  a  plane  wave  (a  wave  in  which  the  wave  fronts  lay  in  a  fixed  line 
parallel  to  the  direction  of  the  propagation) ,  the  sphere  tended  to  propagate 
the  wave  outward  and  reflect  it  as  a  divergent  wave.  Echo  did  an  adequate 
job  reflecting  radio  signals  transmitted  from  the  ground,  but  it  did  a  poor 
job  of  focusing  them.  As  a  result,  everybody  received  some  of  the  reflected 
signal,  but  nobody  received  very  much  of  it. 

Thus,  the  Echo  balloon  served  primarily  as  a  demonstration  model, 
showing  how  a  simple  passive  comsat  might  work.  For  actual  operations,  a 
better  concept,  which  NASA  and  the  companies  involved  in  the  development 
of  commercially  viable  satellites  were  already  working  on,  was  satellites  that 
could  communicate  with  active  electronics.  Because  the  force  or  intensity 
of  a  radio  wave  is  weakened  or  attenuated  by  the  square  of  the  distance  it 
must  travel  through  space,  an  active  communications  system  has  a  distinct 
advantage  over  the  passive  system:  the  active  system  receives  the  signal  at 
one  frequency  and  retransmits  it  at  another.  In  effect,  the  signal  travels 
the  earth-to-satellite  distance  only  once;  the  signal  in  a  passive  system  must 
travel  the  distance  twice,  and  thus  is  more  seriously  attenuated,  as  the  fourth 
power  of  the  distance.75 

The  demise  of  the  passive  satellite  communication  system  and  the 
emergence  of  the  active  communication  system,  however,  also  need  to  be 
explained  in  the  context  of  broader  economic,  political,  and  institutional 
realities.  In  the  beginning,  satellite  communications  research  was  funded  by 
the  U.S.  government  because  the  military  required  worldwide  instantaneous 
communications  for  national  defense.  The  military  was  interested  in  the 
passive  system  because  it  could  not  be  electronically  jammed.  On  the  other 
hand,  the  private  telecommunications  companies  were  not  yet  interested 
in  a  satellite  communications  system,  partly  because  they  were  investing 
heavily  in  ground  relay  stations  and  under-the-ocean  cable  systems  and 
partly  because  their  engineers  strongly  suspected  that  radio  signals  passing 
through  the  earth's  ionosphere  would  be  seriously  weakened  in  intensity. 

In  an  ironic  twist  of  fate,  given  the  history  that  was  to  follow,  the  Echo 
balloon  actually  changed  this  thinking  about  the  potential  of  a  communica- 
tions system  in  space.  When  Echo  1  demonstrated  that  the  ionosphere  was 
not  going  to  be  a  problem  in  satellite  communications,  the  private  sector 
jumped  on  the  bandwagon  and  demanded  their  own  geosynchronous  satellite 
system,  but  the  private  sector  wanted  an  active  rather  than  a  passive  system. 
Many  of  the  companies  involved  had  the  technical  knowledge  to  develop  an 
active  system,  but  this  was  not  the  sole  reason  for  their  interest;  money  was 
another  factor.  Individual  companies  could  charge  for  sending  a  message 

190 


The  Odyssey  of  Project  Echo 

through  the  system  since  they  would  own  the  frequency  channels  located 
in  the  particular  satellites.  As  Bressette  comments,  "The  active  communi- 
cations people  used  the  capitalistic  approach  for  the  success  of  a  project: 
'Does  it  make  money?'  On  the  other  hand,  the  few  people  [like  Bressette] 
who  were  promoting  the  passive  system  were  thinking  more  democratically. 
Just  think  how  inexpensive  satellite  communications  would  be  today,  if  it 
were  possible  to  replace  all  the  active  communications  satellites  with  just 
three  nonmaintenance  passive  satellites."76 

To  overcome  the  problem  of  radio- wave  attenuation  from  geosynchronous 
orbit,  the  Echo  satelloon  would  need  to  be  many  times  larger.  Since  the 
technology  did  not  exist  in  the  early  1960s  to  put  such  a  large  satelloon 
in  orbit,  even  the  military  began  to  opt  for  the  active  system.  Given  the 
logistical  difficulties  and  tremendous  costs  of  flying  high-altitude  radio-relay 
stations  over  the  oceans  inside  giant  aircraft  such  as  the  B-52,  the  DOD 
was  excited  by  the  promise  of  a  space-based  geosynchronous  system,  which 
could  move  the  high-altitude  radar-relay  flights  into  a  backup  position. 

For  its  part,  NASA  Langley  did  not  easily  give  up  on  the  passive 
system.  Between  1963  and  1965,  in  conjunction  with  Goodyear  Aerospace 
Corporation,  a  team  of  Langley  researchers  performed  a  study  showing  that 
as  little  as  a  10°  segment  cut  from  a  very  large  sphere  in  geosynchronous 
orbit  would  be  satisfactory  for  passive  communications  between  two  remote 
stations  on  earth.77 

William  J.  O'Sullivan's  original  concept  for  the  inflatable  satellite,  which 
was  to  serve  as  an  air-density  experiment,  was  not  forgotten.  The  long-term 
orbiting  of  the  satelloon  allowed  scientists  to  measure  accurately,  for  the 
first  time,  the  density  of  the  air  in  the  far  upper  atmosphere.  With  the 
data  came  some  important  insights  into  the  effects  of  solar  pressure  on  the 
motion  of  satellites,  information  that  was  helpful  in  predicting  the  behavior 
and  lifetime  of  future  satellites.  Several  versions  of  the  basic  experiment 
were  carried  out  at  a  high  altitude  over  both  low  and  high  latitudes  of  the 
earth's  surface  as  part  of  four  Explorer  missions:  Explorer  9  in  February 
1961,  Explorer  19  in  December  1963,  Explorer  24  in  November  1964,  and 
Explorer  39  in  August  1968.  NASA  launched  these  satellites  at  regular 
intervals  to  provide  continual  coverage  of  density  variation  throughout  a 
solar  cycle.  With  the  findings  from  these  worthwhile  missions,  scientists 
were  able  to  improve  their  measurements  of  atmospheric  density,  better 
understand  variations  in  density  caused  by  variations  in  the  solar  cycle,  and 
study  the  MPD- related  phenomena  of  geomagnetically  trapped  particles  and 
their  down-flux  into  the  atmosphere. 

O'Sullivan's  1956  concept  led  to  not  just  a  single  experiment  but  an  entire 
program  of  inflatable  satellites,  all  of  which  involved  Langley  in  some  central 
way.  This  program  included,  in  addition  to  the  Echo  satelloons  and  the  air- 
density  Explorers,  a  Langley-managed  passive  geodetic  satellite  known  as 
"Pageos"  (Passive  Geodetic  Earth-Orbiting  Satellite).  A  Thor-Agena  lifting 
off  from  the  Pacific  Missile  Range  in  June  1966  took  Pageos  1,  which  was 

191 


Space/light  Revolution 


L-64-9269 

This  satellite,  Explorer  24,  was  a  12- foot- diameter  inflatable  sphere  developed 
by  an  engineering  team  at  Langley.  It  provided  information  on  complex  solar 
radiation/ air- density  relationships  in  the  upper  atmosphere. 


very  similar  to  Echo  1,  into  a  near  polar  orbit  some  200  nautical  miles  above 
the  earth.  This  orbit  was  required  by  the  U.S.  Coastal  and  Geodetic  Survey 
to  use  the  triangulation  technique  developed  from  Echo  1  for  determining 
the  location  of  38  points  around  the  world.  More  than  five  years  and  the 
work  of  12  mobile  tracking  stations,  which  waited  for  favorable  weather 
conditions  during  a  few  minutes  of  twilight  each  evening,  were  required  to 
complete  the  project.  Finally,  the  geodetic  experts  were  able  to  fix  the  38 
points  into  a  grid  system  helpful  in  determining  the  precise  location  of  the 
continents  relative  to  each  other.  Some  of  this  information,  that  which  was 
not  classified  as  secret,  enabled  the  U.S.  scientific  community  to  determine 
geometrically  the  shape  and  the  size  of  the  earth.  This,  in  turn,  was  useful  to 
scientists  studying  the  theory  of  continental  drift.  Data  that  the  U.S.  Army 
Map  Service  classified  as  secret  proved  helpful  to  U.S.  military  planners 
concerned  with  the  accuracy  of  intercontinental  ballistic  missiles.  Thus, 
although  initially  conceived  to  tell  us  about  the  upper  atmosphere,  NASA's 
inflatable  satellite  program  told  us  perhaps  even  more  about  the  military 
buildup  here  on  earth.  9 

O'Sullivan  became  one  of  NASA's  most  highly  publicized  scientists.  In 
December  1960,  the  U.S.  Post  Office  Department  issued  a  commemorative 
4-cent  stamp  in  honor  of  his  beloved  Echo  balloon.  For  his  concept  of  the 
inflatable  space  vehicle,  NASA  would  award  him  one  of  its  distinguished 
service  medals,  in  addition  to  $5000  cash.  In  1962,  O'Sullivan  would  appear 
as  a  guest  on  the  popular  TV  game  show  "What's  My  Line?";  all  four  of 

192 


The  Odyssey  of  Project  Echo 


Hanging  from  the  ceiling  of  the 
Weeksville  blimp  hangar  like  a 
shiny  Christmas  tree  ornament, 
Langley  's  Pageos  satelloon  was  vir- 
tually identical  to  Echo  1. 


L-65-6894 


the  celebrity  panelists  correctly  picked  him  from  the  lineup  as  the  father  of 
the  Echo  satelloons. 

As  is  nearly  always  the  case  in  the  history  of  a  large-scale  technological 
development,  however,  many  other  individuals,  mostly  overlooked,  deserved 
a  significant  share  of  the  credit.  Jesse  Mitchell  was  one  of  those  individ- 
uals. In  late  1959,  Mitchell,  who  had  been  responsible  for  the  program 
development  plan  for  Echo  1,  left  Langley  for  a  special  assignment  on  an 
important  space  advisory  committee  chaired  by  Dr.  James  Killian,  Presi- 
dent Eisenhower's  special  assistant  for  science  and  technology.  After  this 
assignment,  Mitchell  became  the  head  of  the  Geophysics  and  Astronomy 
Division  at  NASA  headquarters.  In  following  years,  his  office  funded  the 
last  three  air-density  satellites  and  the  Langley-managed  Pageos  geodetic 
survey  satellites. 


The  Hegemony  of  Active  Voice 

Project  Echo  continued  for  several  more  years.    In  1962,  Langley  en- 
gineers staged    "Big  Shot" — two  space  deployment   tests  of  the   Echo  2 


193 


Spaceflight  Revolution 

balloon.*  The  first  test  was  a  disaster,  with  the  balloon  tearing  apart 
because  of  a  structural  load  problem.  The  second  test  was  a  success.  Echo  2 
was  launched  into  orbit  in  1964,  serving,  like  its  predecessor,  as  a  passive 
communications  relay.  By  the  mid-1960s,  however,  the  active  satellite  had 
proved  itself  the  better  method  for  communications  in  space.  In  July  1962,  a 
little  more  than  two  years  after  the  launch  of  Echo  1  and  some  20  years  after 
the  publication  of  Arthur  C.  Clarke's  speculative  essay  on  the  potential  of 
"extraterrestrial  relays,"  NASA  had  launched  its  first  active  communications 
satellite,  Telstar  1.  This  experimental  "comsat,"  which  belonged  to  AT&T, 
sent  the  first  direct  television  signals  ever  between  two  continents  (North 
America  and  Europe).  In  December  1962,  while  Langley  and  Goddard 
were  still  quarreling  over  what  to  do  with  Echo  2,  NASA's  own  Relay  1 
satellite  went  into  action.  Within  days,  Relay  1,  which  was  developed  at 
NASA  Goddard,  was  transmitting  civilian  television  broadcasts  between 
the  United  States  and  Europe.  When  TV  viewers  saw  astronaut  L.  Gordon 
Cooper  being  recovered  from  his  capsule  on  16  May  1963  at  the  end  of  the 
last  Mercury  mission,  they  were  seeing  a  signal  from  Relay  I.80 

The  age  of  the  active  comsat  had  arrived,  and  with  it  came  a  revolution 
in  telecommunications  that  would  have  an  enormous  impact  worldwide.  On 
25  February  1963,  NASA  announced  that  it  was  canceling  its  plans  for 
any  advanced  passive  communications  satellites  beyond  Echo  2  and  cutting 
off  funding  for  several  feasibility  study  contracts  aimed  at  determining  the 
best  shape,  structure,  and  materials  of  future  communications  balloons  in 
space.  In  light  of  the  formation  of  the  national  Communications  Satellite 
Corporation  (ComSatCorp),  the  space  agency  instead  would  focus  its  efforts 
on  the  development  of  synchronous-orbit  active  satellites.81 

The  next  American  active  comsat,    Telstar  2,  went  into  space  in  May 

1963,  which  was  still  before  the  launch  of  Echo  2.    Telstar  2  sent  the  first 
color  television  pictures  across  the  Atlantic  Ocean.  On  22  November  1963, 
NASA's  Relay  1  was  scheduled  to  transmit  color  television  pictures  across 
the  Pacific.     An  audience  in  Japan  waited  to  see  a  ceremonial  meeting 
between  NASA  Administrator  James  E.  Webb  and  the  Japanese  ambassador 
in  Washington.  The  audience  in  Tokyo  was  also  supposed  to  receive  a  taped 
greeting  from  President  Kennedy;  instead,  Relay  1  transmitted  the  shocking 
news  of  his  assassination.  Thanks  to  Relay  2,  which  was  launched  in  January 

1964,  TV  viewers  were  able  to  witness  Pope  Paul  VFs  visit  later  that  year 


The  management  of  Big  Shot  and  Echo  2  proved  more  quarrelsome  than  Shotput  and  Echo  1. 
Langley  and  Goddard  personnel  disagreed  strongly  about  many  engineering  details  and  fought  over 
budgetary  and  procurement  matters.  The  Langley  engineers  were  angry  that  Goddard  officials  were 
in  charge  of  Echo  when  Langley  was  doing  the  basic  planning  leading  to  launch.  Goddard's  satellite 
experts,  on  the  other  hand,  were  already  involved  in  the  development  of  active  electronic  comsats  and 
were  not  much  interested  in  improving  the  performance  of  passive  reflectors.  Thus,  the  tug-of-war 
between  Langley  and  Goddard  was  more  than  a  turf  battle;  it  was  a  technical,  debate  between  advocates 
of  passive  and  active  satellites. 

194 


The  Odyssey  of  Project  Echo 

to  the  Middle  East  as  well  as  Soviet  Premier  Nikita  Khrushchev's  tour  of 
Poland.  Thanks  to  another  NASA-sponsored  communications  satellite,  the 
Hughes  Aircraft-developed  Syncom  3,  Americans  enjoyed  live  TV  coverage 
of  the  Olympic  games  taking  place  on  the  other  side  of  the  world  in  Tokyo.  2 

In  1964,  10  nations  (plus  the  Vatican)  formed  the  International  Telecom- 
munications Satellite  Consortium,  or  Intelsat.  In  the  next  12  years,  Intelsat 
built  something  close  to  the  integrated  system  of  global  communications 
that  Arthur  Clarke  had  suggested.  By  the  late  1960s,  Intelsat's  member- 
ship included  80  countries.  Individual  nations  owned  and  operated  their 
own  ground  stations  and  reaped  dividends  in  proportion  to  their  investment 
shares,  while  a  large,  new  American  joint-stock  company,  ComSatCorp, 
whose  operations  were  private  but  heavily  subsidized  by  the  U.S.  govern- 
ment, managed  the  financial  and  operations  end  of  the  satellite  communica- 
tions system.  (NASA  simply  launched  the  satellites  and  was  reimbursed  for 
its  costs.)  By  the  early  1970s,  Intelsat's  sophisticated  network  was  enabling 
rapid  long-distance  telephoning  and  distribution  of  TV  programs  as  never 
before.  One  NASA  historian  has  written,  "Before  these  satellites  existed, 
the  total  capability  for  transoceanic  telephone  calls  had  been  500  circuits; 
in  1973  the  Intelsat  satellites  alone  offered  more  than  4000  transoceanic  cir- 
cuits. Real-time  TV  coverage  of  events  anywhere  in  the  world — whether 
Olympics,  wars,  or  coronations — had  become  commonplace  in  the  world's 
living  rooms."83 

Arthur  Clarke's  prophecy  of  "global  TV"  and  "citizens  of  the  world"  had 
arrived.  By  the  1980s,  satellite  television  had  grown  so  popular,  especially 
in  rural  and  mountainous  areas  where  standard  TV  reception  was  poor  or 
cable  TV  business  did  not  reach,  that  hundreds  of  thousands  of  people  in 
the  United  States  and  around  the  world  were  installing  their  own  personal 
satellite  dishes  in  their  backyards,  thereby  receiving  into  their  homes  directly 
from  space  a  seemingly  boundless  number  of  channels  and  programs,  only 
a  small  fraction  of  which  they  would  have  had  access  to  through  their  local 
ultra-high  frequency  (UHF),  VHF,  or  even  cable  stations.  By  the  early 
1990s,  many  people  and  governments  around  the  world  were  relying  for 
their  news  not  on  local  or  even  national  stations,  but  on  Ted  Turner's  Cable 
News  Network  (CNN)  via  satellite  from  Atlanta. 

In  just  a  few  decades,  Arthur  Clarke's  idea  for  a  global  communications 
system  (for  which  the  British  radio  journal  paid  him  the  equivalent  of  a 
measly  $40)  exploded  into  a  multibillion-dollar  industry,  leading  Clarke  to 
pen  a  facetious  little  article,  "A  Short  Pre-History  of  Comsats,  Or:  How  I 
Lost  a  Billion  Dollars  in  My  Spare  Time."84  None  of  this,  not  even  Clarke's 
humorous  lament,  would  have  been  possible  with  just  the  passive  reflectors. 

Others  besides  Clarke  also  came  to  recognize  the  missed  opportunities.  In 
1962  and  1963,  when  members  of  the  Project  Echo  Task  Group  first  learned 
in  detail  about  the  capabilities  of  the  inaugural  active  comsats  Telstar  and 
Relay,  they  were  a  little  disappointed  that  they  had  spent  so  much  time  on 
the  passive  reflector.  "I  remember  thinking,  damn,  we  worked  on  the  wrong 

195 


Space/light  Revolution 

one!"  recalls  Norman  Crabill.  "Except  I  really  didn't  because  I  had  learned 
a  lot.  Whether  it  was  active  or  passive,  I  had  a  job  to  do."85  In  the  late 
1950s  and  early  1960s,  before  the  advent  of  the  silicon  chip,  which  completely 
altered  the  scale  of  electronic  devices  and  made  possible  the  miniaturized 
amplifiers  required  for  actively  transmitting  satellites,  the  passive  reflector 
seemed  to  be  the  only  "do-able"  technology.  Because  of  their  work  on 
passive  satellite  technology,  Crabill  and  many  other  Langley  researchers  had 
prepared  themselves  well  for  the  management  of  more  significant  unmanned 
spaceflight  and  satellite  programs,  such  as  Lunar  Orbiter  and  the  Viking 
landing  on  Mars. 


196 


Learning  Through  Failure: 

The  Early  Rush  of  the  Scout 

Rocket  Program 


Failure  analysis  is  basically  research,  when  you  get 
down  to  it.  You  recover  and  learn  from  mistakes; 
you  don't  do  that  with  success. 

— Eugene  Schult,  head  of  guidance 
and  control  work  for  the  Scout  Project 
at  NASA  Langley 


Nothing  demonstrates  the  pitfalls  of  rushing  into  space  more  dramatically 
than  the  early  history  of  the  Scout  rocket  program.  This  relatively  small, 
four-stage  solid-fuel  rocket  was  conceived  in  1956  by  NACA  engineers  in 
Langley's  PARD  as  a  simple  but  effective  way  of  boosting  light  payloads 
into  orbit.  Scout  eventually  proved  to  be  one  of  the  most  economical, 
dependable,  and  versatile  launch  vehicles  ever  flown — not  just  by  NASA 
but  by  anyone,  anywhere.  The  program  did  not  begin,  however,  with  an 
impressive  performance;  it  began  with  four  years  of  confidence-crushing 
failures.  To  make  Scout  a  success,  researchers  had  to  climb  a  long  and 
torturous  learning  curve,  which  resembled,  at  least  to  those  involved,  the 
infernal  hill  up  which  Sisyphus  eternally  pushed  his  uncooperative  rock. 


"Itchy"  for  Orbit 

Max  Faget,  Joseph  G.  Thibodaux,  Jr.,  Robert  O.  Piland,  and  William 
E.  Stoney,  Jr.,  formed  the  core  of  a  notoriously  freethinking  group  within 
Langley's  PARD.  Early  in  1956,  a  year -and  a  half  before  Sputnik  1,  this 
group  began  playing  with  the  idea  of  developing  a  multistage  hypersonic 

197 


Spaceflight  Revolution 

rocket.1  These  engineers  had  been  launching  dozens  of  rockets  each  year 
from  the  lonely  beach  at  Wallops  Island.  To  them,  the  idea  of  building  one 
powerful  enough  to  reach  orbit  did  not  seem  at  all  farfetched.2 

Moreover,  the  organizers  of  the  IGY  in  1955  had  asked  expressly  for 
someone  to  put  up  the  first  artificial  satellite  as  the  highlight  of  the  upcoming 
celebration.  In  response,  the  governments  of  the  United  States  and  the 
Soviet  Union,  respectively,  on  two  consecutive  days,  29  and  30  July  1955, 
had  announced  their  rival  intentions  to  launch  satellites.  Each  country,  given 
its  burgeoning  ballistic  missile  program,  expressed  confidence  that  it,  and 
not  the  other,  would  be  the  first  to  put  an  object  in  space.  A  few  months 
later,  in  the  fall  of  1955,  the  Eisenhower  administration  made  the  ultimately 
history-turning  (and  in  the  opinion  of  some  critics,  disastrous)  decision  to 
endorse  the  navy's  Vanguard  proposal — and  Viking  booster — as  the  way  to 
launch  America's  first  satellite.  Viking's  competitor,  the  army's  Jupiter  C 
rocket,  the  darling  of  von  Braun  and  associates  in  Alabama,  had  to  wait  in 
the  wings,  ready  to  perform  when  the  Vanguard  program  flopped.3 

But  von  Braun's  rocket  experts  were  not  the  only  ones  "itchy"  for  or- 
bit. The  PARD  group  felt  that  the  boosting  of  a  small,  lightweight  pay  load 
into  orbit  would  require  only  an  extension  of  the  hypersonic  solid-fuel  rocket 
technologies  that  they  had  been  developing  at  Wallops  Island  and  Langley 
since  the  early  1950s.  "Solid-fueled  rockets  have  always  had  the  advantage 
over  liquid- fueled  as  far  as  simplicity,  cost,  and  possibly  reliability,"  remem- 
bers PARD  engineer  and  later  Scout  Project  team  member  Roland  D.  "Bud" 
English.  The  PARD  group's  idea  was  to  employ  solid  propulsion  and  use 
as  many  existing  solid-fuel  rockets  for  the  various  stages  of  the  proposed 
launch  vehicle  as  possible.  "It  was  the  logical  extension  of  the  work  going 
on  in  PARD  on  solid  rockets,"  says  English.  "It  was  a  natural  progression 
from  Mach  15  [ballistic  velocity]  to  the  audacity  to  think  in  terms  of  orbit," 
agrees  his  colleague  James  R.  Hall.4 

In  the  mid-1950s,  large  solid- fuel  rocket  motors  such  as  the  Cherokee  and 
the  Jupiter  Senior — the  latter  being  the  largest  solid-fuel  rocket  motor  up  to 
that  time — were  undergoing  rapid  development  to  meet  the  need  to  power 
the  U.S.  military's  growing  fleet  of  ballistic  missiles.  The  PARD  engineers 
were  convinced  that  by  combining  a  few  of  these  new  motors  intelligently 
into  a  three-  or  four-stage  booster  configuration,  the  NAG  A  in  a  relatively 
short  period  could  develop  a  launch  vehicle  that  would  have  enough  power  to 
shoot  past  ballistic  velocity  and  fly  into  orbit.  This  would  require  a  speed 
of  at  least  Mach  18.  The  Honest  John  rocket,  a  five-stage  vehicle  under 
development  for  the  army,  had  achieved  speeds  of  Mach  15  in  flight  tests  at 
Wallops  in  the  summer  of  1956,  and  the  Sergeant,  a  five-stage  rocket  also 
under  development,  was  supposed  to  be  capable  of  Mach  18.  Other  rocket- 
stage  motors  were  under  way  for  the  navy's  Polaris  and  Vanguard  project 
missiles.  From  this  promising  menu,  PARD  engineers  believed  they  could 
assemble  a  stack  of  rocketry  that  could  achieve  orbit.5 


198 


The  Early  Rush  of  the  Scout  Rocket  Program 

The  only  problems  were  that  this  stack  would  amount  to  "the  most 
expensive  vehicle  ever  developed  by  PARD"  and  "no  funds  were  immediately 
available."6  Furthermore,  as  the  rocket  was  to  serve  as  a  satellite  launch 
vehicle,  it  directly  competed  not  only  with  the  navy's  presidentially  anointed 
Vanguard  and  the  army's  overlooked  Jupiter  C,  but  also  with  the  air 
force's  Thor-Able,  which  was  rapidly  nearing  completion.  These  long-range 
military  rockets,  all  of  them  liquid  rather  than  solid-fuel,  made  the  case 
for  the  little  Scout  harder  to  advocate.  The  modest  PARD  proposal  for  a 
simpler,  cheaper,  and  potentially  more  reliable  bantam  rocket  simply  could 
not  compete  with  such  heavyweights. 

Then  came  Sputnik,  complicating  this  contest  among  American  rocket 
initiatives.  Engineer  William  Stoney,  perhaps  the  earliest  champion  of  what 
became  the  Scout  Project,  remembers  feelings  within  PARD  about  Sputnik, 
"We  were  disappointed  we  weren't  the  first  but  in  another  sense  it  reassured 
us  that  we  were  really  on  the  right  track — that,  boy,  we  really  could  get 
supported  from  now  on,  because  this  was  important  that  the  U.S.  continue 
to  try  to  catch  up,  and  we  were  part  of  that  game."  Sputnik  made  the 
PARD  rocketeers  think  "at  a  whole  new  level  of  exploration  that  heretofore 
was  beyond  consideration."7 

In  the  hectic  and  uncertain  months  following  Sputnik,  PARD  tried 
to  push  a  formal  proposal  for  its  rocket  development  through  Langley 
management  for  consideration  by  NACA  headquarters.  In  January  1958, 
however,  Ira  H.  Abbott,  one  of  the  NACA's  assistant  directors  for  research  in 
Washington  who  had  excellent  connections  to  Langley,  informed  PARD  that 
"NACA  Headquarters  would  not  be  receptive  to  a  proposal  for  development 
of  another  satellite  vehicle."8  However,  the  political  environment  for  such 
proposals  was  in  a  state  of  flux  in  early  1958,  and  Langley  engineers  knew 
it;  therefore,  they  kept  design  studies  for  their  rocket  going  even  after  such 
an  emphatic  refusal. 

In  late  March  1958,  another  Langley  veteran,  John  W.  "Gus"  Crowley, 
associate  director  for  research  at  NACA  headquarters,  revived  hopes  for  the 
rocket  when  he  asked  Langley  to  prepare  a  "Space  Technology  Program"  for 
the  prospective  new  space  agency.  In  its  report,  submitted  on  15  May,  the 
Langley  senior  staff,  "without  any  opposition,"  included  the  PARD  concept 
"as  a  requirement  of  the  program  for  the  investigation  of  manned  space 
flight  and  reentry  problems."  The  report  stated  that,  for  $4  million,  Langley 
could  develop  a  booster  that  launched  "small-scale  recoverable  orbiters"  into 
space,  and  could  do  it  in  a  matter  of  months.9 

Even  before  the  report  circulated,  on  6  May,  Langley  requested  a  research 
authorization  to  cover  "the  investigation  of  a  four-stage  solid-fuel  satellite 
system  capable  of  launching  a  150-pound  satellite  in  a  500-mile  orbit." 
Formal  approval,  which  took  just  a  few  weeks,  meant  that  PARD's  vehicle 
had  officially  made  it  into  the  space  program.10  The  air  force's  interest 
in  an  advanced  solid-fuel  rocket  test  vehicle,  with  mutually  acceptable 
specifications  for  a  joint  system  negotiated  in  July,  further  secured  Scout's 

199 


Space/light  Revolution 

position.  Such  a  deal  eventually  complicated  the  Scout  Project  greatly, 
however,  because  Langley  had  to  take  on  the  added  burden  of  handling  many 
of  the  contractual  details  for  the  coordinated  NASA/DOD  project.  The 
DOD  objective  was  to  obtain  a  fleet  of  solid-fuel  boosters  for  support  of  the 
air  force's  wide  range  of  space  research  projects,  which  at  that  time  included 
Dyna-Soar  support,  anti-ICBM  research,  and  nuclear  weapons.  The  last  of 
these  was  to  lead  to  the  development  of  the  so-called  "Blue  Scout"  rocket.11 

After  Scout  won  further  approval,  engineering  analysis  of  the  rocket 
system  indicated  that  the  proposed  third-stage  motor  (an  ABL  X248)  had 
to  be  replaced  with  a  larger  motor  of  the  same  type.  This  was  a  problem 
that  could  have  killed  an  earlier  proposal  but  now  bothered  no  one.  As 
one  PARD  veteran  remembers,  "The  overall  space  plans  for  NASA  were  so 
grandiose  when  compared  with  NACA  operations"  that  such  changes,  and 
such  costs,  were  now  relatively  minor  items.12 


Little  Big  Man 

Sometime  during  1958,  PARD's  William  Stoney,  soon  to  be  assigned 
overall  responsibility  for  development  of  the  new  rocket,  named  it  "Scout." 
Given  engineers'  propensity  for  acronyms,  some  believed  Scout  stood  for 
"Solid  Controlled  Orbital  Utility  Test  System";  however,  Stoney  insists  to- 
day that  the  various  acronyms  that  have  appeared  attached  to  the  name 
"Scout"  (even  in  official  publications)  have  all  been  "after-the-fact  addi- 
tions." According  to  him,  Scout  was  named  in  the  spirit  of  the  contemporary 
Explorer  series  of  satellites  with  which  the  rocket  would  often  be  paired.  He 
and  his  colleagues  gave  no  thought  at  the  time  to  deriving  its  name  from  a 
functional  acronym.13 

As  for  the  technical  definition  of  the  rocket,  as  suggested  earlier,  the 
Langley  engineers  tried  to  keep  developmental  costs  and  time  to  a  minimum 
by  selecting  components  from  off-the-shelf  hardware.  The  majority  of 
Scout's  components  were  to  come  from  an  inventory  of  solid- fuel  rockets 
produced  for  the  military,  although  everyone  involved  understood  that  some 
improved  motors  would  also  have  to  be  developed  under  contract.*  By  early 
1959,  after  intensive  technical  analysis  and  reviews,  Langley  settled  on  a 
design  and  finalized  the  selection  of  the  major  contractors.  The  rocket's 
40-inch-diameter  first  stage  was  to  be  a  new  "Algol"  motor,  a  combination 
of  the  Jupiter  Senior  and  the  navy  Polaris  produced  by  the  Aerojet  General 
Corporation,  Sacramento,  California.  The  31-inch-diameter  second  stage, 


The  only  new  technology  required  for  Scout  was  its  hydrogen  peroxide  reaction-jet  control  system, 
developed  by  contractor  Walter  Kidde  and  Company,  which  enabled  controlled  flight  outside  the 
atmosphere.  Later  versions  of  this  technology  would  be  used  for  various  purposes  in  space  programs, 
including  the  spatial  orientation  and  stabilization  of  "Early  Bird,"  ComSatCorp's  first  experimental 
satellite. 


200 


The  Early  Rush  of  the  Scout  Rocket  Program 


In  this  photo  from  October 
1960,  Scout  Test-2  (ST-2) 
stands  ready  for  launch  at 
Wallops  Island. 


L-60-6627 


"Castor,"  was  derived  from  the  army's  Sergeant  and  was  to  be  manufactured 
by  the  Redstone  Division  of  the  Thiokol  Company  in  Huntsville,  Alabama. 
The  motor  for  the  30-inch-diameter  third  stage,  "Antares,"  evolved  under 
NASA  contract  from  the  ABL  X248  design  into  a  new  version  called  the 
X254  (and  subsequently  into  the  X259);  it  was  built  under  contract  to 
NASA  by  ABL,  a  U.S.  Navy  Bureau  of  Ordnance  facility  operated  by  the 
Hercules  Powder  Company,  Cumberland,  Maryland.  The  final  upper-stage 
propulsion  unit,  "Altair,"  which  was  25.7  inches  in  diameter  (34  inches  at  the 
heat  shield),  amounted  to  an  improved  edition  of  the  X248  that  was  also 
manufactured  by  ABL.  Joining  these  four  stages  were  transition  sections 
containing  ignition,  guidance  and  attitude  controls,  spin-up  motors,  and 
separation  systems. 

Upon  assembly  of  the  vehicle,  which  was  to  be  done  by  Chance  Vought 
of  Dallas,  the  rocket's  airframe  and  control-system  contractor,  the  original 
Scout  stood  only  72  feet  high  from  the  base  of  its  fins  to  the  tip  of  its  nose 
cone  and  weighed,  at  first-stage  ignition,  a  mere  37,000  pounds.  The  thrust 
of  the  four  stages  added  together  totaled  just  over  200,000  pounds,  which  was 
easily  enough  to  carry  the  proposed  150-pound  payload  into  space,  although 
at  a  300-mile  rather  than  a  500- mile  orbit.  (Later  versions  of  Scout  would 


201 


Spaceflight  Revolution 

eventually  fly  missions  with  300-pound  and  even  450-pound  pay  loads,  using 
an  optional  fifth  stage.)14 

For  such  a  comparatively  small  rocket,  Scout  turned  into  something  quite 
significant — the  first  large  NASA  project  that  Langley  ran  in-house.  The 
only  previous  in-house  project  to  match  Scout  was  the  Bell  X-l  supersonic 
research  airplane  of  a  decade  earlier.  The  X-l,  however,  was  a  joint  effort 
with  the  air  force  and  was  physically  remote  from  Langley  at  faraway  Muroc 
Dry  Lake  in  California.  The  plane  never  got  to  Langley  Field,  although 
NACA  Langley  was  primarily  responsible  for  its  development.15  The  Scout 
project,  on  the  other  hand,  was  conceived,  designed,  and  for  the  most  part, 
built  at  Langley.  Components  were  brought  to  nearby  Wallops  for  launch 
and  flight  testing,  thus  making  a  "very  tight  Langley  loop."16 

A  formal  "Scout  Project  Group"  was  not  organized  at  Langley  until 
February  1960  after  a  recommendation  was  made  by  a  NASA  headquar- 
ters review  committee  chaired  by  NASA  Lewis  Research  Center's  Bruce 
Lundin.17  Until  that  time,  all  the  work  on  the  rocket  had  been  overseen 
first  by  regular  PARD  management  and  later,  after  the  creation  of  the  STG 
in  1958,  by  Bob  Gilruth  and  his  staff.  Gilruth  already  had  his  hands  full 
with  Project  Mercury,  but  he  reluctantly  took  over  the  responsibility  for 
a  short  period  because  Abe  Silverstein,  whose  Office  of  Space  Flight  De- 
velopment initially  funded  Scout,  insisted  on  it.*  Given  Gilruth's  intimate 
knowledge  of  PARD  and  its  personnel,  he  trusted  the  Scout  engineers  to 
manage  themselves.  So  did  Langley  Associate  Director  Floyd  Thompson, 
who  gave  Scout  personnel  "remarkable  freedom"  to  operate  almost  indepen- 
dently. "[Our  work]  was  of  course  part  of  the  race  to  catch  the  Russians," 
James  Hall  has  stated.  "But  more  important  it  was  to  prove  something  to 
ourselves.  People  worked  hard  and  were  selfless  about  helping  each  other." 
They  were  mostly  young  men  "trying  to  do  something  that  had  never  been 
done  before."1  Such  naive  enthusiasts  neither  cared  for  nor  would  have 
benefited  from  top-heavy  management. 

The  project  office  started  small  with  nine  personnel:  Project  Office 
Head  Bill  Stoney,  Technical  Assistant  Bud  English,  Administrative  Assis- 
tant Abraham  Leiss,  three  project  engineers  (C.  T.  Brown,  Jr.,  Eugene  D. 
Schult,  and  William  M.  Moore),  Field  Director  James  Hall,  Project  Co- 
ordinator Elmer  J.  Wolfe,  Secretary  Edith  R.  Horrocks,  plus  two  resident 
representatives  from  industry.  Each  division  of  the  laboratory  also  made 
one  employee  responsible  for  coordinating  support  for  Scout  whenever  it 
was  required.19 

This  skeletal  crew  and  associated  shadow  organization  began  to  race 
against  the  calendar  to  build  and  launch  Scout.  The  project  team  grew 
in  size  rather  quickly,  so  by  1962  more  than  200  Langley  staff  members 


!H 

With  the  establishment  of  the  Scout  Project  Group  in  February  1960,  the  Scout  team  began  to 
report  instead  to  Donald  R.  Ostrander,  director  of  the  new  Office  of  Launch  Vehicle  Programs  at  NASA 
headquarters,  which  was  established  on  29  December  1959  and  had  responsibility  for  all  launch  vehicles. 

202 


The  Early  Rush  of  the  Scout  Rocket  Program 


James  Hall,  Langley's  original  field  director 
for  the  Scout  Project  Group. 
L-61-5176 


were  working  almost  exclusively  on  Scout,  which  even  project  leaders  had 
to  concede  was  "a  very  large  segment  of  people  to  work  on  anything 
at  Langley."20  At  Wallops,  Scout  work  dominated,  taking  over  several 
assembly  shops  and  other  buildings.  The  core  staff  in  the  project  office 
stayed  relatively  small,  however,  reaching  its  peak  of  55  employees  in  1965 
and  then  dropping  back  to  34  by  the  time  of  Scout  flight  number  75  in  1971. 
The  involvement  of  contractors  was  essential.  Especially  helpful  were 
the  people  from  Chance  Vought — soon  to  be  organized  into  the  LTV  Missile 
Group  of  the  Chance  Vought  Corporation — who  had  won  the  bid  to  develop 
the  Scout  airframe  and  launching  capability.21  The  partnership  between 
Langley  and  LTV  grew  into  one  of  the  most  cooperative,  fruitful,  and  long- 
lasting  (30  years)  in  NASA  history.  As  James  Hall  remembers  so  fondly, 
from  almost  the  beginning,  the  Scout  Project  Office  made  "no  distinction 
between  government  people  and  contractors.  We  were  all  on  the  same  team 
and  did  what  we  had  to  do  regardless  of  the  color  of  our  badges."  The 
feeling  was  mutual.  According  to  Ken  Jacobs,  who  worked  many  years  on 
Scout  for  LTV,  "We  were  very  much  more  liable  to  work  together  than  we 
were  to  work  apart.  If  your  counterpart  in  the  government  had  a  problem  or 
a  question,  he  would  contact  you  on  the  telephone  and  [we  would]  be  able  to 
come  up  with  a  mutual  agreement  or  solution.  The  end  result  was  that  the 
program  would  be  much  better  off  for  experiencing  this  degree  of  cooperation 
between  the  two  individuals  who  had  the  task."  Milt  Green,  another  LTV 
employee,  remarks,  "We  all  had  one  common  goal."  The  teamwork  resulted 
from  "a  mutual  respect  for  each  other.  It  wasn't  an  adversarial  relation  with 
a  lot  of  gnarling  of  hands.  [It  was]  strictly  a  job  that  had  to  be  done,  and 
done  in  the  most  reliable  manner."22 


203 


Space/light  Revolution 


L-67-95 

In  this  picture  from  June  1967,  32  LTV  employees  pose  in  front  of  the  Scout  S-159C, 
which  a  few  months  later  on  19  October  1967  would  successfully  carry  the  RAM  C-l 
experiment  into  orbit. 


This  deeply  felt  sense  of  mutual  reliance  and  cooperation,  at  least  on 
Langley's  part,  related  to  the  deeply  ingrained,  40-year-old  NACA  culture. 
If  Langley's  work  included  Scout,  and  Scout  needed  LTV  to  succeed,  then 
Langley  needed  LTV  and  would  consider  it  a  member  of  the  family;  that 
was  the  formula.  The  LTV  staff  appreciated  it.  "It  was  just  a  close-knit, 
dedicated  group,"  remembers  Larry  Tant,  a  Scout  operations  manager  for 
Langley.  "We  had  a  lot  of  pride  in  what  we  were  doing.  We  were  like 
brothers."  "There  was  something  about  the  program,"  declares  Jon  Van 
Cleve,  an  early  Scout  team  member  from  Langley.  "You  worked  in  it  for  a 
little  while  and  really  got  involved  in  it.  When  that  happened,  you  lost  the 
lines  of  whether  you  were  agency,  LTV,  or  air  force.  You  became  a  Scout 
person."23 

These  warm  testimonials,  which  came  years  after  Scout  had  amassed 
its  remarkable  record,  were  mostly  made  in  the  early  1990s  when  NASA 
honored  members  of  the  Scout  team  in  official  ceremonies.  Both  government 
and  industry  Scout  staff  reminisced  about  Scout's  record,  which  no  other 
booster,  large  or  small,  foreign  or  American,  had  surpassed.  Langley's  Scout 
Project  Group  enjoyed  incredible  strings  of  22  and  37  consecutive  launches 
without  failure  during  two  long  periods,  the  first  lasting  from  July  1964 
until  January  1967  and  the  second  from  September  1967  until  December 
1975.  In  1991,  when  NASA  Langley  reluctantly  turned  over  the  direction 
of  the  Scout  Project  to  NASA  Goddard  and  the  commercial  production  of 
the  vehicle  to  LTV,  the  scorecard  of  113  launches  showed  an  overall  success 
rate  of  an  astounding  96  percent. 

204 


The  Early  Rush  of  the  Scout  Rocket  Program 

The  retrospective  comments  about  the  great  teamwork  on  the  Scout 
Project  need  to  be  understood  within  the  context  of  that  final  glorious 
record.  The  feelings  between  government  and  contractor  could  not  have 
been  so  positive  in  the  early  years  of  the  Scout  program,  when  one  rocket 
after  another  self-destructed  or  otherwise  failed.  In  the  Scout  program, 
such  can-do,  throw-your-arm-around-your-buddy  camaraderie  developed 
only  gradually  and  was  tested  severely  by  frequent  early  experiences  with 
misfortune.  These  are  trials  that  Scout  team  members  understandably 
prefer  to  forget. 


Little  Foul-Ups 

On  18  April  1960,  only  14  months  after  the  creation  of  the  Langley  Scout 
Project  Office,  the  first  experimental  Scout  sat  ready  to  be  fired  from  a  new 
launch  tower  at  Wallops  Island.  For  weeks  NASA  headquarters  had  been 
demanding  that  "some  type  of  flight  test  be  made  in  the  Scout  program 
as  soon  as  possible."24  The  only  way  for  Langley  and  its  contractors  to 
meet  this  demand  was  to  work  hours  of  overtime.  James  Hall  recollects  the 
days  before  this  and  other  early  Scout  launches:  "It  was  schedule-driven. 
People  worked  very  hard  and  long  hours.  This  was  such  a  dynamic  program, 
people  felt  compelled  to  work  however  long  it  took.  The  closer  to  launch, 
the  more  demanding  the  schedule  became.  At  the  launch  site,  people  would 
often  go  without  sleep  to  make  up  time  or  make  something  work  or  correct 
a  problem.  People  were  consumed  by  the  program's  schedule."25 

On  7  March,  after  deflecting  the  demands  of  NASA  headquarters  for  as 
long  as  he  could,  Langley  Director  Floyd  Thompson  gave  in  and  conceded 
that  an  unguided  Scout,  one  greatly  reduced  in  scope  from  later  Scouts, 
could  be  fired  under  the  direction  of  Langley's  Applied  Materials  and 
Physics  Division  (the  old  PARD)  to  obtain  "some  information  on  the  overall 
configuration,"  but  preparations  would  take  a  month.  The  second  stage  of 
the  rocket  would  have  to  be  a  weighted  dummy,  one  of  several  supplied  by 
the  contractors  for  fitting  transition  sections  during  construction  and  for 
checking  "overall  alignment  and  general  suitability  including  freedom  from 
interference  with  components  supplied  by  other  contractors."  Thompson, 
reflecting  the  concerns  of  his  Scout  Project  team,  wanted  it  to  be  known 
that  this  was  "not  an  official  Scout  test."  It  was  an  "expedited  launch,"  a 
"Cub  Scout,"  meant  only  to  obtain  engineering  data  on  the  vehicle.26 

Several  problems  occurred  during  this  hurried,  "unofficial"  test  flight  of 
Cub  Scout.  The  rocket  rolled  more  than  anticipated  during  ascent,  thus 
causing  a  structural  failure  near  the  burnout  of  the  first  stage.  This  failure 
prevented  the  third  stage  (atop  the  second-stage  dummy  motor)  from  test- 
firing.  In  addition,  the  heat-shield  design  proved  defective  by  breaking  away 
from  the  fourth  stage  as  the  vehicle  passed  through  the  transonic  region. 
This  was  not  the  start  everyone  had  hoped  for.  Scout  Project  personnel 

205 


Space/light  Revolution 


L-63-5567 

Spectators — mainly  the  families  of  NASA  employees — usually  filled  the  makeshift 
grandstand  at  Wallops  Island  to  witness  the  launch  of  NASA  's  small  unmanned 
rockets. 

tried  to  put  on  a  happy  face,  remarking  that  the  test  provided  valuable 
experience  assembling  the  rocket's  components  on  the  new  launcher  and 
actually  firing  from  it.  No  one,  however,  was  fooled.  The  launch  had  been 
important  to  the  test  program  and  was  meant  to  develop  confidence  in  the 
systems.  As  one  Langley  engineer  at  the  launch  later  recorded,  "The  failure 
was  a  blow  to  the  prestige  of  the  project,  and  efforts  to  complete  the  first 
actual  Scout  were  redoubled."27 

With  these  efforts,  the  first  launch  of  a  full-fledged  Scout —  ST-1 — was 
to  take  place  on  1  July  1960,  less  than  three  months  after  Cub  Scout's  little 
foul-up.  The  anticipation  level  was  extremely  high.  By  the  time  of  Scout 
Test-1,  Langley  had  been  firing  rockets  at  Wallops  for  16  years,  since  1944. 
To  Langley  engineers  launching  rockets  might  have  been  "old  hat,"  but 
this  launch  was  different.  "Everybody  was  excited,"  James  Hall  remembers 
with  a  gleam  in  his  eye.  "The  concept  of  launching  an  orbital  vehicle  was 
a  new  and  a  really  exciting  challenge.  That  launch  was  the  culmination 
of  two  years  of  intensive  work.  We  had  a  number  of  practice  countdowns 
and  dry  runs.  We  got  our  timing  down  and  got  things  all  set  up.  In  fact, 
we  were  almost  wearing  that  thing  out  testing  it.  That's  what  you're  up 


206 


The  Early  Rush  of  the  Scout  Rocket  Program 


The  first  Scout  was  launched  on  the 
evening  of  1  July  1960. 
L-60-2565 


against  in  space.  But  you  reach  a  point  where  you  have  to  come  down  to 
the  countdown."28 

The  countdown  lasted  11  hours.  As  it  progressed,  the  Scout  launch 
team  gradually  moved  away  from  the  vehicle.  During  the  last  hour  or  so, 
the  rocketeers  moved  into  a  little  cinder  block  building  with  three  racks  of 
switches,  controls,  and  displays.  By  later  launch  system  standards,  it  was 
simple,  crude  technology,  but  it  was  enough  to  light  the  fuse.  Compared 
with  ICBMs,  Scout  was  tiny,  but  compared  with  all  the  previous  rockets 
launched  at  Wallops,  it  was  quite  large.  When  the  first-stage  Algol  motor 
was  lit  up,  with  its  approximately  100,000  pounds  of  thrust,  an  awesome 
energy  was  released.  As  Hall  describes  it,  "The  ground  shakes  and  the  fire 
and  smoke  appear.  It's  a  very  splendid  thing."29  It  was  similar  to  being 
close  to  the  heart  of  an  earthquake,  with  massive  pressure  waves  bouncing 
off  the  chest. 

The  rocket  was  to  ascend  into  space  and  then  use  its  last-stage  Altair 
motor  to  fire  a  193-pound  acceleration  and  radiation  package  back  through 
the  atmosphere  as  a  probe.  But  Scout  did  not  reach  a  high  enough  altitude 
to  fire  the  package.  One  of  the  new  built-in  features  of  the  full-fledged  Scout 
system  was  a  destruct  capability  to  be  used  if  the  rocket  flew  off  course  and 
endangered  populated  areas.  Scout  Test-1  would  appear  to  do  just  that. 

At  136  seconds  after  launch,  radar  tracking  on  the  ground  showed  that 
the  rocket  had  gone  off  course.  A  rolling  moment  (i.e.,  an  aerodynamic 

207 


Space/light  Revolution 


L-62-6440 
The  cone-shaped  cinder  block  building  was  the  site  of  Scout 's  launch  control. 


L-67-212 

Worried  looks  on  the  faces  of  the  NASA  men  in  the  control  room  at  Wallops  bear 
witness  to  the  vicissitudes  of  launching  rockets  into  space. 

208 


The  Early  Rush  of  the  Scout  Rocket  Program 

tendency  to  rotate  the  body  about  its  longitudinal  axis)  developed  with  the 
Antares  motor,  and  then  it  just  as  quickly  dissipated.  However,  the  rolling 
caused  a  very  slight  disorient  at  ion  of  the  radar  tracking.  As  the  postflight 
test  analysis  would  later  show,  the  shift  of  the  radar  that  was  indicated  on 
the  plotboard  meant  that  "the  vehicle  had  taken  a  violent  turn  in  azimuth 
and  a  dip  down  in  elevation."  The  rocket  seemed  to  be  about  50  degrees  off 
course  and  heading  somewhere  it  definitely  was  not  supposed  to  go.  This 
deviation  forced  the  radar  safety  officer  inside  the  blockhouse  to  take  action. 
He  actuated  the  "hold-fire"  signal  for  the  fourth  stage,  then,  as  James  Hall, 
who  was  also  in  the  blockhouse  bitterly  recalls,  the  range  officer  "waited 
as  long  as  he  could,  looked  over  to  us,  and  we  had  to  concur.  He  hit  the 
destruct  button."30  The  radar  tracking  recovered  quickly,  thus  showing  that 
there  really  never  had  been  a  significant  problem. 

"That  was  a  crushing  blow  to  destroy  a  rocket  that  was  doing  exactly 
what  it  was  programmed  to  do,"  Hall  laments  30  years  later,  "but  which  just 
indicated  on  a  range  safety  plotboard  that  it  was  on  an  incorrect  trajectory. 
You  can't  imagine  how  hard  people  worked  as  a  group  to  bring  this  to  the 
launch  point."  Inside  the  blockhouse  the  men  kicked  cans  and  cussed  the 
unfortunate  safety  officer.  "But  after  an  hour,"  according  to  Hall,  "most 
people  recognized  there  was  only  one  thing  to  do.  That  was  to  work 
and  build  the  next  vehicle,  which  we  did  in  three  or  four  months."  The 
spaceflight  revolution  demanded  nothing  less.31 

In  fact,  circumstances  also  demanded  that  the  test  be  labeled  a  success 
even  when  everyone  knew  better.  The  many  subsequent  chronicles  of  the 
Scout  program  all  classified  ST-1  as  a  success.  "The  fourth  stage  never  had 
a  chance  to  perform,"  but  "radiation  measurements  were  successfully  made 
to  an  altitude  of  875  miles."32 


"3-2-1  Splash" 

On  4  October  1960,  ST-2  proved  to  be  the  first  real  success  of  the  project. 
Also  launched  as  a  probe  with  a  radiation  pay  load  on  board,  the  rocket 
reached  a  maximum  altitude  of  3500  miles  and  achieved  a  total  range  of 
5800  miles.  The  newspaper  headlines  underscored  the  elation  surrounding 
this  successful  launch.  In  the  Newport  News  Times  Herald,  a  large  typeface 
banner  headline  celebrated  the  feat.  The  headline  on  page  12  of  section 
A  of  the  Washington  Post  read:  "  'Poor  Man's  Rocket'  Fired  Successfully." 
The  Washington  Star  followed  with  a  feature  article,  "Versatile  Scout  to 
Get  Space  Chores."  During  this  period,  Scout  also  received  additional 
positive  publicity  for  the  air  force's  successful  launch  of  two  of  its  "Blue 
Scouts."33  The  Scout  Project  was,  indeed,  looking  up. 

With  this  one  successful  but  nonorbital  mission  behind  them,  the  Scout 
leaders  believed  that  testing  was  complete  and  that  the  missile  was  ready  to 
start  operations  as  one  of  NASA's  launch  vehicles.  As  such,  it  would  be  used 

209 


Spaceflight  Revolution 

for  three  types  of  missions:  placing  small  satellites  in  orbit,  making  high- 
velocity  reentry  studies  and  testing  heat-resistant  materials,  and  launching 
high-altitude  and  space  probes.  The  Scout  rockets  were  scheduled  to  take  off 
not  only  from  Wallops  but  also,  beginning  in  early  1962,  from  a  Scout  launch 
site  being  prepared  at  the  Western  Test  Range  located  on  Vandenberg  AFB 
in  California.  NASA  headquarters  was  so  optimistic  about  Scout  that  it 
arranged  for  a  full-scale  72-foot  model  of  the  rocket  to  be  displayed  at  the 
15th  annual  meeting  and  "astronautical  exposition"  of  the  American  Rocket 
Society  in  mid-December  1960.  More  than  5500  attendees  viewed  the  model 
outside  the  Shoreham  Hotel  in  Washington,  B.C.,  and  were  impressed  that 
it  stood  almost  as  high  as  the  nine-story  building. 

Such  a  celebration  of  Scout  proved  premature.  The  first  orbital  flight 
from  Wallops  on  4  December  1960  failed.  Now  the  headlines  read,  "NASA 
Fizzles  Orbit  Attempt"  (Virginian-Pilot,  5  Dec.  1960),  "Scout  Sinks  After 
Fizzle"  (Norfolk-Portsmouth  Ledger-Star,  5  Dec.  1960),  "Feeling  of  Unsuc- 
cess  Persists  at  Rocket  Site"  (Norfolk  Ledger- Dispatch,  5  Dec.  1960),  and 
"Rocket,  Satellite  Lie  Under  Deep  Waters"  (Richmond  News  Leader,  5  Dec. 
1960).  Three  of  the  first  six  flights  were  in  fact  failures.  Not  long  there- 
after, Lt.  Col.  George  Rupp,  formerly  project  officer  on  the  Bullpup  missile 
weapons  system,  came  to  NASA  from  the  U.S.  Marine  Corps  to  replace 
a  disheartened  Bill  Stoney  as  director  of  Langley's  Scout  Project  Office.* 
This  change  solved  nothing  because  Stoney's  management  had  not  been  the 
problem.  In  the  first  four  months  following  Stoney's  replacement,  three  out 
of  four  launches  failed.  A  NASA  investigation  found  faults  with  the  elec- 
trical systems,  the  heat  shield,  the  ignition  systems,  and  much  more.  As  a 
later  Scout  program  brochure  recalls,  "This  was  a  time  of  exhilarating  suc- 
cesses and  heart  breaking  failures.  The  space  age  was  in  its  infancy  and  the 
participants  were  learning  about  the  operation  of  complex  systems  in  the 
unforgiving  environment  of  a  high  speed  flight  through  the  atmosphere  to 
the  border  of  space."34  Personal  accounts  come  closer  to  the  truth.  Roland 
"Bud"  English,  one  of  the  original  nine  members  of  the  Scout  Project  Group 
and  the  fourth  head  of  the  Scout  Project  Office,  remembers:  "The  Scout 
program  was  done  in  a  rush.  Unquestionably,  everything  was  behind  sched- 
ule, and  there  was  pressure  on  NASA  to  perform.  The  Space  Act  had  been 
passed,  and  NASA  was  supposed  to  be  going  up  to  do  a  job,  and  Scout  was 
part  of  that.  So  there  was  very  definitely  a  pressure  to  do  it  in  a  hurry, 
too  much  of  a  hurry,  and  not  enough  emphasis  on  proper  quality  and  really 
getting  ready  for  an  operational  flight."35 

Even  with  the  many  failures,  the  launch  dates  just  kept  coming.  "None 
of  us  liked  to  slip  a  commitment,"  James  Hall  admits,  "and  slippages  were 
relatively  modest  considering  the  complexity  of  the  program.  As  things 
got  down  to  deadline,  completion  of  ground  system  checkout,  completion  of 


M 

Rupp  stayed  in  this  post  until  his  retirement  from  the  military  in  June  1963,  whereupon  he  was 
succeeded  by  Eugene  D.  Schult  and  later  by  Roland  English. 

210 


The  Early  Rush  of  the  Scout  Rocket  Program 


Succeeding  Rupp  in  June  1963  was 
Langley  engineer  Eugene  D.  Schult, 
who  had  been  with  the  Scout  Project 
Office  from  the  start. 


L-63-4479 


launch  tower  checkout,  and  then  the  actual  practice  countdown  and  final 
launch  countdown — those  critical  milestones — didn't  slip  that  much,  but 
people  had  to  work  24  hours  a  day  to  hold  them."36 

Not  until  20  July  1963  and  the  launch  of  Scout  flight  number  22  did  the 
problem  come  to  a  head.  (Preceding  this  flight,  ironically,  three  consecutive 
missions  had  been  successful,  and  two  of  three  had  been  orbital.)  Two 
and  one-half  seconds  after  liftoff  at  Wallops,  a  flame  appeared  above  the 
first-stage  fins.  Two  seconds  later,  the  Algol  stage  became  engulfed  by  fire. 
"It  was  obvious  something  terrible  had  happened,"  Bud  English  recalls, 
frowning.  "You  could  tell  from  the  communications  coming  from  the  range 
safety  net  [work].  There  had  been  a  burnthrough  of  the  first  stage  nozzle  a 
few  seconds  after  takeoff.  The  vehicle  went  through  some  wild  gyrations.  It 
got  about  300  feet  high  and  broke  into  three  parts:  the  first  stage  went  in  one 
direction;  the  second  stage  went  in  another;  and  the  third  and  fourth  stages 
fell  more  or  less  back  on  the  launch  pad  and  burned.  It  was  a  disaster."37 

Langley's  Scout  engineer  Tom  Perry  was  part  of  the  recovery  team  that 
slogged  through  the  salt  marshes  a  mile  off  the  coastal  island  to  pick  up 
bits  and  pieces  of  the  rocket  to  help  NASA  determine  what  went  wrong.  He 
found  one  large  chunk  of  the  fiery  debris  in  an  unexpected  place.  "Someone 
had  parked  a  small  car  inside  one  of  the  assembly  buildings  and  it  just  so 
happened  that  a  flaming  piece  of  the  rocket  had  come  right  down  through 
the  roof  and  into  the  front  seat,  burning  that  car  to  a  crisp."38 

211 


Space/light  Revolution 


L-61-5743 

As  this  sequence  of  photos  demonstrates,  the  launch  of  ST-5  on  30  June  1961  went 
well;  however,  a  failure  of  the  rocket's  third  stage  doomed  the  payload,  a  scientific 
satellite  known  as  S-55  designed  for  micrometeorite  studies  in  orbit. 


NASA  headquarters  launched  a  formal  investigation.  A  seven-man 
review  board  found  flaws  in  a  rocket  nozzle  that  had  gone  undetected 
during  production  and  testing.  Following  the  board's  recommendation,  the 
space  agency  imposed  a  three-month  moratorium  on  the  launch  schedule; 
no  more  Scouts  would  fly  until  a  comprehensive  study  of  all  the  data  from 
the  previous  21  Scouts  had  been  completed.39 

Significantly,  the  in-depth  investigations  of  the  rocket's  subsystems  made 
during  this  review  revealed  that  each  Scout  failure  had  been  caused  by  a 
different  problem.  That  in  itself  was  the  essential  problem.  "We  never  had 
the  same  failure  twice,"  James  Hall  underscores,  "but  it  was  clear  from  the 
early  record  of  Scout  that  there  was  enough  miscellaneous  failure  that  we 
had  to  sit  down  and  rethink  the  whole  thing  very  seriously."40 

Certain  institutional  and  bureaucratic  factors  also  had  contributed  to 
Scout's  failures.  As  much  as  the  Langley  engineers  had  wanted  to  make 
Scout  contractors  and  air  force  partners  members  of  one  integrated  team, 
in  many  key  essentials  they  simply  were  not.  Former  PARD  engineer  and 


212 


The  Early  Rush  of  the  Scout  Rocket  Program 


L-62-1729  L-63-5790 

What  kept  the  Scout  engineers  going  through  the  tough  times  was  the  occasional 
spectacular  success.  In  this  photo  from  1  March  1962  (left),  ST-8  streaks  into  the 
night  sky  above  Wallops,  carrying  a  reentry  heating  experiment.  The  bumthrough 
of  ST-110's  first-stage  nozzle  just  seconds  after  firing  on  20  July  1963  resulted  in 
significant  damage  to  the  launch  tower  (right).  Remnants  of  the  third  and  fourth 
stages  of  the  erratic  Scout  can  be  seen  on  the  launchpad. 

member  of  Langley's  original  Scout  Project  Office  Eugene  Schult  remembers, 
"We  did  things  differently  at  Wallops  than  at  the  Western  Test  Range. 
The  air  force  had  its  own  way  of  doing  things;  the  contractor  had  his 
ways;  and  we  had  our  ways.  It  was  a  problem  trying  to  coordinate 
them.'"1  L  Essentially,  each  organization  employed  its  own  safety  procedures: 
an  assembly  checkout  line  at  the  LTV  plant  in  Dallas,  other  checkout  lines 
on  the  ground  at  Wallops  and  Vandenberg,  and  yet  two  more  lines  in  the 
towers  at  the  launchers,  both  in  California  and  Virginia.  Each  line  used 
different  equipment  and  procedures. 

However,  the  principal  cause  of  Scout's  mishaps  was  simply  the  need  to 
make  everything  happen  so  fast.  The  LTV  mission  integrator,  Ken  Jacobs, 
recalls  how  engineers  scrambled  to  assemble  the  rocket:  "Back  in  those  days, 
if  you  needed  a  part,  you  did  what  we  called  a  'midnight  requisition.'  We'd 
go  get  the  part  from  the  space  vehicle  in  inventory."  This  was  obviously  one 
of  the  shortcomings  of  the  system.  "People  were  robbing  Peter  to  pay  Paul, 
and  the  result  was  we  had  an  unsuccessful  vehicle."  Over  and  above  this 
"cannibalizing"  of  the  hardware,  Bud  English  feels  that  "there  simply  were 


213 


Spaceflight  Revolution 


L-60-7980 

In  this  photo  from  December  1960,  employees  of  Vought  Astronautics,  Scout's 
prime  contractor,  work  with  NASA  technicians  to  prepare  ST-3  for  launch.  Un- 
fortunately, this  rocket  would  fail  because  of  a  second-stage  misfire. 

not  good  standardized  vehicle  safeguards  and  checkout  procedures,  which 
were  needed  to  have  a  successful  vehicle."42 

"Our  record  was  not  good,"  Jacobs  has  to  admit.  "Our  reliability  was 
3-2-1  splash,  3-2-1  splash."  The  time  had  come  to  "blow  the  whistle  and 
take  a  look  at  this  program  and  see  what  our  problem  was."  The  Scout 
Project  Office,  LTV,  the  supporting  engineers  at  Langley,  the  related  air 
force  personnel,  and  everyone  involved  had  to  sit  down  and  do  some  "deep 
thinking"  about  what  had  to  be  done  to  fix  not  only  the  rocket  but  also  the 
entire  program.43 


Recertification 

The  Scout  team  decided  that  a  14- month  reliability  improvement  pro- 
gram to  recertify  the  rocket  was  needed.  The  effort  was  spearheaded  by 
a  NASA/LTV/air  force  "tiger  team,"  whose  mission  was  "to  revise  com- 
pletely how  [the  project  office]  handled  the  vehicle  and  to  standardize  the 

214 


The  Early  Rush  of  the  Scout  Rocket  Program 

process  to  the  ultimate  degree."44  The  tiger  team  concept,  which  in  essence 
was  a  technological  commando  squad,  had  already  proved  effective  in  indus- 
trial settings.  NASA  was  beginning  to  use  it  more  frequently  in  the  1960s 
to  attack  particularly  troublesome  problems.  The  tiger  team's  activities 
started  at  Langley  when  James  Hall,  operations  manager  for  Scout,  wrote 
an  inch-thick  specification  that  laid  out  a  single  set  of  test  equipment,  a 
single  checkout  procedure,  and  the  rigorous  standards  for  using  both. 

Such  an  approach  proved  to  be  exactly  what  was  needed.  Under  the 
direction  of  the  tiger  team,  all  27  of  the  Scout  rockets  already  manufactured 
for  the  program  were  returned  to  LTV  in  Dallas  to  be  taken  apart  and 
inspected.  Weld  seams  were  X-rayed,  and  solder  joints  were  inspected  under 
microscopes.  Everything  that  could  be  standardized  was  standardized.  Even 
the  lengths  of  the  cable  in  Vought's  laboratories  now  had  to  match  those 
at  the  two  launch  sites.  The  launch  countdown  now  included  more  than 
800  items.  Additional  tiger  teams  were  put  together  at  Wallops  and  at 
Vandenberg  to  assure  compliance  with  the  new  standards.  No  Scout  was 
to  leave  Dallas  until  an  inspection  team  had  done  a  complete  worthiness 
review  of  the  whole  vehicle  and  given  it  a  clean  bill  of  health.45 

At  the  end  of  the  long  recertification  process,  nearly  all  members  of  the 
Scout  team  were  confident  that  they  now  understood  why  things  had  gone 
wrong:  from  the  time  that  NASA  had  adopted  the  concept  for  the  little  solid- 
fuel  rocket  and  made  it  an  agenda  item  for  the  spaceflight  revolution,  the 
Scout  Project  Group  simply  had  had  neither  the  time,  nor  the  inclination,  to 
look  before  they  leaped.  "We  all  underestimated  the  magnitude  of  the  job  at 
that  time,"  Milt  Green  admits.  "The  biggest  problem  we  had  was  denying 
the  existence  of  problems  that  we  did  not  understand."46  The  problem  was, 
of  course,  all  too  human. 

The  process  of  honestly  facing  up  to  fundamental  mistakes  and  moving 
beyond  them  was  probably  what  made  the  Scout  Project  Group  the  remark- 
ably successful  organization  it  eventually  became.  Certainly  the  experience 
turned  the  project's  leaders  into  some  of  the  most  reflective  of  NASA's  engi- 
neer/philosophers. Eugene  Schult,  who  was  responsible  for  Scout's  guidance 
and  control,  ponders  the  project,  "We  wouldn't  learn  anything  if  we  didn't 
have  problems;  that's  basic  in  engineering  training.  .  .  .  Success  doesn't  tell 
us  anything.  It  doesn't  tell  us  where  the  limits  are,  or  what  the  limiting 
aspects  of  the  envelope  are.  But  when  you  hit  a  mistake,  you  dig  into  it 
and  you  find  out  there's  a  weakness.  And  by  curing  weaknesses  you  get 

11 47 

success. 

Schult  and  his  Scout  group  did  indeed  recover  from  failure.  The  first  three 
launches  after  the  recertification — in  December  1963  (from  Vandenberg), 
March  1964  (from  Wallops),  and  June  1964  (from  Vandenberg) — were  all 
resounding  orbital  successes.  Between  July  1964  and  January  1967,  Scout 
established  a  record  of  22  consecutive  launches.  Only  one  of  the  16  recertified 
rockets  experienced  a  failure.  The  pressure  to  succeed  was  now  off.  Scout 
workers  no  longer  had  to  perform  failure  reviews  every  other  month,  and 

215 


Space/light  Revolution 

they  no  longer  had  to  work  the  endless  overtime  and  spend  weekends  away 
from  their  families.  In  such  a  positive  environment,  success  bred  success. 

"Now  we  really  had  the  kind  of  vehicle  we'd  set  out  to  develop,"  boasts 
Bud  English.  "Reliable.  It  was  still  simple  and  inexpensive,  but  we  could 
launch  [it]  quickly."48  In  fact,  the  Scout  group  needed  only  six  weeks  to 
process  one  of  the  rockets  for  a  successful  launch.  Even  with  this  short 
turnaround  time,  NASA  would  launch  this  little  rocket  for  10  years  without 
a  problem.  English  and  his  colleagues  had  indeed  done  the  job  they  set  out 
to  do. 


An  Unsung  Hero 

Scout  made  a  total  of  113  flights  under  NASA  Langley's  direction;  the 
last  one  before  the  official  transfer  of  the  program  to  NASA  Goddard  and 
LTV  took  place  on  9  May  1990  from  Vandenberg  AFB.  As  a  result  of  these 
flights,  NASA  engineers  and  their  contractors  authored  more  than  1300 
technical  and  scientific  reports  on  various  aspects  of  the  rocket's  design, 
performance,  and  mission  findings.49 

The  pride  that  the  Scout  Project  Group  felt  for  the  rocket's  performance 
sprang  not  only  from  its  phenomenal  post-recertification  accomplishment 
rate  of  22  and  37  successful  launches  but  also  from  the  critical  roles  played 
by  Scout  payloads  in  the  advancement  of  atmospheric  and  space  science. 
Early  Scout  missions  helped  researchers  study  the  density  of  the  atmosphere 
at  various  altitudes,  the  properties  of  the  Van  Allen  radiation  belts,  and  the 
possible  dangers  of  the  micrometeoroid  environment  on  spacecraft.  Scouts 
in  the  1970s  tested  Einstein's  theory  of  relativity  by  carrying  an  extremely 
accurate  atomic  clock  into  space,  and  they  also  helped  to  confirm  the  theory 
of  the  "black  hole." 

In  support  of  NASA's  early  space  program,  Scout  was  critical  to  the 
important  research  into  reentry  aerodynamics  for  the  manned  space  mis- 
sions. With  the  resulting  data,  NASA  researchers  determined  what  mate- 
rials best  withstood  the  heat  of  reentry.  This  information  as  well  as  other 
data  acquired  by  Scout  missions  contributed  directly  to  test  programs  such 
as  Projects  Fire  and  RAM  and  to  the  successes  of  Mercury,  Gemini,  and 
Apollo.  In  one  notable  mission  in  November  1970,  the  rocket  carried  two 
male  bullfrogs  into  orbit.  This  turned  out  to  be  the  only  time  a  Scout  satel- 
lite was  to  carry  a  living  payload.  The  unusual  mission  enabled  NASA  to 
study  the  effects  of  space  on  the  inner  ear  and  thereby  better  understand 
the  causes  of  the  space  sickness  experienced  by  astronauts. 

Scout  also  delivered  into  space  several  reconnaissance  and  communica- 
tions satellites.  For  the  DOD,  the  rocket  launched  classified  payloads;  for 
the  navy,  it  put  into  orbit  the  satellites  needed  for  its  Transit  system,  which 
by  the  late  1960s  provided  instantaneous  global  navigation  data  not  only  for 
the  operational  fleet  but  also  for  commercial  shipping  worldwide. 

216 


The  Early  Rush  of  the  Scout  Rocket  Program 

Much  of  Scout's  contribution  was  international:  the  rocket  launched  23 
satellites  for  foreign  countries,  including  Germany,  the  Netherlands,  France, 
and  the  United  Kingdom,  and  the  European  Space  Agency.  Based  on  a  1961 
agreement  between  the  United  States  and  Italy,  NASA  Langley  supplied 
Scouts  for  an  innovative  Italian  launch  operation  known  as  San  Marco,  which 
was  established  on  two  huge  mobile  platforms  in  the  Indian  Ocean,  3  miles  off 
the  coast  of  Kenya.  From  this  unusual  location  in  Ngwana  Bay,  the  Centre 
Italiano  Ricerche  Aerospaziali  (the  Center  for  Italian  Aerospace  Research), 
starting  in  April  1967,  used  NASA  Scouts  to  boost  an  international  series 
of  eight  spacecraft  into  orbit.  The  flights  of  these  spacecraft,  many  of 
which  were  placed  into  equatorial  orbits,  gathered  valuable  data  about  the 
ionosphere  and  the  magnetosphere,  about  the  galactic  sources  of  radiation 
and  X-rays,  and  especially  about  the  nature  of  the  earth's  atmosphere  in  the 
region  of  the  equator.  Participation  in  the  San  Marco  project  incidentally 
gave  some  Langley  engineers  their  first  opportunity  for  foreign  travel  and 
international  cooperation.  In  1966  it  even  afforded  some  of  them  the  rare 
opportunity  of  an  audience  with  Pope  Paul  VI,  who  blessed  the  rocket. 
Fortunately,  the  launch  of  the  anointed  Scout  went  well.50 

Over  the  years,  through  the  waning  of  the  Apollo  program  and  into 
the  era  dominated  by  the  Space  Shuttle,  Scout  became  more  of  a  bargain. 
Improvements  in  its  stage  motors  enabled  the  rocket  to  carry  larger  pay  loads, 
but  costs  remained  low.*  Using  the  consumer  price  index,  Langley  employees 
hoping  to  retain  the  Scout  program  calculated  that  a  Scout  cost  less  when 
NASA  Goddard  took  over  the  program  (and  LTV  took  over  the  rocket)  in 
1991  than  the  original  $4  million  invested  in  it  in  1958. 51 

In  summary,  Scout,  although  virtually  unknown  outside  NASA  circles, 
developed  into  one  of  the  finest  pieces  of  technology  in  the  history  of  space 
exploration.  As  Tom  Perry  has  observed  about  the  evolution  of  his  most 
cherished  rocket,  "The  Scout  became  so  reliable  that  mission  planners  could 
take  it  for  granted.  They  focused  on  the  science  of  the  satellite  pay  load 
rather  than  on  its  transportation  system.  ...  It  happens  to  be  NASA's 
smallest  launch  vehicle  and  it  does  not  receive  the  same  level  of  notoriety 
you  would  with  a  larger  system.  But  over  the  years  it  has  proven  to  be  a  very 
reliable,  consistent,  performing  warhorse."  As  Perry  and  other  Scout  people 
at  Langley,  Houston,  Wallops  Island,  Vandenberg  AFB,  and  San  Marco  are 
still  fond  of  saying,  more  than  30  years  after  its  first  countdown,  Scout  is 
"the  unsung  hero  of  space."52 


sk 

For  the  first  10  production  Phase  II  Scouts  (Phase  I  was  the  developmental  phase),  the  vehicle 
hardware  costs  amounted  to  $0.96  million  per  vehicle;  for  the  next  14  production  Scouts  (Phase  III),  the 
cost  per  vehicle  rose  to  $1.42  million.  Costs  decreased  for  the  25  Phase  IV  rockets  (provided  by  LTV) 
to  $1.19  million  per  vehicle.  Costs  for  later  Scouts  rose  only  slightly,  and  stayed  under  $1.5  million  per 
vehicle. 


217 


Space/light  Revolution 


From  Italy's  innovative  San  Marco 
launch  operation  in  the  Indian 
Ocean,  NASA  Langley  helped  to 
launch  an  international  series  of 
eight  spacecraft  into  orbit.  A  huge 
mobile  launcher  lifts  Scout  into  fir- 
ing position  (right);  the  San  Marco 
platform  floats  securely  in  inter- 
national waters  in  Ngwana  Bay 
(below). 


L-74-7828 


L-71-1197 


218 


The  Early  Rush  of  the  Scout  Rocket  Program 
Postscript 

A  cynic  might  suggest  that  it  was  entirely  in  keeping  with  Scout's 
difficult  and  publicly  unappreciated  sojourn  into  space  that  the  project 
ended  as  it  did.  In  the  late  1970s,  NASA  policymakers  proposed  to 
launch  all  future  NASA  satellites  using  the  Space  Transportation  System 
(STS)  still  under  development  and  abolish  all  expendable  launch  vehicles; 
the  Space  Shuttle,  when  fully  operational,  could  do  it  all.  Only  the 
Challenger  explosion  in  1986,  which  underscored  the  need  for  alternative 
launch  capabilities,  reversed  the  shortsighted  policy.  In  the  aftermath  of 
the  Challenger  accident,  and  in  league  with  the  Reagan  administration's 
objectives  for  the  commercialization  of  space  and  the  privatization  of  many 
government  services,  NASA  created  the  "Mixed  Fleet"  concept.  Under  this 
plan,  NASA  was  to  give  up  its  other  launch  services  to  commercial  firms, 
which  from  then  on  were  to  handle  whatever  NASA  payloads  the  Shuttle 
could  not  carry.  Essentially,  this  meant  the  end  of  the  expendable  launch 
vehicle  business  as  NASA's  Scout  Project  Group  had  known  and  developed 
it.53 

Scout  engineers  sorely  lamented  the  loss  of  Scout.  For  them,  the  ven- 
ture into  space  had  come  to  mean  an  all-enveloping  system  and  a  rigorous 
discipline:  a  government-driven  version  for  rockets  of  Henry  Ford's  mass  pro- 
duction. "Other  programs  are  full  of  changes  and  improvisations,"  declares 
James  Hall;  they  are  always  "borrowing  from  other  missiles  and  assembling 
something  just  to  get  it  delivered  on  schedule" — which  is  exactly  what  the 
Scout  team  itself  had  been  doing  in  the  pre-recertification  days.54  Over 
time,  however,  the  "cannibalizing"  became  minimal  in  Scout.  The  program 
for  rocket  assembly  matured  beyond  the  practice,  thus  becoming  standard 
almost  to  the  point  of  stereotype.  Scout  engineers  wanted  to  produce  a 
launch  vehicle  that  was  as  reliable  for  a  trip  to  space  as  an  automobile  was 
for  a  trip  to  town.  Scout,  like  the  Ford  Model  T,  was  the  "poor  man's 
rocket." 

Learning  hard  lessons  through  failure  and  then  enjoying  such  incredible 
long-term  success  made  losing  the  rocket  especially  difficult  for  the  Scout 
Project  Group.  Scout  had  been  giving  the  country  access  to  space  for  more 
than  30  years.  It  succeeded  in  spite  of — and  ironically  perhaps  because  of— 
its  hurried  early  development.  Not  many  programs  born  of  the  spaceflight 
revolution  survived  the  spaceflight  revolution;  Scout  was  one. 


219 


8 


Enchanted  Rendezvous: 
The  Lunar- Or  bit  Rendezvous  Concept 


There  was  a  reluctance  to  believe  that  the  rendezvous 
maneuver  was  an  easy  thing.  In  fact,  to  a  layman, 
if  you  were  to  explain  what  you  had  to  do  to  perform 
a  rendezvous  in  space,  he  would  say  that  sounds  so 
difficult  we'll  never  be  able  to  do  it  this  century. 

—Clinton  E.  Brown,  head,  Langley 
Lunar  Mission  Steering  Group 
on  Trajectories  and  Guidance 

I'm  not  so  sure  we  ever  thought  of  rendezvous  as  very 
complicated.  It's  an  amazing  thing.  We  thought  that 
if  our  guys  could  work  out  the  orbital  mechanics  and 
we  gave  the  pilot  the  right  controls  and  stuff,  then  he  'd 
land  it  and  make  the  rendezvous.  We  didn't  think  it 
was  very  complicated. 

— Arthur  W.  Vogeley,  head,  Langley 
Guidance  and  Control  Branch 


On  Thursday  morning,  25  May  1961,  in  a  speech  to  a  joint  session  of 
Congress,  President  John  F.  Kennedy  challenged  the  American  people  to 
rebound  from  their  recent  second-place  finishes  in  the  space  race:  "First,  I 
believe  that  this  nation  should  commit  itself  to  achieving  the  goal,  before 
this  decade  is  out,  of  landing  a  man  on  the  moon  and  returning  him 
safely  to  earth.  No  single  space  project  . . .  will  be  more  exciting,  or 
more  impressive  ...  or  more  important  . . .  and  none  will  be  so  difficult  or 
expensive."  "It  will  not  be  one  man  going  to  the  Moon,"  the  dynamic 
43-year-old  president  told  his  countrymen,  "it  will  be  an  entire  nation.  For 
all  of  us  must  work  to  put  him  there."1 

221 


Space/light  Revolution 

At  first  no  one  at  Langley  could  quite  believe  it.  If  President  Kennedy 
had  in  fact  just  dedicated  the  country  to  a  manned  lunar  landing,  he  could 
not  be  serious  about  doing  it  in  less  than  nine  years.  NASA  had  been 
studying  the  feasibility  of  various  lunar  missions  for  some  time,  but  had 
never  dreamed  of  a  manned  mission  that  included  landing  on  and  returning 
from  the  surface  of  the  moon  by  the  end  of  the  1960s.  NASA  was  not  exactly 
sure  how  such  a  lunar  mission  could  be  achieved,  let  alone  in  so  little  time. 

Not  even  Bob  Gilruth,  the  leader  of  the  STG,  was  prepared  for  the  sen- 
sational announcement.  He  heard  the  news  in  a  NASA  airplane  somewhere 
over  the  Midwest  on  his  way  to  a  meeting  in  Tulsa.  He  knew  that  Kennedy 
planned  to  say  something  dramatic  about  the  space  program  in  his  speech, 
and  so  he  asked  the  pilot  to  patch  it  through  on  the  radio.  Looking  out 
the  window  over  the  passing  clouds,  he  had  heard  every  incredible  word. 
Only  one  word  described  Gilruth's  feelings  at  that  moment:  "aghast."  The 
first  manned  Mercury  flight  by  Alan  Shepard  had  taken  place  only  three 
weeks  before,  on  5  May.  NASA  had  made  this  one  brief  15-minute  subor- 
bital  flight,  and  suddenly  the  President  was  promising  Americans  the  moon. 
The  audacity  of  the  goal  was  stunning.2  American  astronauts  would  fly  a 
quarter  of  a  million  miles,  make  a  pinpoint  landing  on  a  familiar  but  yet  so 
strange  heavenly  body,  blast  off,  and  return  home  safely  after  a  voyage  of 
several  days  through  space,  and  do  it  all  by  the  end  of  the  decade.  Only 
one  thought  was  more  daunting  to  Gilruth,  and  that  was  that  he  was  one 
of  the  main  people  who  would  have  to  make  it  happen.  Already  the  STG 
had  its  hands  full  preparing  for  another  suborbital  flight  (Virgil  I.  "Gus" 
Grissom's,  on  21  July)  and  for  the  first  orbital  flight  sometime  early  in  the 
next  year  (John  Glenn's,  on  20  February  1962).  Gilruth  himself,  before 
the  president's  announcement,  "had  spent  almost  no  time  at  all"  on  lunar 
studies,  so  demanding  were  the  activities  of  Project  Mercury.3 

Only  the  project  managers  directly  responsible  for  making  Mercury  a 
success  felt  burdened  by  the  prospects  of  now  having  to  fulfill  the  lunar 
commitment.  Other  planners  and  dreamers  about  space  exploration  within 
NASA  were  elated. 

When  they  heard  about  Kennedy's  announcement,  Clinton  E.  Brown  and 
his  adventurous  colleagues  cheered,  "Hooray,  let's  put  on  full  speed  ahead, 
and  do  what  we  can."  To  them,  landing  astronauts  on  the  moon  as  quickly 
as  possible  was  obviously  the  next  step  if  the  United  States  was  going  to 
win  the  space  race.  Furthermore,  Brown  and  his  little  band  of  men — plus 
one  other  key  Langley  researcher,  Dr.  John  C.  Houbolt — were  confident  that 
they  already  knew  the  best  way  to  accomplish  the  lunar  goal.4 

Brown's  Lunar  Exploration  Working  Group 

After  Sputnik,  a  small  circle  of  Langley  researchers  had  plunged  into  the 
dark  depths  of  space  science.  "We  were  aeronautical  engineers,"  remembers 

222 


Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 

William  H.  Michael,  Jr.,  a  member  of  Brown's  division  who  had  recently 
returned  to  Langley  after  a  two-year  stint  in  the  aircraft  industry.  "We 
knew  how  to  navigate  in  the  air,  but  we  didn't  know  a  thing  about  orbital 
mechanics,  celestial  trajectories,  or  interplanetary  travel,  so  we  had  to  teach 
ourselves  the  subjects."  In  the  Langley  Technical  Library,  Michael  could  find 
only  one  pertinent  book,  An  Introduction  to  Celestial  Mechanics,  written 
in  1914  by  British  professor  of  astrophysics  Forrest  R.  Moulton  (someone 
Michael  had  never  heard  of).5  With  this  out-of-date  text,  Michael  and  a 
few  associates  taught  themselves  enough  about  the  equations  of  celestial 
mechanics  to  grow  confident  in  their  computations.  Before  long,  the  novices 
had  transformed  themselves  into  experts  and  were  using  their  slide  rules  and 
early  electronic  computers  to  calculate  possible  paths  to  the  moon. 

In  anticipating  the  trajectories  for  lunar  missions  in  the  late  1950s, 
Brown,  Michael,  and  the  few  others  were  leapfrogging  over  what  most  people 
considered  to  be  "the  logical  next  step"  into  space:  an  earth-orbiting  space 
station.  Little  did  they  know  that  their  mental  gymnastics  would  set  the 
direction  of  the  U.S.  space  program  for  the  next  30  years. 

Following  the  wisdom  of  Konstantin  Tsiolkovskii,  Hermann  Oberth, 
Guido  von  Pirquet,  Wernher  von  Braun,  and  other  spacefaring  visionaries, 
most  proponents  of  space  travel  believed  that  the  first  step  humans  would 
take  into  the  universe  would  be  a  relatively  timid  one  to  some  sort  of  space 
station  in  earth  orbit.  The  station  could  serve  as  a  research  laboratory 
for  unique  experiments  and  valuable  industrial  enterprises,  and  from  this 
outpost,  human  travelers  could  eventually  venture  into  space  using  craft 
for  trips  to  the  moon,  the  planets,  and  beyond.  Most  NASA  researchers 
believed  that  the  space  station  was  the  perfect  target  project  because  it 
could  focus  NASA's  space-related  studies  as  well  as  its  plans  for  future 
space  exploration.6 

Clint  Brown  and  associates  felt  differently:  they  thought  that  the  space 
station  step  must  be  skipped.  The  politics  of  the  space  race,  not  the  inspired 
prophecies  of  the  earliest  space  pioneers,  were  dictating  the  terms  of  our 
space  program.  The  Russians  had  already  demonstrated  that  they  had 
larger  boosters  than  the  United  States.  This  meant  that  they  had  the 
capability  of  establishing  a  space  station  first.  As  Brown  explains,  "If  we  put 
all  our  efforts  into  putting  a  space  station  around  the  world,  we'd  probably 
find  ourselves  coming  in  second  again."  The  "obvious  answer"  was  that 
"you  had  to  take  a  larger  bite  and  decide  what  can  really  give  us  leadership 
in  the  space  race."  To  him  "that  clearly  seemed  the  possibility  of  going  to 
the  moon  and  landing  there."7 

Inside  Brown's  Theoretical  Mechanics  Division,  the  conviction  that  lu- 
nar studies  should  take  precedence  over  space  station  studies  grew.  In  early 
1959,  Langley's  assistant  director,  Eugene  Draley,  agreed  to  form  a  Langley 
working  group  to  study  the  problems  of  lunar  exploration.  Brown,  the  cat- 
alytic group  leader,  asked  for  the  participation  of  six  of  Langley's  most 
thoughtful  analysts:  David  Adamson,  Supersonic  Aerodynamics  Division; 

223 


Space/light  Revolution 

Paul  R.  Hill,  PARD;  John  C.  Houbolt,  Dynamic  Loads  Division;  Albert  A. 
Schy,  Stability  Research  Division;  Samuel  Katzoff,  Full-Scale  Research  Divi- 
sion; and  Bill  Michael  of  his  own  Theoretical  Mechanics  Division.  Leonard 
Roberts,  a  talented  young  mathematician  from  England,  eventually  joined 
the  group.  Brown  assembled  these  researchers  for  the  first  time  in  late  March 
1959  and  periodically  into  1960.  Besides  advising  Langley  management  on 
the  establishment  of  lunar-related  research  programs,  Brown's  group  also 
organized  a  course  in  space  mechanics  for  interested  employees.  For  many, 
this  course  provided  their  first  real  brush  with  relativity  theory.  The  Brown 
study  group  even  worked  to  disseminate  information  about  the  moon  by 
holding  public  seminars  led  by  experts  from  Langley  and  from  the  nearby 
universities.8 

Everything  about  this  original  lunar  study  group  was  done  quietly 
and  without  much  fuss.  In  those  early  days  of  NASA,  the  management 
of  research  was  still  flexible  and  did  not  always  require  formal  research 
authorizations  or  approval  from  NASA  headquarters  in  Washington.  When 
Brown  expressed  his  desire  to  work  more  on  lunar  exploration  than  on  the 
space  station,  Draley  simply  told  him,  "Fine,  go  ahead."  Henceforth,  he  and 
his  lunar  working  group  proceeded  with  their  efforts  to  solve  the  problems  of 
sending  an  American  to  the  moon.  Brown's  group  was  doing  what  Langley 
researchers  did  best:  exploring  an  interesting  new  idea  and  seeing  how  far 
they  could  go  with  it. 

Langley  researchers  were  not  the  only  people  in  the  United  States  think- 
ing seriously  about  lunar  missions.  Officers  in  the  air  force,  scientists  in  think 
tanks,  professors  at  universities,  and  other  engineers  and  researchers  in  and 
around  NASA  were  all  contemplating  a  journey  to  the  moon.  In  February 
1959,  a  month  before  the  creation  of  Brown's  Lunar  Exploration  Working 
Group  at  Langley,  NASA  headquarters  had  created  a  small  "Working  Group 
on  Lunar  and  Planetary  Surfaces  Exploration"  (evolving  later  into  the 
"Science  Committee  on  Lunar  Exploration")  chaired  by  Dr.  Robert  Jastrow, 
the  head  of  NASA  headquarters'  new  Theoretical  Division.  This  group  in- 
cluded such  leaders  in  planetology  and  lunar  science  as  Harold  C.  Urey, 
professor  at  large  at  the  University  of  California  at  San  Diego,  several  lead- 
ing scientists  from  JPL  in  Pasadena,  and  a  few  from  Langley.  In  their 
meetings  Jastrow's  group  looked  into  the  feasibility  of  both  "rough"  (later 
usually  called  "hard")  and  "soft"  landings  on  the  moon.  In  a  rough  landing, 
a  probe  would  crash  onto  the  surface  and  be  destroyed,  but  only  after  an 
on-board  camera  had  sent  back  dozens  of  valuable  pictures  to  earth.  In  a 
soft  landing,  a  spacecraft  would  actually  land  intact  on  the  moon.  Langley's 
Bill  Michael  sat  in  on  one  of  the  first  meetings  of  the  Jastrow  Committee. 
In  reaction  to  what  he  heard,  Michael  and  others  at  Langley  began  develop- 
ing ideas  for  photographic  reconnaissance  of  the  moon's  surface  from  lunar 
orbit  as  well  as  for  lunar  impact  studies.9  Houbolt,  of  Langley's  Dynamic 
Loads  Division,  also  attended  some  of  these  meetings  to  share  his  budding 
knowledge  of  the  requirements  for  spacecraft  rendezvous. 

224 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 


Committees  Reviewing  Lunar  Landing  Modes 


Date 
Formed 

Location 

Title 

Chairman 

Feb.  1959 

NASA  HQ 

Working  Group  on  Lunar 
and  Planetary  Surfaces 
Exploration 

Jastrow 

Mar.  1959 

Langley 

Lunar  Exploration  Working 
Group 

Brown 

Apr.  1959 

NASA  HQ 

Research  Steering  Committee 
on  Manned  Space  Flight 

Goett 

Summer 
1959 

Langley 

Manned  Space  Lab  Group 
Subcommittee:  Rendezvous 

Nichols 
Houbolt 

May  1960 

Langley 

Intercenter  Review  of 
Rendezvous  Studies 

Maggin 

May  1960 

Langley 

Lunar  Mission  Steering  Group 
Subcommittee: 
Trajectories  and  Guidance 
Subcommittee: 
Rendezvous 

Becker 
Brown 

Houbolt 

Oct.  1960 

NASA  HQ 

Manned  Lunar  Landing  Task 
Group  (Low  Committee) 

Low 

May  1961 

NASA  HQ 

Ad  Hoc  Task  Group  for  a 
Lunar  Landing  Study 
(Fleming  Committee) 

Fleming 

May  1961 

NASA  HQ 

Lundin  Committee 

Lundin 

June  1961 

NASA  HQ 

Ad  Hoc  Task  Group  for 
Study  of  Manned  Lunar 
Landing  by  Rendezvous 
Techniques 
(Heaton  Committee) 

Heaton 

July  1961 

NASA  HQ 

NASA/DOD  Large  Launch 
Vehicle  Planning.  Group 
(Golovin  Committee) 

Golovin 

Dec.  1961 

NASA  HQ 

Manned  Space  Flight 
Management  Council 

Holmes 

225 


Space/light  Revolution 

Two  months  later,  in  April  1959,  NASA  headquarters  formed  a  Research 
Steering  Committee  on  Manned  Space  Flight.  Chaired  by  Harry  J.  Goett 
of  NASA  Goddard,  this  committee  was  to  review  man-in-space  problems, 
recommend  the  missions  to  follow  Project  Mercury,  and  outline  the  research 
programs  to  support  those  missions.10 

In  its  final  report,  which  came  at  the  end  of  1959,  the  Goett  Committee 
called  for  a  manned  lunar  landing  as  the  appropriate  long-term  goal  of 
NASA's  space  program.  Between  that  goal  and  the  present  Project  Mercury, 
however,  a  major  interim  program  designed  to  develop  advanced  orbital 
capabilities  and  a  manned  space  station  was  needed.  Before  that  program, 
to  be  named  Gemini,  took  shape,  however,  basic  priorities  would  change. 

Langley's  representative  on  the  Goett  Committee,  Laurence  K.  Loftin, 
Jr.,  the  technical  assistant  to  Associate  Director  Thompson,  agreed  that  the 
space  station  should  be  NASA's  immediate  goal.  But  two  other  members 
disagreed:  the  STG's  Max  Faget  and  George  Low,  NASA's  director  of 
spacecraft  and  flight  missions  in  Washington.  During  meetings  from  May 
to  December,  they  voiced  what  turned  out  to  be  the  minority  opinion  that 
the  moon  should  be  NASA's  next  objective.  George  Low  was  particularly 
vocal  in  making  the  point.  Not  only  did  he  want  to  go  to  the  moon,  Low 
also  wanted  to  land  on  it,  with  men,  and  the  sooner  the  better.11 


Michael's  Paper  on  a  "Parking  Orbit" 

At  Langley,  members  of  Brown's  lunar  exploration  group  were  studying 
ways  of  accomplishing  Low's  dream.  One  of  these  studies,  by  Bill  Michael, 
examined  the  benefits  of  "parking"  the  earth-return  propulsion  portion  of  a 
spacecraft  in  orbit  around  the  moon  during  a  landing  mission. 

The  spark  for  Michael's  interest  in  what  came  to  be  called  a  "parking 
orbit,"  a  spacecraft  in  a  waiting  orbit  around  the  moon  or  some  other 
celestial  body,  was  calculations  he  had  made  to  see  whether  any  advantage 
could  be  gained  in  a  lunar  mission  from  additional  "staging."  First  explained 
by  Tsarist  Russia's  space  visionary  Tsiolkovskii  in  the  late  1800s,  staging 
was  the  proven  technological  concept  by  which  a  self-propelled,  staged- 
rocket  vehicle  (Tsiolkovskii  called  it  a  rocket  "train" )  could  ascend  to  greater 
heights  as  its  stages  expended  their  fuel  and  separated. 

In  a  lunar  landing  mission,  Michael  speculated,  flying  a  big  rocket  ship 
directly  from  the  earth  to  the  moon  would  be  impractical.  (Jules  Verne's 
popular  book  and  other  science-fiction  fantasies  had  pictured  this  method 
for  a  lunar  landing.)  Too  much  unnecessary  weight  would  have  to  be 
transported  to  the  moon's  surface.  How  much  wiser  it  would  be  to  make  "an 
intermediate  step"  and  place  the  vehicle  in  lunar  orbit  where  much  of  the 
total  weight  remained  behind  including  the  structure  of  the  interplanetary 
spacecraft,  its  heavy  fuel  load  for  leaving  lunar  orbit  and  returning  home, 
and  its  massive  heat  shield  necessary  for  a  safe  reentry  into  the  earth's 

226 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 


L-89-8683 

At  a  colloquium  held  at  Langley  on  20  July  1989  to  celebrate  the  20th  anniversary  of 
the  first  lunar  landing,  William  H.  Michael,  Jr.,  (center)  reviews  the  evolution  of  his 
parking  orbit  concept  with  Clinton  E.  Brown  (right),  head  of  the  Lunar  Exploration 
Working  Group  and  Arthur  Vogeley  (left),  mastermind  of  Langley 's  rendezvous  and 
docking  simulators  of  the  1960s. 


atmosphere.  "It's  very  expensive  to  accelerate  any  type  of  mass  to  high 
velocity,"  Michael  thought.  "Any  time  you  do  not  have  to  do  that,  you  save 
a  lot  of  fuel  and  thus  a  lot  of  weight." 

Michael  wrote  his  calculations  in  1960  in  a  never-to-be-published  paper, 
"Weight  Advantages  of  Use  of  Parking  Orbit  for  Lunar  Soft  Landing 
Mission."  In  the  paper,  Michael  identified  the  most  basic  advantage  of  what 
came  to  be  known  as  lunar-orbit  rendezvous  (LOR).  His  results  implied 
that  LOR  could  save  NASA  an  impressive  50  percent  or  more  of  the  total 
mission  weight.  Figuring  the  numbers  did  not  require  any  difficult  or 
sophisticated  calculations.  Nor  did  it  require  any  knowledge  of  the  writings 
of  Russian  rocket  theoretician  Yuri  Kondratyuk  and  British  scientist  and 
Interplanetary  Society  member  H.  E.  Ross,  both  of  whom  had  expressed  the 
fundamentals  of  the  LOR  concept  years  earlier  (Kondratyuk  in  1916,  and 
Ross  in  1948). 14  Neither  Michael  nor  anyone  else  at  Langley  at  this  point, 
so  they  have  always  maintained,  had  any  knowledge  of  those  precursors. 

They  also  knew  nothing  about  competition  from  contemporaries;  how- 
ever, they  soon  would.  The  same  morning  that  Michael  first  showed  his 

227 


Spaceflight  Revolution 

rough  parking-orbit  calculations  to  Clint  Brown,  a  team  led  by  Thomas  E. 
Dolan  from  Vought  Astronautics,  a  division  of  the  Chance  Vought  Cor- 
poration in  Dallas,  gave  a  briefing  at  Langley.  The  briefing  concerned 
Vought 's  ongoing  company- funded,  confidential  study  of  problems  related 
to  Manned  Lunar  Landing  and  Return  (MALLAR)  and  specifically  its  plans 
for  a  manned  spaceflight  simulator  and  its  possible  application  for  research 
under  contract  to  NASA.15 

During  the  briefing,  Dolan's  staff  mentioned  an  idea  for  reaching  the 
moon.  Although  the  Vought  representatives  focused  their  analysis  on  the 
many  benefits  of  what  they  called  a  "modular  spacecraft" — one  in  which 
several  parts,  including  a  lunar  landing  module,  were  designed  for  certain 
tasks — Brown  and  Michael  understood  that  Vought  was  advertising  the 
essentials  of  the  LOR  concept.  "They  got  up  there  and  they  had  the  whole 
thing  laid  out,"  Brown  remembers.  "They  had  scooped  us"  with  their  idea 
of  "designing  a  spacecraft  so  that  you  can  throw  away  parts  of  it  as  you  go 
along."  For  the  next  several  days,  Michael  walked  around  "with  his  face 
hanging  down  to  the  floor."16 

Nevertheless,  the  chagrined  Langley  engineer  decided  to  write  a  brief 
paper  because  he  was  confident  that  he  had  come  up  with  his  idea  indepen- 
dently. Furthermore,  the  word  around  Langley  later  came  to  be  that  Dolan 
had  developed  the  idea  of  using  a  detachable  lunar  landing  module  for  the 
landing  operation  after  an  earlier  visit  to  Langley  when  PARD  engineers  fa- 
miliar with  Michael's  embryonic  idea  had  suggested  a  parking  orbit  to  him. 
This  explanation  may  simply  be  "sour  grapes."  On  the  other  hand,  Dolan 
had  been  visiting  Langley  in  late  1959  and  early  1960,  and  Michael  does 
remember  having  already  mentioned  his  idea  to  a  few  people  at  the  center, 
"so  it  shouldn't  have  been  any  surprise  to  anybody  here  at  Langley  that 
such  a  possibility  existed."17  The  truth  about  the  origin  of  Dolan's  idea 
will  probably  never  be  known. 

Michael's  paper,  at  least  in  retrospect,  had  some  significant  limitations. 
It  was  only  two  pages  long  and  presented  little  analysis.  Its  charts  were  dif- 
ficult to  follow  and  interpret.  He  did  not  mention  "earth-escape  weights," 
though  an  informed  reader  could  infer  such  numbers.  Perhaps  most  impor- 
tantly, the  paper  did  not  explicitly  mention  either  the  need  for  a  separate 
lunar  lander  or  the  additional  weight  savings  derived  from  using  one  and 
discarding  it  before  the  return  trip  home.  A  reader  would  already  have  to 
be  familiar  with  the  subject  even  to  recognize,  let  alone  fully  fathom,  what 
was  being  implied.  Michael's  paper  was  hardly  a  fully  developed  articulation 
of  a  lunar  landing  mission  using  LOR.  Nonetheless,  it  made  a  fundamentally 
important  contribution:  it  made  rendezvous  the  central  theme  for  Langley 
researchers  contemplating  lunar  missions.  As  his  paper  concluded,  the  chief 
problems  in  a  lunar  landing  mission  were  the  "complications  involved  in 
requiring  a  rendezvous  with  the  components  left  in  the  parking  orbit."18 

Although  disappointed  by  the  news  that  Vought.  had  scooped  them 
with  the  idea  of  LOR,  the  Langley  researchers  were  hardly  demoralized. 

228 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 


Three  days  before  President  Kennedy's 
lunar  commitment,  John  D.  Bird, 
"Jaybird"  (left),  captured  Langley's  en- 
thusiasm for  a  moonshot  in  his  sketch 
"TO  THE  MOON  WITH  C-l's  OR 
BUST"  (below).  In  essence,  his  plan 
called  for  a  mission  via  earth- orbit  ren- 
dezvous (EOR)  requiring  the  launch  of 
10  C-l  rockets. 


L-62-8189 


/    >-*.J 

/      r'v      '       y 

-".  £4  ^Tnr 


I    C-l 


^c- 


-3 


229 


Spaceflight  Revolution 

Researchers  in  and  around  Brown's  division  quickly  began  making  lunar 
and  planetary  mission  feasibility  studies  of  their  own.  John  P.  Gapcynski, 
for  example,  considered  factors  involved  in  the  departure  of  a  vehicle  from  a 
circular  orbit  around  the  earth.  Wilbur  L.  Mayo  calculated  energy  and  mass 
requirements  for  missions  to  the  moon  and  even  to  Mars.  Robert  H.  Tolson 
studied  the  effects  on  lunar  trajectories  of  such  geometrical  constraints  as 
the  eccentricity  of  the  moon's  orbit  and  the  oblate  shape  of  the  earth,  and 
also  looked  into  the  influence  of  the  solar  gravitational  field.  John  D.  Bird, 
"Jaybird,"  who  worked  across  the  hall  from  Michael,  began  designing  "lunar 
bugs,"  "lunar  schooners,"  and  other  types  of  small  excursion  modules  that 
could  go  down  to  the  surface  of  the  moon  from  a  "mother  ship."  Jaybird 
became  a  particularly  outspoken  advocate  of  LOR.  When  a  skeptical  visitor 
to  Langley  offered,  with  a  chuckle,  that  LOR  was  "like  putting  a  guy  in 
an  airplane  without  a  parachute  and  having  him  make  a  midair  transfer," 
Bird  set  the  visitor  straight.  "No,"  he  corrected,  "it's  like  having  a  big  ship 
moored  in  the  harbor  while  a  little  rowboat  leaves  it,  goes  ashore,  and  comes 
back  again."19 


The  Rendezvous  Committees 

A  feeling  was  growing  within  NASA  in  late  1959  and  early  1960  that 
rendezvous  in  space  was  going  to  be  a  vital  maneuver  no  matter  what 
NASA  chose  as  the  follow-on  mission  to  Project  Mercury.  If  the  next 
step  was  a  space  station,  a  craft  must  meet  and  dock  with  that  station 
and  then  leave  it;  if  the  next  step  was  a  lunar  mission,  that,  too,  would 
require  some  sort  of  rendezvous  either  in  lunar  orbit,  as  Michael's  study 
suggested,  or  in  earth  orbit,  where  a  lunar-bound  spacecraft  might  be 
assembled  or  at  least  fueled.  Even  if  neither  of  these  projects  was  adopted, 
communications  and  military  "spy"  satellites  would  require  inspection  and 
repair,  thus  necessitating  rendezvous  maneuvers.  Rendezvous  would  be  a 
central  element  of  all  future  flight  endeavors — whatever  NASA  decided. 

By  late  summer  1959,  Langley's  senior  staff  was  ready  to  proceed  with 
detailed  studies  of  how  best  to  perform  rendezvous  maneuvers  in  space.  Two 
rendezvous  study  committees  eventually  were  formed,  both  chaired  by  Dr. 
John  C.  Houbolt,  the  assistant  chief  of  Langley's  Dynamic  Loads  Division. 

Houbolt  was  an  aircraft  structures  expert  who  had  begun  work  at  Langley 
in  1942  with  a  B.S.  and  M.S.  in  civil  engineering  from  the  University 
of  Illinois.  In  contrast  to  most  Langley  researchers,  he  had  spent  a 
significant  amount  of  time  conducting  research  abroad.  He  had  been  an 
exchange  research  scientist  at  the  British  Royal  Aircraft  Establishment  at 
Farnborough,  England,  in  1949,  and  in  1958,  Houbolt  had  only  recently 
returned  from  a  year  at  the  Swiss  Federal  Polytechnic  Institute  in  Zurich, 
where  his  dissertation  on  the  heat-related  aeroelastic  problems  of  aircraft 
structures  in  high-speed  flight  had  earned  him  a  Ph.D.20 

230 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

Upon  returning  from  his  graduate  work  in  Switzerland,  Houbolt  had 
found  himself  becoming  more  curious  about  spaceflight  as  were  other 
Langley  researchers.  On  his  own,  largely  independent  of  the  conversations 
taking  place  within  Brown's  group,  Houbolt  learned  the  fundamentals  of 
space  navigation.  "I  racked  down  and  went  through  the  whole  analysis  of 
orbital  mechanics  so  I  could  understand  it."  Prom  his  own  preliminary  stud- 
ies of  trajectories,  he  saw  the  vital  importance  of  rendezvous  and  began  to 
recognize  and  evaluate  the  basic  problems  associated  with  it.  During  the 
STG's  training  of  the  Mercury  astronauts  at  Langley,  Houbolt  taught  them 
their  course  on  space  navigation.21 

Houbolt  focused  on  one  special  problem  related  to  rendezvous — the 
timing  of  the  launch.  NASA  could  not  launch  a  mission  at  just  any  time  and 
be  assured  of  effecting  a  rendezvous  with  an  orbiting  spacecraft.  In  order 
to  visualize  the  problem,  Houbolt  built  a  gadget  with  a  globe  for  the  earth 
and  a  small  ball  on  the  end  of  a  short  piece  of  coat  hanger  for  the  satellite. 
He  connected  it  all  to  a  variable-ratio  gearbox.  The  gadget  simulated  a 
satellite  at  different  altitudes  and  in  different  orbital  planes.  With  this 
little  machine  Houbolt  could  figure  the  time  that  satellites  would  take  at 
varying  altitudes  to  orbit  the  revolving  earth.  From  his  considerations  of 
orbital  mechanics,  Houbolt  found  that  a  change  in  orbital  plane  at  25,000 
feet  per  second  without  the  help  of  aerodynamic  lift  would  require  such  an 
enormous  amount  of  energy  that  it  could  not  be  made.  With  this  simple 
but  ingenious  model,  Houbolt  saw  how  long  NASA  might  have  to  wait — a 
period  of  many  days — in  order  to  launch  a  rendezvous  mission  from  Cape 
Canaveral.  However,  he  also  found  a  way  to  circumvent  the  problem:  "if  the 
orbital  plane  of  the  satellite  could  be  made  just  one  or  two  degrees  larger 
than  the  latitude  of  the  launch  site,"  the  launch  "window"  could  be  extended 
to  four  hours  every  day.  Thus,  he  began  to  understand  how  NASA  could 
avoid  the  long  waiting  periods.22 

The  word  quickly  spread  through  Langley  that  Houbolt,  the  aircraft 
structures  specialist,  was  now  "the  rendezvous  man."  He  even  had  a 
"license  to  rendezvous"  issued  to  him  by  the  Rand  Corporation,  a  nonprofit 
think  tank  (affiliated  with  Douglas  Aircraft)  in  southern  California.  The 
Rand  Corporation,  which  had  an  interest  in  space  rendezvous  and  a  space 
rendezvous  simulator,  presented  Houbolt  with  this  "license"  in  November 
1959  after  he  successfully  linked  two  craft  on  the  Douglas  rendezvous 
simulator.23  Thus,  when  NASA  Langley  created  its  steering  groups  to 
study  the  problems  of  orbital  space  stations  and  lunar  exploration  missions, 
Houbolt  naturally  was  asked  to  provide  the  input  about  rendezvous. 

The  first  of  Houbolt 's  rendezvous  committees  was  linked  to  Langley 's 
Manned  Space  Laboratory  Group.  Headed  by  the  Full- Scale  Research 
Division's  Mark  R.  Nichols,  an  aerodynamics  specialist  who  was  reluctant 
to  accept  the  assignment,  this  group  was  formed  late  in  the  summer  of 
1959.  It  was  similar  to  Brown's  interdivisional  Lunar  Exploration  Working 
Group,  except  that  it  was  larger  and  had  committees  of  its  own.  One  of 

231 


Space/light  Revolution 

them,  Houbolt's  committee,  was  to  look  into  the  matter  of  rendezvous  as 
it  pertained  to  earth-orbit  operations.  This  it  did  in  a-  "loosely  organized 
and  largely  unscheduled"  way  during  the  first  months  of  1960.  Serving 
on  the  committee  were  John  M.  Eggleston,  Arthur  W.  Vogeley,  Max  C. 
Kurbjun,  and  W.  Hewitt  Phillips  of  the  Aero-Space  Mechanics  Division; 
John  A.  Dodgen  and  William  Mace  of  IRD;  and  John  Bird  and  Clint  Brown 
of  the  Theoretical  Mechanics  Division.24  The  overlapping  memberships  and 
responsibilities  of  the  committees  and  study  groups  created  during  this  busy 
and  chaotic  period  have  caused  much  confusion  in  the  historical  record  about 
where  the  concept  of  LOR  first  arose  at  Langley  and  about  who  deserves 
the  credit. 

At  one  of  the  early  meetings  of  the  Manned  Space  Laboratory  Group 
on  18  September  1959,  Houbolt  made  a  long  statement  on  the  rendezvous 
problem.  In  this  statement,  one  of  the  first  made  on  this  subject  anywhere 
inside  NASA,  Houbolt  insisted  that  his  committee  be  allowed  to  study 
rendezvous  "in  the  broadest  terms"  possible  because,  as  he  argued  correctly, 
the  technique  was  certain  to  play  a  major  role  in  almost  any  advanced 
space  mission  NASA  might  initiate.25  Three  months  later,  in  December 
1959,  Houbolt  appeared  with  other  leading  members  of  the  Manned  Space 
Laboratory  Group  before  a  meeting  of  the  Goett  Committee  held  at  Langley. 
He  urged  the  adoption  of  a  rendezvous-satellite  experiment — an  experiment, 
in  essence,  similar  to  NASA's  later  Project  Gemini — which  could  "define 
and  solve  the  problems  more  clearly."  The  Goett  Committee  members,  the 
majority  of  whom  were  still  narrowly  focusing  on  a  space  station  and  a 
circumlunar  mission,  showed  little  interest  in  Houbolt's  experiment  idea.26 

Representatives  from  Goddard,  Marshall,  and  JPL  met  at  Langley  on 
16-17  May  1960  for  an  intercenter  review  of  NASA's  current  rendezvous 
studies.  At  this  meeting,  Houbolt  gave  the  principal  Langley  presentation 
based  on  a  paper  he  had  just  delivered  at  the  National  Aeronautical  Meeting 
of  the  Society  of  Automotive  Engineers  in  New  York  City,  5-8  April.  All 
representatives  were  in  "complete  agreement"  that  rendezvous  was  "an 
important  problem  area"  that  opened  "many  operational  possibilities"  and 
that  warranted  "significant  study."  The  strength  of  Houbolt's  presentation 
demonstrated  that  of  all  the  NASA  centers,  Langley  was  "expending  the 
greatest  effort  on  rendezvous."  Eleven  studies  were  under  way  at  the 
center  compared  with  three  at  Ames  and  two  each  at  Lewis  and  the  Flight 
Research  Center.  Marshall  had  an  active  interest  in  rendezvous  but  only 
in  connection  with  advanced  Saturn  missions.  With  their  "leanings  toward 
orbital  operations,"  von  Braun's  people  had  done  little  work  specifically  on 
rendezvous  and  were  not  prepared  to  talk  about  what  little  they  had  done.2 

One  week  after  the  intercenter  review,  a  second  rendezvous  committee 
met  for  the  first  time.  It  was  part  of  a  Lunar  Mission  Steering  Group 
created  by  Director  Floyd  Thompson.  Chairing  this  group  was  hypersonics 
specialist  John  V.  Becker,  chief  of  the  Aero-Physics  Division.28  Much  larger 
and  more  formal  than  Brown's  original  little  band  of  lunar  enthusiasts,  the 

232 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

group  chaired  by  Becker  incorporated  the  Brown  group,  with  the  dynamic 
Brown  himself  serving  as  chairman  of  the  new  group's  subcommittee  on 
trajectories  and  guidance.  Five  other  subcommittees  were  quickly  organized: 
Howard  B.  Edwards  of  IRD  chaired  an  instrumentation  and  communications 
committee;  Richard  R.  Heldenfels  of  the  Structures  Research  Division 
headed  a  committee  on  structures  and  materials;  Paul  R.  Hill  of  the  Aero- 
Space  Mechanics  Division  was  in  charge  of  a  committee  on  propulsion,  flight 
testing,  and  dynamic  loads;  Eugene  S.  Love,  Becker's  assistant  chief  of  the 
Aero-Physics  Division,  led  a  committee  on  reentry  aerodynamics,  heating, 
configuration,  and  aeromedical  issues;  and  John  C.  Houbolt  headed  the 
rendezvous  committee.  Serving  with  Houbolt  were  Wilford  E.  Silvertson, 
Jr.,  of  IRD  and  John  Bird  and  John  Eggleston,  who  were  also  members  of 
his  other  rendezvous  committee  for  the  Manned  Space  Lab  Group. 

Becker's  Lunar  Mission  Steering  Group  was  to  take  a  "very  broad  look 
at  all  possible  ways  of  accomplishing  the  lunar  mission."  At  the  time  NASA 
envisioned  a  circumlunar  rather  than  a  landing  mission.  (By  late  summer 
1960,  Lowell  E.  Hasel,  secretary  of  Becker's  study  group,  was  referring  to  it 
in  his  minutes  as  the  "LRC  Circumlunar  Mission  Steering  Group.")  More 
specifically,  the  Becker  group  was  to  decide  whether  it  approved  of  the 
general  guidelines  for  lunar  missions  as  established  by  the  STG  in  meetings 
a  month  earlier,  in  April  I960.29  In  the  next  six  months,  Becker's  group  met 
six  times,  sent  representatives  to  NASA  headquarters  and  Marshall  Space 
Flight  Center  for  consultation  and  presentation  of  preliminary  analyses,  and 
generally  educated  itself  in  the  relevant  technical  areas.  Its  exploratory 
experimental  data  eventually  appeared  in  12  Langley  papers  presented  at 
the  first  NASA/Industry  Apollo  Technical  Conference  held  in  Washington 
from  18-20  July  1961.  Long  before  that  time,  however,  Langley 's  Lunar 
Mission  Steering  Group  discontinued  its  activities.  In  mid-November  1960, 
when  the  STG  developed  its  formal  Apollo  Technical  Liaison  Plan,  which 
organized  specialists  in  each  problem  area  from  every  NASA  center,  the 
group  was  no  longer  needed  and  simply  stopped  meeting.30 


Houbolt  Launches  His  First  Crusade 

In  his  paper  presented  before  the  Society  of  Automotive  Engineers  in 
April  1960,  Houbolt  had  focused  on  "the  problem  of  rendezvous  in  space, 
involving,  for  example,  the  ascent  of  a  satellite  or  space  ferry  as  to  make 
a  soft  contact  with  another  satellite  or  space  station  already  in  orbit." 
His  analysis  of  soft  rendezvous  could  have  applied  to  a  lunar  mission,  but 
Houbolt  did  not  specifically  refer  to  that  possibility.31 

He  had  been  seriously  studying  it,  as  revealed"  in  the  minutes  of  a  meeting 
of  Langley's  Manned  Space  Laboratory  Group  held  on  5  February  1960.  On 
that  occasion  Houbolt  discussed  the  general  requirements  of  a  "soft  landing 
device"  in  a  lunar  mission  involving  LOR.  He  did  so  in  spite  of  the  fact  that 

233 


Spaceflight  Revolution 


Houbolt's  Early  Crusades 


Date 

Location 

Presentation  Audience 

Sept.  1959 

Langley 

Manned  Space  Lab  Group 

Dec.  1959 

Langley 

Goett  Committee 

Feb.  1960 

Langley 

Manned  Space  Lab  Group 

Apr.  1960 

New  York 

Society  of  Automotive  Engineers 

Spring  1960 

Langley 

Robert  Piland  and  STG  members 
(informal) 

Spring  1960 

Langley 

William  Mrazek 

May  1960 

Langley 

Intercenter  Review 

Sept.  1960 

Langley 

Seamans  (informal) 

Nov.  1960 

Pentagon 

Air  Force  Scientific  Advisory 
Board 

Dec.  1960 

Langley 

STG  leaders 

Dec.  1960 

NASA  HQ 

Headquarters  staff 
including  Glennan,  von  Braun, 
Seamans,  and  Faget 

this  particular  rendezvous  committee  was  supposed  to  be  focusing  more 
narrowly  on  a  rendezvous  with  an  earth-orbiting  space  station.32 

From  this  point  on,  Houbolt  began  to  advertise  LOR  in  meetings  and 
conversations.  In  the  spring  of  1960,  he  talked  about  LOR  with  Robert  O. 
Piland  and  other  members  of  the  STG  at  Langley.  During  the  same  period, 
Houbolt  mentioned  LOR  to  William  A.  Mrazek,  director  of  the  Structures 
and  Mechanics  Division  at  Marshall.  Houbolt  had  been  helping  Mrazek  to 
evaluate  the  S-IV  stage  (consisting  of  four  uprated  Centaur  engines)  of  the 
Saturn  rocket.33 

In  the  summer  of  1960,  while  making  back-of-the-envelope  calculations  to 
confirm  the  savings  in  rocket-boosting  power  gained  by  the  LOR  approach, 
Houbolt  experienced  a  powerful  technological  epiphany.  Three  years  later, 
in  a  1963  article,  he  described  what  happened:  "Almost  simultaneously,  it 
became  clear  that  lunar-orbit  rendezvous  offered  a  chain  reaction  simplifi- 
cation on  all  'back  effects':  development,  testing,  manufacturing,  erection, 
count-down,  flight  operations,  etc."  In  this  moment  of  revelation,  Houbolt 
made  an  ardent  resolve:  "I  vowed  to  dedicate  myself  to  the  task."  From 


234 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

that  instant  until  NASA's  selection  of  the  mission  mode  for  Project  Apollo 
in  July  1962,  he  tirelessly  crusaded  for  the  LOR  concept.34 

On  1  September  1960,  Dr.  Robert  C.  Seamans,  who  had  a  Ph.D.  in 
aeronautical  engineering  and  was  a  former  member  of  an  NACA  technical 
subcommittee,  became  NASA's  new  associate  director.  One  of  his  first 
official  duties  was  visiting  all  the  agency's  field  centers  for  orientations 
about  their  programs  and  introductions  to  their  personnel.  During  his 
visit  to  Langley,  one  of  the  many  people  he  encountered  was  an  excited 
John  Houbolt,  who  seized  the  moment  to  say  something  privately  about  the 
advantages  of  LOR.  He  told  the  associate  administrator,  "We  ought  to  be 
thinking  about  using  LOR  in  our  way  of  going  to  the  moon."35 

Bob  Seamans,  in  his  previous  job  as  chief  engineer  for  RCA's  Missile 
and  Electronics  Division  in  Massachusetts,  had  been  involved  in  an  air  force 
study  known  as  Project  Saint — an  acronym  for  satellite  interceptor.  This 
"quiet  but  far-reaching"  classified  military  project  involved  the  interception 
of  satellites  in  earth  orbit.  Project  Saint  predisposed  Seamans  to  enter- 
tain ideas  about  rendezvous  techniques  and  maneuvers.  Houbolt  explained 
to  him  that  LOR  would  work  even  if  less  weight  than  that  of  the  entire 
spacecraft  was  left  in  a  parking  orbit.  If  only  the  weight  of  the  spacecraft's 
heat  shield  was  parked,  NASA  could  realize  some  significant  savings.  Im- 
pressed with  the  importance  of  leaving  weight  in  orbit,  and  equally  impressed 
with  Houbolt's  zeal,  Seamans  invited  the  impassioned  Langley  researcher  to 
present  his  ideas  formally  before  his  staff  in  Washington.36 

Before  that  presentation,  however,  Houbolt  gave  two  other  briefings 
on  rendezvous:  the  first,  in  November  1960,  to  the  Air  Force  Scientific 
Advisory  Board  at  the  Pentagon;  the  second,  on  10  December,  to  leading 
members  of  the  STG  including  Paul  Purser,  Robert  Piland,  Owen  Maynard, 
Caldwell  Johnson,  James  Chamberlin,  and  Max  Faget.  (Gilruth  was  not 
present.)  In  both  talks,  Houbolt  spoke  about  all  the  possible  uses  of 
rendezvous.  For  LOR,  the  uses  included  a  manned  lunar  landing,  and 
for  earth-orbit  rendezvous  (EOR)  they  included  assembly  of  orbital  units, 
personnel  transfer  to  a  space  station,  proper  placement  of  special  purpose 
satellites,  and  inspection  and  interception  of  satellites.  Houbolt  stressed 
that  rendezvous  would  be  both  inherently  useful  and  technically  feasible  in 
many  space  missions.  Historians  have  missed  this  key  point  about  Houbolt: 
he  was  advocating  rendezvous  generally,  not  just  LOR. 

If  humans  were  going  to  land  on  the  moon  using  existing  rocket  boosters, 
or  even  the  boosters  that  were  then  on  the  drawing  boards,  a  combination 
of  EOR  and  LOR  would  be  required.  "We  would  put  up  a  component 
with  a  first  booster;  we  would  put  up  another  component  with  another 
booster;  then  we  would  rendezvous  the  two  of  them  in  earth  orbit.  Then  we 
would  go  to  the  moon  with  this  booster  system  and  perform  the  lunar-orbit 
rendezvous  with  the  remaining  spacecraft.  The  whole  reason  for  doing  it 
this  way  would  be  because  the  boosters  were  still  too  small." 


235 


Space/light  Revolution 

Although  he  presented  several  rendezvous  concepts,  Houbolt  championed 
LOR.  With  charts  showing  a  soft  manned  lunar  landing  accomplished  both 
with  the  Saturn-class  rockets  then  in  development  and  with  existing  launch 
vehicles  such  as  Atlas  or  Langley's  Scout,  Houbolt  concluded  his  lecture 
by  emphasizing  the  "great  advantage"  of  LOR.  In  a  lunar  landing  mission, 
the  earth-boost  payload  would  be  reduced  two  to  two-and-a-half  times.  "I 
pointed  out  over  and  over  again"  that  if  these  boosters  could  be  made  bigger, 
then  NASA  "could  dispense  with  the  earth-orbit  rendezvous  portion  and  do 
it  solely  by  lunar-orbit  rendezvous."37 

Houbolt  recalls  that  neither  the  Scientific  Advisory  Board  nor  the  STG 
seemed  overly  interested;  however,  they  did  not  seem  overly  hostile.  He  was 
to  experience  this  passive  reaction  often  in  the  coming  months.  But  not  all 
the  reactions  were  so  passive.  Some  of  them,  from  intelligent  and  influential 
people  inside  the  space  program,  were  loud,  harshly  worded,  and  negative. 

On  14  December  1960,  Houbolt  traveled  to  Washington  with  a  group  of 
Langley  colleagues  to  give  the  staff  at  NASA  headquarters  the  briefing  he 
had  promised  Bob  Seamans  three  months  earlier.  All  NASA's  important 
people  were  in  the  audience,  including  Keith  Glennan,  Seamans,  Wernher 
von  Braun,  and  the  leadership  of  the  STG.  For  15  minutes,  Houbolt  moved 
carefully  through  his  charts  and  analysis.  He  concluded,  as  he  had  done 
in  the  earlier  briefings,  with  an  enthusiastic  statement  about  LOR's  weight 
savings — a  reduction  of  earth  payload  by  a  "whopping"  two  to  two-and-a- 
half  times. 

When  he  finished,  a  small  man  with  a  receding  hairline  and  a  bow  tie 
jumped  up  from  the  audience.  Houbolt  knew  all  too  well  who  he  was:  the 
hot-blooded  Max  Faget,  his  longtime  Langley  associate  and  present  member 
of  the  STG.  "His  figures  lie."  Faget  accused.  "He  doesn't  know  what  he's 
talking  about."  Even  in  a  bull  session  at  Langley,  Faget's  fiery  accusation 
would  have  been  upsetting.  But  "in  an  open  meeting,  in  front  of  Houbolt 's 
peers  and  supervisors,"  it  was  "a  brutal  thing  for  one  Langley  engineer  to 
say  to  another."38  Faget  had  not  bothered  to  voice  these  doubts  four  days 
earlier  during  the  more  private  STG  management  briefing  at  Langley,  when 
Houbolt  and  the  others  who  were  to  give  talks  at  headquarters  had  rehearsed 
their  presentations.  Faget  continued  his  vocal  objections  in  the  hallway  after 
the  headquarters  briefing  was  over.  Houbolt  tried  to  stay  calm,  but  clearly 
he  was  agitated.  He  answered  the  charge  simply  by  telling  Faget  that  he 
"ought  to  look  at  the  study  before  [making]  a  pronouncement  like  that."39  It 
was  an  "ought  to"  that  Houbolt  would  be  passing  on  to  many  LOR  skeptics 
before  it  was  all  over. 

Curiously,  earlier  at  the  same  NASA  headquarters  meeting,  Clint  Brown 
had  made  a  presentation  based  on  a  study  he  had  done  with  Ralph  W. 
Stone,  Jr.,  of  the  Theoretical  Mechanics  Division.  Brown  had  explained  a 
general  operational  concept  for  an  LOR  plan  for  a  manned  lunar  mission. 
Brown's  basic  idea  was  to  develop  an  early  launch  capability  by  combining 
several  existing  rocket  boosters,  specifically  the  Atlas,  Centaur,  and  Scout. 

236 


Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 

He  also  illustrated  the  advantage  of  rendezvous  for  weight  reduction  over 
direct  ascent.  But  oddly,  Brown's  talk — unlike  Houbolt's — did  not  provoke 
a  strong  negative  reaction.40  Perhaps  this  was  because  Houbolt  gave  a 
more  explicit  analysis  of  the  advantages  of  LOR  over  the  direct  approach, 
or  perhaps  it  was  because  Brown  had  given  his  presentation  first  and  Faget 
needed  to  build  up  some  steam,  or  perhaps  it  was  personal,  with  Faget 
simply  liking  Brown  better  than  he  liked  Houbolt. 

The  Feelings  Against  LOR 

The  basic  premise  of  the  LOR  concept  that  NASA  would  eventually 
develop  into  Project  Apollo  was  to  fire  an  assembly  of  three  spacecraft  into 
earth  orbit  on  top  of  a  single  powerful  (three-stage)  rocket,  the  Saturn  V. 
This  50,000-pound-plus  assembly  would  include  a  mother  ship  or  command 
module  (CM);  a  service  module  (SM)  containing  the  fuel  cells,  attitude 
control  system,  and  main  propulsion  system;  and  a  small  lunar  lander  or 
excursion  module.  Once  in  earth  orbit,  the  last  stage  of  the  Saturn  rocket 
would  fire  and  expend  itself,  thus  boosting  the  Apollo  spacecraft  with  its 
crew  of  three  astronauts  into  its  trajectory  to  the  moon.  Braking  into  lunar 
orbit  via  the  small  rockets  aboard  the  service  module,  two  members  would 
don  space  suits  and  climb  into  the  lunar  excursion  module  (LEM),  detach 
it  from  the  mother  ship,  and  pilot  it  to  the  lunar  surface.  The  third  crew 
member  would  remain  in  the  CM,  maintaining  a  lonely  but  busy  vigil  in 
lunar  orbit.  If  all  went  well,  the  top  half  or  ascent  stage  of  the  LEM  would 
rocket  back  up,  using  the  ascent  engine  provided,  and  redock  with  the  CM. 
What  remained  of  the  lander  would  then  be  discarded  to  the  vast  void 
of  space — or  crashed  on  the  moon  as  was  done  in  later  Apollo  missions  for 
seismic  experiments — and  the  three  astronauts  in  their  command  ship  would 
head  for  home.* 

Knowing  what  we  know  now,  that  is,  that  the  United  States  would 
land  Americans  on  the  moon  and  return  them  safely  before  the  end  of 
the  decade  using  LOR,  the  strength  of  feeling  against  the  concept  in  the 
early  1960s  is  hard  to  imagine.  In  retrospect,  we  know  that  LOR  enjoyed — 
as  Brown,  Michael,  Dolan,  and  especially  John  Houbolt  had  said — several 
advantages  over  its  competitors.  It  required  less  fuel,  only  half  the  payload, 
and  somewhat  less  new  technology;  it  did  not  require  a  monstrous  rocket 
such  as  the  proposed  Nova  for  a  direct  flight;  and  it  called  for  only  one  launch 
from  earth,  whereas  one  of  LOR's  chief  competitors,  EOR,  required  at  least 
two.  Only  the  small  lightweight  LEM,  not  the  entire  spacecraft,  would  have 
to  land  on  the  moon.  This  was  perhaps  LOR's  major  advantage.  Because 


One  can  summarize  the  LOR  concept  with  three  specifications:  (1)  Only  a  specially  designed  LEM 
would  actually  descend  to  the  moon's  surface;  (2)  Only  a  portion  of  that  LEM,  the  ascent  stage,  would 
return  to  dock  with  the  CM  in  lunar  orbit;  and  (3)  Only  the  CM  or  Apollo  capsule  itself,  with  its 
protective  heat  shield,  would  fall  back  to  earth. 

237 


Space/light  Revolution 

the  lander  was  to  be  discarded  after  use  and  would  not  be  needed  to  return 
to  earth,  NASA  could  customize  the  design  of  the  LEM  for  maneuvering 
flight  in  the  lunar  environment  and  for  a  soft  lunar  landing.  In  fact,  the 
beauty  of  LOR  was  that  NASA  could  tailor  all  the  modules  of  the  Apollo 
spacecraft  independently — without  those  tailorings  having  to  compromise 
each  other.  One  spacecraft  unit  to  do  three  jobs  would  have  forced  some 
major  concessions,  but  three  units  to  do  three  jobs  was  another  plus  for  LOR 
that  no  one  at  NASA,  finally,  could  overlook. 

In  the  early  1960s,  these  advantages  were  theoretical,  but  the  fear  that 
American  astronauts  might  be  left  in  an  orbiting  coffin  some  240,000  miles 
from  home  was  quite  real.  If  rendezvous  had  to  be  part  of  the  lunar  mission, 
many  people  felt  it  should  be  attempted  only  in  earth  orbit.  If  rendezvous 
failed  there,  the  threatened  astronauts  could  be  brought  home  simply  by 
allowing  the  orbit  of  their  spacecraft  to  deteriorate.  If  a  rendezvous  around 
the  moon  failed,  the  astronauts  would  be  too  far  away  to  be  saved.  Nothing 
could  be  done.  The  specter  of  dead  astronauts  sailing  around  the  moon 
haunted  those  responsible  for  the  Apollo  program.  This  anxiety  made 
objective  evaluation  of  LOR  by  NASA  unusually  difficult. 

John  Houbolt  understood  NASA's  fears,  but  he  recognized  that  all  the 
alternative  schemes  had  serious  pitfalls  and  dreadful  possibilities  of  their 
own.  He  was  certain  that  all  the  other  options  would  be  more  perilous  and 
did  not  really  offer  rescue  possibilities.  The  LOR  concept,  in  contrast,  did 
offer  the  chance  of  a  rescue  if  two  small  landing  modules,  rather  than  one, 
were  included.  One  lander  could  be  held  in  reserve  with  the  orbiting  mother 
ship  to  go  down  to  the  lunar  surface  if  the  number  one  lander  encountered 
serious  trouble.  Or,  in  the  case  of  an  accident  inside  the  command-service 
module,  one  attached  LEM  could  serve  as  a  type  of  "lifeboat."*  Houbolt 
just  could  not  accept  the  charge  that  LOR  was  inherently  more  dangerous, 
but  neither  could  he  easily  turn  that  charge  aside. 

The  intellectual  and  emotional  climate  in  which  NASA  would  have  to 
make  perhaps  the  most  fundamental  decision  in  its  history  was  amazingly 
tempestuous.  The  psychological  obstacle  to  LOR's  progress  made  the 
entire  year  of  1961  and  the  first  seven  months  of  1962  the  most  hectic  and 
challenging  period  of  John  Houbolt's  life.41 

On  5  January  1961,  Houbolt  spoke  again  on  rendezvous  during  the 
first  afternoon  session  of  a  historic  two-day  Space  Exploration  Program 
Council  (SEPC)  in  Washington.  This  council  had  been  created  by  NASA 
for  "smoothing  out  technical  and  managerial  problems  at  the  highest  level." 
Chaired  by  the  associate  administrator,  this  council  meeting — the  first  that 


This  scenario  would  indeed  happen  during  the  mission  of  Apollo  13,  when  outward  bound  and 
200,000  miles  from  earth,  an  explosion  in  one  of  the  oxygen  tanks  within  the  service  module  caused  a 
leak  in  another  oxygen  tank  and  confronted  NASA  with  an  urgent  life-threatening  problem.  NASA  solved 
the  problem  by  having  the  astronauts  head  home,  without  landing,  and  by  moving  them  temporarily 
into  the  atmosphere  of  the  LEM. 

238 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

Seamans  presided  over — included,  as  always,  all  program  office  heads  at 
headquarters,  the  directors  of  all  NASA  field  centers,  and  their  invited  guests 
and  speakers.  The  SEPC  had  been  meeting  quarterly  since  early  1960,  but 
this  first  meeting  of  1961  was  by  far  the  most  historic  to  date;  it  was  the 
first  meeting  inside  NASA  to  feature  an  agencywide  discussion  of  a  manned 
lunar  landing.42 

At  the  end  of  the  first  day  of  this  meeting,  everyone  agreed  that  the 
mission  mode  for  a  manned  lunar  landing  by  NASA  could  be  reduced  to 
three  options:  direct  ascent,  which  was  still  the  front-runner;  EOR,  which 
was  gaining  ground  quickly;  and  LOR,  the  dark  horse  on  which  only  the 
most  capricious  gamblers  in  NASA  would  have  ventured  a  bet. 

One  speaker  had  presented  each  option.  First,  Marshall's  impressive 
rocket  pioneer  from  Germany,  Wernher  von  Braun,  reviewed  NASA's  launch 
vehicle  program,  with  an  eye  to  the  advantages  of  EOR.  Von  Braun  ex- 
plained how  two  pieces  of  hardware  could  be  launched  into  space  indepen- 
dently using  advanced  Saturn  rockets  then  under  development;  how  the  two 
pieces  could  rendezvous  and  dock  in  earth  orbit;  how  a  lunar  mission  vehi- 
cle could  be  assembled,  fueled,  and  detached  from  the  joined  modules;  and 
how  that  augmented  ship  could  proceed  directly  to  the  surface  of  the  moon 
and  after  exploring,  return  to  earth.  The  clearest  immediate  advantage  of 
EOR,  as  von  Braun  pointed  out,  was  that  it  required  a  pair  of  less  powerful 
rockets  that  were  already  nearing  the  end  of  their  development.  Two  of  his 
early  Saturns  would  do  the  job.  The  biggest  pitfall  of  EOR,  as  with  direct 
ascent,  was  that  no  one  knew  how  the  spacecraft  would  actually  make  its 
landing.  On  the  details  of  that  essential  maneuver,  von  Braun  said  nothing 
other  than  to  admit  that  more  serious  study  would  have  to  be  done  very 
quickly.43 

Next,  Melvyn  Savage  of  the  Office  of  Launch  Vehicle  Programs  at  NASA 
headquarters  explained  direct  ascent.  A  massive  rocket  roughly  the  size  of  a 
battleship  would  be  fired  directly  to  the  moon,  land,  and  blast  off  for  home 
directly  from  the  lunar  surface.  The  trip  required  one  brute  of  a  booster 
vehicle,  the  proposed  12-million-pound  thrust  Nova  rocket. 

Late  in  the  afternoon,  Houbolt  came  to  the  podium  to  discuss  rendezvous 
and  highlight  the  unappreciated  strengths  of  his  dark-horse  candidate.  To 
him  the  advantages  of  LOR  and  the  disadvantages  of  the  other  two  options 
were  obvious.  Any  single  rocket  such  as  Nova  that  had  to  carry  and  lift 
all  the  fuel  necessary  for  leaving  the  earth's  gravity,  braking  against  the 
moon's  gravity  as  well  as  leaving  it,  and  braking  against  the  earth's  gravity 
was  clearly  not  the  most  practical  choice — especially  if  the  mission  was  to 
be  accomplished  in  the  near  future.  The  development  of  a  rocket  that 
mammoth  would  take  too  long,  and  the  expense  would  be  enormous.  In 
Houbolt 's  opinion,  EOR  was  a  more  reasonable  choice  than  direct  ascent  but 
not  as  sensible  as  LOR.  After  the  lunar-bound  spacecraft  left  its  rendezvous 
station  around  the  earth,  the  rest  of  an  EOR  mission  would  be  accomplished 
in  exactly  the  same  way  as  direct  ascent.  NASA's  crew  of  astronauts  would 

239 


Spaceflight  Revolution 


Houbolt's  Later  Crusades 


Date 

Location 

Presentation  Audience 

Jan.  1961 

NASA  HQ 

Space  Exploration  Program 
Council  Meeting 

Jan.  1961 

Langley 

STG  members 

Jan.  1961 

Langley 

Pearson  (Low  Committee) 

Jan.  1961 

NASA  HQ 

NASA  Headquarters  staff 

Apr.  1961 

Langley 

STG 

May  1961 

Letter  to  Seamans  (NASA  HQ) 

June  1961 

NASA  HQ 

Lundin  Committee 

June  1961 

France 

International  Space  Flight 
Symposium 

July  1961 

Langley 

Rehearsal  with  STG 

July  1961 

Washington,  B.C. 

NASA/Industry  Apollo  Technical 
Conference 

Aug.  1961 

NASA  HQ 

Golovin  Committee 

Aug.  1961 

Langley 

STG 

Nov.  1961 

Letter  to  Seamans  (NASA  HQ) 

Jan.  1962 

Langley 

Shea  and  STG 

Jan.  1962 

MSC,  Houston 

Manned  Spacecraft  Center 
personnel 

Feb.  1962 

NASA  HQ 

Manned  Space  Flight  Management 
Council 

Apr.  1962 

Report  and  papers  sent  to 
von  Braun  (MSFC) 

have  to  land  an  incredibly  heavy  and  large  vehicle  on  the  surface  of  the 
moon.  The  business  of  backing  such  a  large  stack  of  machinery  down  onto 
the  moon  and  "eyeballing"  a  pinpoint  soft  landing  on  what  at  the  time 
was  still  a  virtually  unknown  lunar  surface  would  be  incredibly  tricky  and 
dangerous.  Those  few  NASA  researchers  who  had  been  thinking  about  the 
terrors  of  landing  such  a  behemoth  (and  getting  the  astronauts  down  from 


240 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

the  top  of  it  using  some  little  inside  elevator!)  knew  that  no  satisfactory 
answers  to  these  problems  were  on  the  horizon.44 

Other  talks  were  given  that  day,  including  an  introduction  by  George 
Low,  chair  of  NASA  headquarters  Manned  Lunar  Landing  Task  Group 
(formed  in  October  1960),  and  a  technical  talk  by  Houbolt's  nemesis 
Max  Faget  outlining  the  hardware  and  booster  requirements  for  several 
possible  types  of  lunar  missions.  Everyone  walked  away  from  the  meeting 
understanding  that,  if  the  United  States  was  someday  to  reach  the  moon, 
NASA  would  have  to  choose  a  plan  soon.45  At  this  point,  the  odds  were 
excellent  that  the  choice  would  be  direct  ascent,  which  seemed  simplest  in 
concept.  Coming  in  second,  if  a  vote  had  been  taken  that  day,  would  have 
been  EOR.  LOR,  to  many  NASA  officials  present,  was  an  option  almost 
unworthy  of  mention. 

The  Early  Skepticism  of  the  STG 

In  the  early  months  of  1961,  the  STG  was  preoccupied  with  the  first 
manned  Mercury  flight  and  the  hope — soon  to  be  crushed  by  Vostok  1— 
that  an  American  astronaut  would  be  the  first  human  in  space.  When 
any  of  its  members  had  a  rare  moment  to  consider  rendezvous,  they  were 
typically  thinking  about  it  "as  one  of  several  classes  of  missions  around 
which  a  Mercury  program  follow-on  might  be  built."46 

At  Langley  on  10  January  1961,  five  days  after  the  meeting  of  the  SEPC, 
Houbolt  went  with  three  members  of  the  Theoretical  Mechanics  Division — 
the  division  chief  Clint  Brown,  Ralph  Stone,  and  Manuel  J.  "Jack"  Queijo — 
to  an  informal  meeting  at  the  center  with  three  members  of  the  STG's 
Flight  Systems  Division — H.  Kurt  Strass,  Owen  E.  Maynard,  and  Robert 
L.  O'Neal.  Langley  Associate  Director  Charles  Donlan,  Gilruth's  former 
chief  assistant,  also  attended.  At  this  meeting  Houbolt,  Brown,  and  the 
others  tried  to  persuade  representatives  from  the  STG  that  a  rendezvous 
experiment  belonged  in  the  Apollo  program  and  that  LOR  was  the  way  to 
go  if  any  plans  for  a  manned  lunar  landing  were  to  be  made.47 

They  were  not  persuaded.  Although  the  STG  engineers  received  the 
analysis  more  politely  than  had  cohort  Max  Faget  the  month  earlier, 
all  four  men  admitted  quite  frankly  that  the  claims  about  the  weight 
savings  were  "too  optimistic."  Owen  Maynard  remembers  that  he  and 
his  colleagues  initially  viewed  the  LOR  concept  as  "the  product  of  pure 
theorists'  deliberations  with  little  practicality."  In  essence,  they  agreed  with 
Faget's  charge,  though  they  did  not  come  out  and  say  it,  that  Houbolt's 
figures  did  "lie."  The  STG  engineers  believed  that  in  advertising  the  earth- 
weight  savings  of  LOR  and  the  reduction  in  the  size  of  the  booster  needed 
for  the  lunar  mission,  Houbolt  and  the  others  were  failing  to  factor  in,  or 
they  were  at  least  greatly  underestimating,  the  significant  extra  complexity, 
and  thus  added  weight,  of  the  systems  and  subsystems  that  LOR's  modular 
spacecraft  would  require.48 

241 


Spaceflight  Revolution 

This  criticism  was  central  to  the  early  skepticism  toward  the  LOR  concept 
both  inside  and  outside  the  STG.  Even  Marshall's  Wernher  von  Braun 
initially  shared  the  sentiment:  "John  Houbolt  argued  that  if  you  could  leave 
part  of  your  ship  in  orbit  and  don't  soft  land  all  of  it  on  the  moon  and  fly  it 
out  of  the  gravitational  field  of  the  moon  again,  you  can  save  takeoff  weight 
on  earth."  "That's  pretty  basic,"  von  Braun  recalled  later  in  an  oral  history. 
"But  if  the  price  you  pay  for  that  capability  means  that  you  have  to  have  one 
extra  crew  compartment,  pressurized,  and  two  additional  guidance  systems, 
and  the  electrical  supply  for  all  that  gear,  and  you  add  up  all  this,  will  you 
still  be  on  the  plus  side  of  your  trade-off?"  Until  the  analysis  was  done  (and 
some  former  NASA  engineers  still  argue  today  that  "this  trade-off  has  never 
been  realistically  evaluated"),  no  one  could  be  sure.  Many  NASA  people 
suspected  that  LOR  would  prove  far  too  complicated.  "The  critics  in  the 
early  debate  murdered  Houbolt,"  von  Braun  remembered  sympathetically.49 

Houbolt  recalls  this  January  1961  meeting  with  the  men  from  the  STG  as 
a  "friendly,  scientific  discussion."  He,  Brown,  and  the  others  did  what  they 
could  to  counter  the  argument  that  the  weight  of  a  modular  spacecraft  would 
prove  excessive.  Using  an  argument  taken  from  automobile  marketing, 
they  stated  that  the  lunar  spacecraft  would  not  necessarily  have  to  be 
"plush" ;  an  "economy"  or  even  "budget"  model  might  be  able  to  do  the  job. 
Houbolt  offered  as  an  example  one  of  John  Bird's  lunar  bugs,  "a  stripped- 
down,  2,500-pound  version  in  which  an  astronaut  descended  on  an  open 
platform,"50  but  the  STG  engineers  did  not  take  the  budget  model  idea 
seriously.  In  answer  to  the  charge  that  a  complicated  modular  spacecraft 
would  inevitably  grow  much  heavier  than  the  LOR  advocates  had  been 
estimating,  Houbolt  retaliated  with  the  argument  that  the  estimated  weight 
of  a  direct-ascent  spacecraft  would  no  doubt  increase  during  development, 
making  it  an  even  less  competitive  option  in  comparison  with  rendezvous. 

But  in  the  end,  the  substantive  differences  between  the  two  groups 
of  engineers  went  out  the  window.  All  Houbolt  could  say  to  the  STG 
representatives  was  "you  don't  know  what  you're  talking  about,"  and  all 
they  could  say  to  him  was  the  same.  "It  wasn't  a  fight  in  the  violent  sense," 
reassures  Houbolt.  "It  was  just  differences  in  scientific  opinion  about  it."51 

Whether  the  skeptical  response  to  that  day's  arguments  in  favor  of  LOR 
was  indicative  of  general  STG  sentiment  in  early  1961  has  been  a  matter 
of  some  serious  behind-the-scenes  debate  among  the  NASA  participants. 
Houbolt  has  argued  that  the  STG  consistently  opposed  LOR  and  had  to  be 
convinced  from  the  outside,  by  Houbolt  himself,  after  repeated  urgings,  that 
it  was  the  best  mission  mode  for  a  lunar  landing.  Leading  members  of  the 
STG,  notably  Gilruth  and  Faget,  have  argued  that  was  not  really  the  case. 
They  say  that  the  STG  was  too  busy  preparing  for  the  Mercury  flights  to 
think  seriously  about  lunar  studies;  they  began  considering  such  missions 
only  after  Kennedy's  commitment.  Gilruth  recalls  that  when  Houbolt  first 
approached  him  "with  some  ideas  about  rendezvousing  Mercury  capsules 
in  earth  orbit"  as  "an  exercise  in  space  technology,"  he  did  in  fact  react 

242 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 


Looking  like  a  birdie  for  a  badminton  game, 
this  early  lunar  excursion  model  was  pro- 
posed by  Langley  researchers  in  the  spring 
of  1961  for  the  suggested  Project  MALLIR 
(Manned  Lunar  Landing  Involving  Ren- 
dezvous). 
L-61-6790 


negatively.  It  was  a  "diversion  from  our  specified  mission"  according  to 
Gilruth  and,  therefore,  not  something  about  which  he  as  the  head  of  Project 
Mercury  then  had  any  time  on  which  to  reflect.52 

According  to  Gilruth,  he  did  not  know  of  Houbolt's  interest  in  LOR  until 
early  1961.  By  that  time,  NASA  had  begun  studying  the  requirements  of 
a  manned  lunar  landing  through  such  committees  as  the  Manned  Lunar 
Landing  Task  Group  chaired  by  George  Low  (the  Low  Committee).  The 
STG,  although  still  overwhelmed  with  work,  did  its  best  to  follow  suit.  When 
it  did  begin  serious  consideration  of  a  lunar  program,  especially  of  landing 
men  on  the  moon,  LOR  gained  "early  acceptance  . . .  notwithstanding  the 
subsequent  debates  that  erupted  in  numerous  headquarters  committees."53 

"I  was  very  much  in  favor  of  that  mode  of  flight  to  the  moon  from  the  very 
beginning,"  Gilruth  has  since  claimed.  "I  recall  telling  our  people  that  LOR 
seemed  the  most  promising  mode  to  me — far  more  promising  than  either  the 
direct  ascent  or  the  earth  orbital  rendezvous  modes."  The  most  important 
thing  in  planning  for  a  manned  lunar  program  was  to  minimize  the  risk  of 
the  landing  operation.  Thus,  LOR  was  the  best  of  the  contending  modes 
because  it  alone  permitted  the  use  of  a  smaller  vehicle  specifically  designed 
for  the  job.  In  Gilruth's  view,  he  was  always  encouraging  to  Houbolt.  In 
his  estimation,  he  felt  all  along  that  "the  Space  Task  Group  would  be  the 
key  in  carrying  the  decision  through  to  the  highest  echelons  of  NASA"  and 
that,  "of  course,  this  proved  to  be  the  case."5 

243 


Spaceflight  Revolution 


Although  Houbolt  was  not  the  first 
to  foresee  the  advantages  of  a  moon 
landing  via  LOR,  his  total  commit- 
ment and  crusading  zeal  won  the 
support  of  key  people  in  NASA. 


L-62-5208 


Houbolt  accepts  very  little  of  these  ex  post  facto  assertions;  indeed,  he 
violently  disagrees  with  them.  He  points  out  that  on  several  occasions  in 
late  1960  he  had  briefed  leading  members  of  the  STG  about  LOR  and  that 
Gilruth  had  to  know  his  ideas.  According  to  Houbolt,  the  STG  had  ignored 
and  resisted  his  calculations  as  too  optimistic  and  continued  to  ignore  and 
resist  them  while  insisting  on  the  development  of  the  large  Nova  boosters. 
As  evidence,  Houbolt  points  to  many  subsequent  incidents  in  which  his 
ideas  were  summarily  discounted  by  the  STG  and  to  various  statements  of 
resistance  from  key  STG  members.  One  such  statement  came  from  Gilruth 
in  an  official  letter  as  late  as  September  1961.  "Rendezvous  schemes  are 
and  have  been  of  interest  to  the  Space  Task  Group  and  are  being  studied," 
Gilruth  informed  NASA  headquarters  on  12  September.  "However,  the 
rendezvous  approach  itself  will,  to  some  extent,  degrade  mission  reliability 
and  flight  safety."  Rendezvous  schemes  such  as  Houbolt's  "may  be  used 
as  a  crutch  to  achieve  early  planned  dates  for  launch  vehicle  availability," 
Gilruth  warned.  Their  advocates  propose  them  "to  avoid  the  difficulty  of 
developing  a  reliable  Nova  class  launch  vehicle."55 

Houbolt  felt  strongly  that  if  he  could  just  persuade  Gilruth's  people  to  do 
their  homework  on  rendezvous,  "then  they  too  would  become  convinced  of  its 
merits."  But  for  months  he  could  not  get  them,  or  anyone  else,  to  do  that. 
LOR  met  with  "virtually  universal  opposition — no  one  would  accept  it — 
they  would  not  even  study  it."  In  Houbolt's  words,  "my  perseverance,  and 
solely  mine"  caused  the  STG  and  various  other  groups  finally  to  study  and 


244 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

realize  "the  far-sweeping  merits  of  the  plan."  "My  own  in-depth  analysis, 
. . .  my  crusading,  . . .  paved  the  way  to  the  acceptance  of  the  scheme."56 

In  early  1961,  when  the  Low  Committee  announced  its  plan  for  a  manned 
lunar  landing  and  its  aspiration  for  that  bold  mission  to  be  made  part  of 
Project  Apollo,  NASA  still  appeared  to  be  resisting  LOR.  In  outlining  the 
requirements  for  a  manned  lunar  flight,  the  committee's  chief  recommen- 
dation was  to  focus  on  the  direct  approach  to  the  moon,  thus  leaving  ren- 
dezvous out;  LOR  was  not  discussed  at  all.  Low  remembers  that  during  the 
time  of  his  committee's  deliberations,  he  asked  one  of  the  committee  mem- 
bers, E.  O.  Pearson,  Jr.,  to  visit  John  Houbolt  at  Langley  and  "to  advise 
the  Committee  whether  we  should  give  consideration  to  the  Lunar  Orbit 
Rendezvous  Mode."  Pearson,  the  assistant  chief  of  the  Aerodynamics  and 
Flight  Mechanics  Research  Division  at  NASA  headquarters,  returned  with 
the  answer,  "No,"  LOR  "was  not  the  proper  one  to  consider  for  a  lunar  land- 
ing." A  rendezvous  240,000  miles  from  home,  when  rendezvous  had  never 
been  demonstrated — Shepard's  suborbital  flight  had  not  even  been  made 
yet — seemed,  literally  and  figuratively,  "like  an  extremely  far-out  thing  to 
do."57 

Thus  the  Low  Committee  in  early  1961,  recognizing  that  it  would  be 
much  too  expensive  to  develop  and  implement  more  than  one  lunar  landing 
mission  mode,  made  its  "chief  recommendation":  NASA  should  focus 
on  direct  ascent.  "This  mistaken  technical  judgment  was  not  Houbolt 's 
fault,"  Low  admitted  years  later,  "but  rather  my  fault  in  trusting  a 
single  Committee  member  instead  of  having  the  entire  Committee  review 
Houbolt's  studies  and  recommendations."58 


Mounting  Frustration 

Everything  that  happened  in  the  first  months  of  early  1961  reinforced 
John  Houbolt's  belief  that  NASA  was  dismissing  LOR  without  giving  it 
due  consideration.  On  20  January,  Houbolt  gave  another  long  talk  at 
NASA  headquarters  on  rendezvous.  In  this  briefing,  he  displayed  analysis 
showing  a  manned  lunar  landing  using  Saturn  rockets  and  outlined  a 
simplified  rendezvous  scheme  that  had  been  devised  by  Art  Vogeley  and 
Lindsay  J.  Lina  of  the  Guidance  and  Control  Branch  of  Langley 's  Aero- 
Space  Mechanics  Division.  He  also  mentioned  preliminary  ideas  for  the 
development  of  fixed-base  simulators  by  which  to  study  the  requirements 
for  manned  lunar  orbit,  landing,  and  rendezvous.59  On  27  and  28  February, 
NASA  held  an  intercenter  meeting  on  rendezvous  in  Washington,  but 
no  LOR  presentation  was  made  by  Houbolt  or  anyone  else.  As  if  by 
political  consensus,  the  subject  was  not  even  brought  up.  This  prompted 
one  concerned  headquarters  official,  Bernard  Maggin  from  the  Office  of 
Aeronautical  and  Space  Research,  to  write  Houbolt  a  memo  commenting 
on  NASA's,  and  especially  the  STG's,  lack  of  consideration  for  LOR.60 

245 


Spaceflight  Revolution 

Institutional  politics  was  involved  in  the  unfolding  lunar  landing  mission 
mode  debate.  The  politics  centered  around  the  concern  over  where  the  work 
for  the  manned  lunar  program  was  going  to  be  done.  The  organizations 
involved  in  building  the  big  rockets  were  interested  in  direct  ascent,  which 
required  the  giant  Nova,  and  in  EOR,  which  required  two  or  more  Saturns 
per  mission.  Abe  Silverstein,  the  director  of  the  Office  of  Space  Flight 
Programs  at  NASA  headquarters,  was  working  primarily  from  his  experience 
as  the  former  head  of  Lewis  Research  Center,  which  was  the  old  NACA 
propulsion  research  laboratory  now  heavily  involved  in  rocket  development, 
so  he  naturally  favored  direct  ascent.  Wernher  von  Braun  had  to  be  thinking 
about  the  best  interests  of  his  Marshall  Space  Flight  Center,  which  was 
primarily  responsible  at  that  time  for  developing  the  Saturns.61  For  the 
most  part,  the  management  staff  of  Langley  kept  out  of  these  debates.  No 
matter  which  mission  mode  was  implemented,  Langley  researchers  and  wind 
tunnels  would  have  plenty  of  work  to  do  to  support  the  program.62 

In  some  articles  and  history  books  on  Project  Apollo,  LOR  has  been 
called  a  pet  concept  of  Langley,  but  that  was  not  the  case.  Even  within 
Langley,  LOR  was  embraced  only  by  a  small  but  vocal  minority.  Langley 
management  did  not  get  behind  LOR  until  after  the  STG  and  the  rest  of 
NASA  did.  The  personal  opinion  of  Langley  Director  Floyd  Thompson,  as 
well  as  that  of  most  of  his  senior  staff,  mirrored  that  of  the  STG:  LOR  was 
too  complicated  and  risky.  Direct  ascent  or  EOR  was  the  better  choice.63 

Although  a  brilliant  engineering  analyst  and  an  energetic  advocate  of  the 
causes  he  espoused,  Houbolt  was  not  an  overly  shrewd  behind-the-scenes 
player  of  institutional  politics.  Faced  with  the  impasse  of  early  1961,  his 
first  instinct  was  simply  to  find  more  informed  retorts  to  the  criticisms  he 
had  been  hearing.  So,  with  the  help  of  Brown,  Vogeley,  Michael,  Bird, 
Kurbjun,  and  a  few  others,  he  developed  elaborate  and  detailed  studies 
of  the  lunar  landing  mission  he  envisioned  along  with  extensive  analyses 
of  weight  savings.  Somehow,  he  felt,  he  must  find  a  way  to  circumvent 
the  problem  and  convince  the  agency  that  it  was  making  a  big  mistake  by 
dismissing  LOR. 

On  19  April  1961,  Houbolt  was  to  give  another  briefing  on  rendezvous 
to  the  STG.  In  an  effort  to  package  his  argument  more  convincingly,  he 
created  an  "admiral's  page."  This  was  a  short,  visually  convenient  summary 
for  "the  admiral"  designed  to  save  him  wading  through  a  long  report.  For 
his  STG  briefing,  Houbolt  put  16  pages  of  charts,  data  plots,  drawings, 
and  outlined  analyses — taken  from  his  own  analysis  as  well  as  material 
supplied  by  Langley's  John  Bird,  Max  Kurbjun,  and  Art  Vogeley — onto 
one  17  x  22-inch  foldout  sheet.  The  title  of  his  foldout  was  "Manned  Lunar 
Landing  Via  Rendezvous"  and  on  its  cover  was  a  telescopic  photograph  of 
the  moon.  Several  important  people  attending  the  meeting  received  a  copy 
of  the  foldout  which  helped  them  follow  Houbolt 's  talk  more  closely.64 

As  had  been  the  case  in  Houbolt 's  earlier  presentations,  this  one  also 
dealt  with  both  EOR  and  LOR,  but  it  had  a  clearly  stated  preference 

246 


Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 


L-62-5849 

Houbolt  explains  the  critical  weight-saving  advantage  of  the  LOR  scheme.  Because 
the  lunar  excursion  vehicle  ("L.E.V.")  in  Houbolt's  plan  weighed  only  19,320 
pounds,  compared  to  82, 700  pounds  for  the  lander  required  for  direct  ascent  or  EOR, 
the  total  weight  that  must  be  boosted  to  earth  escape  could  be  reduced  by  more  than 
half  using  LOR. 


for  LOR.  In  this  talk,  however,  Houbolt  advocated  for  the  first  time  two 
specific  projects  for  which  he  supplied  project  names  and  acronyms. 
The  first  of  these  ("Project  1")  he  called  "MORAD" —Manned  Orbital 
Rendezvous  and  Docking.  This  was  his  old  idea  for  a  modest  flight  "exper- 
iment" follow-on  to  Mercury  that  would  "establish  confidence"  in  manned 
rendezvous  techniques.  An  unmanned  pay  load  from  a  Scout  rocket  would 
serve  as  a  target  vehicle  for  a  maneuvering  Mercury  capsule  in  earth  orbit. 
The  second  of  the  projects  ("Project  2")  he  called  "MALLIR" —Manned 
Lunar  Landing  Involving  Rendezvous.  This  contained  the  essence  of  the 
controversial  LOR  scheme.65 


247 


Space/light  Revolution 

In  the  last  box  of  his  foldout,  Houbolt  listed  his  recommendations  for 
"Immediate  Action  Required."  For  MORAD,  he  wanted  NASA  to  give 
a  quick  go-ahead  so  that  Langley  could  proceed  with  a  work  statement 
preparatory  to  contracting  with  industry  to  do  a  study.  For  MALLIR,  he 
wanted  NASA  "to  delegate  responsibility  to  the  Space  Task  Group"  so  that 
the  STG  would  have  to  give  "specific  and  accelerated  consideration"  to 
the  possibility  of  including  rendezvous  as  part  of  Project  Apollo.  In  place 
of  the  STG's  apparent  resistance  to  his  rendezvous  ideas  and  its  current 
discretionary  freedom  to  treat  the  matter  of  rendezvous  as  part  of  Apollo  on 
a  will-also-consider  basis,  he  wanted  a  NASA  directive  that  made  rendezvous 
integral  to  an  accepted  project.  Houbolt  wanted  something  that  would  make 
the  STG,  finally,  give  rendezvous  the  attention  it  merited.  "I  simply  wanted 
people  to  study  the  problems  and  look  at  [them],  and  then  make  a  judgment, 
but  they  wouldn't  even  do  that,"  Houbolt  remembers  with  some  of  his  old 
frustration.  "It  was  that  strange  a  position."6 

Nothing  came  immediately  from  either  one  of  his  proposals.  Again,  the 
reaction  seemed  to  him  to  be  mostly  negative,  as  if  the  STG  still  wanted 
no  part  of  his  ideas.  His  frustration  mounted.  "I  could  never  find  a  real 
answer  to  why  they  wouldn't  even  consider  it,"  Houbolt  laments.  Perhaps 
it  was  the  not-invented-here  syndrome,  perhaps  it  was  just  because  he  was  an 
"outsider"  who  was  "rocking  the  boat  on  their  own  thinking,  and  they  didn't 
want  anybody  to  do  that,""7  or  perhaps  the  STG  was  just  not  prepared  to 
think  seriously  about  such  an  incredibly  bold  and  seemingly  treacherous  idea 
when  they  were  still  not  even  sure  that  they  would  be  able  to  make  their  own 
Mercury  program  a  complete  success.  Mercury  "was  proving  so  troublesome 
that  rendezvous,  however  simple  in  theory,  seemed  very  far  away." 

At  this  April  1961  briefing,  however,  a  solitary  STG  engineer  did  demon- 
strate a  clear  and  exceptional  interest  in  Houbolt's  rendezvous  analysis. 
James  Chamberlin  approached  Houbolt  after  the  meeting  and  asked  him  for 
an  extra  copy  of  the  foldout  sheet  and  "for  anything  else  he  had  on  ren- 
dezvous." Interestingly,  both  Houbolt  and  Chamberlin  recall  Chamberlin 
telling  him  that  he  had  known  about  Langley's  rendezvous  work  but  this 
was  the  first  time  he  had  heard  any  of  the  details  about  the  lunar-orbit 
version.69  One  might  indeed  wonder  then  how  widely  the  information  from 
Houbolt's  previous  talks  had  spread  within  the  STG.  It  seems  significant 
that  Chamberlin  was  not  one  of  Gilruth's  old-time  associates  from  the 
NACA;  he  was  one  of  the  relative  newcomers.* 

President  Kennedy's  Commitment 

Houbolt's  April  briefing  to  the  STG  came  at  the  end  of  a  humbling 
week  for  America.  On  12  April  the  Soviets  sent  the  first  human  into  space, 


The  former  chief  of  design  for  the  Avro  Arrow  aircraft,  Chamberlin  had  been  recruited  by  the  STG 
in  late  1959. 

248 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

cosmonaut  Yuri  Gagarin,  beating  the  United  States  in  the  second  leg  of  the 
space  race.  Three  days  later,  with  President  Kennedy's  hesitant  approval,  a 
confused  invasion  force  prepared  by  the  CIA  landed  at  Cuba's  Bay  of  Pigs 
only  to  be  driven  back  quickly  by  an  unexpectedly  efficient  army  of  20,000 
led  by  communist  Fidel  Castro.  Pierre  Salinger,  Kennedy's  articulate  press 
secretary,  later  called  this  period  "the  three  grimmest  days"  of  the  Kennedy 
presidency.  This  national  crisis  proved  in  some  ways  to  be  more  urgent  than 
even  the  troubled  aftermath  of  Sputnik.70 

Up  to  this  time,  NASA  had  been  preparing  for  a  lunar  landing  mission 
as  its  long-term  space  goal.  Some  NASA  visionaries,  such  as  George  Low, 
wanted  to  go  to  the  moon  sooner  rather  than  later  and  were  working  to 
convince  NASA  leadership,  now  headed  by  a  new  administrator,  James  E. 
Webb,  that  such  a  program  should  be  pushed  with  the  politicians.  Not  all 
the  politicians  needed  to  be  pushed.  Most  notably,  Vice-President  Lyndon 
B.  Johnson  was  pressing  NASA  for  a  more  ambitious  space  agenda  that 
included  a  lunar  landing  program.71  President  Kennedy,  however,  needed  to 
be  convinced.  The  one-two  punch  of  the  Gagarin  flight  and  Bay  of  Pigs  fiasco 
followed  by  the  welcome  relief  and  excitement  of  Alan  Shepard's  successful 
Mercury  flight  on  5  May  was  enough  to  persuade  the  president.  Sputnik  1 
and  2  had  taken  place  in  the  previous  Republican  administration  and  had 
helped  the  dynamic  young  senator  from  Massachusetts  nose  by  Eisenhower's 
vice-president,  Richard  M.  Nixon,  in  the  1960  election.  Now,  in  just  a 
month,  Kennedy's  "New  Frontier"  had  itself  been  undermined  by  crisis. 
Something  had  to  be  done  to  provoke  the  country  into  rebounding  from  its 
recent  second-place  finishes  and  national  humiliations.72  On  25  May,  John 
Kennedy  announced  that  American  astronauts  would  be  first  to  land  on  the 
moon. 


Houbolt's  First  Letter  to  Seamans 

Six  days  before  Kennedy's  historic  announcement,  and  unaware  that 
it  was  coming,  John  Houbolt  shot  off  "a  hurried  non-edited  and  limited 
note"  of  three  single-spaced  pages  to  Bob  Seamans.  Confident  from  past 
meetings  that  Associate  Administrator  Seamans  was  greatly  interested  in 
the  subject  of  rendezvous,  Houbolt  took  the  liberty  of  cutting  through 
several  organizational  layers  to  communicate  with  him  directly. 

Houbolt's  message  was  straightforward  and  not  overly  passionate:  the 
situation  with  respect  to  the  development  of  new  launch  vehicles  was 
"deplorable."  The  Saturns  "should  undergo  major  structural  modifications" 
and  "no  committed  booster  plan"  beyond  Saturn  was  in  place.  Furthermore, 
NASA  was  still  not  attending  to  the  use  of  rendezvous  in  the  planned 
performance  of  the  Apollo  mission.  "I  do  not  wish  to  argue"  whether  "the 
direct  way"  or  "the  rendezvous  way"  is  best,  Houbolt  reassured  Seamans. 
But  "because  of  the  lag  in  launch  vehicle  developments,"  it  appeared  to 

249 


Spaceflight  Revolution 

him  that  "the  only  way  that  will  be  available  to  us  in  the  next  few  years 
is  the  rendezvous  way."  For  this  reason  alone  Houbolt  believed  that  it  was 
"mandatory"  that  "rendezvous  be  as  much  in  future  plans  as  any  item,  and 
that  it  be  attacked  vigorously."73  If  NASA  researchers  continued  to  dismiss 
LOR  totally  as  they  had  been  doing,  Houbolt  knew  that  someday  they  would 
be  sorry. 

If  Houbolt  had  known  that  an  ad  hoc  task  group  at  NASA  headquarters 
was  at  that  moment  in  the  midst  of  concluding  that  rendezvous  had  no  place 
in  the  lunar  landing  program,  his  letter  to  Seamans  might  have  been  more 
urgent.  But  nothing  in  his  letter  suggests  that  Houbolt  knew  anything 
about  the  meetings  of  the  Fleming  Committee.  Established  by  Seamans 
on  2  May,  the  job  of  this  committee  was  to  determine,  in  only  four  weeks, 
whether  a  manned  lunar  landing  was  in  fact  possible  and  how  much  it  would 
cost.  Chaired  by  NASA's  assistant  administrator  for  programs,  William 
A.  Fleming,  who — unlike  George  Low — was  known  to  be  neutral  on  the 
ideas  of  both  a  moon  landing  and  the  method  for  accomplishing  it,  this 
committee  eventually  recommended  a  lunar  landing  program  based  on  a 
three-stage  Nova.  In  essence,  the  Fleming  Committee  "avoided  the  question 
of  rendezvous  versus  direct  ascent."  Seeing  "no  reason  to  base  its  study 
on  a  risky  and  untried  alternative" — and  apparently  not  recognizing  that 
using  a  huge  and  unproven  launch  vehicle  was  also  "risky  and  untried" - 
the  committee  spent  all  four  weeks  trying  to  choose  between  solid-fuel  and 
liquid-fuel  propellants  for  the  Nova  stages.74 

Houbolt  and  the  other  LOR  advocates  at  Langley  would  have  been  dis- 
mayed. To  them,  development  of  the  rendezvous  concept  was  "the  obvious 
thing"  to  do  before  a  lunar  mission,  but  to  so  many  others,  space  rendezvous 
was  still  an  absurdly  complicated  and  risky  proposition.  Some,  like  Bob 
Seamans,  were  not  sure  what  to  think.  On  25  May,  after  hearing  President 
Kennedy's  speech,  the  associate  administrator  called  for  the  appointment 
of  yet  another  ad  hoc  committee,  this  one  "to  assess  a  wide  variety  of  pos- 
sible ways  for  executing  a  manned  lunar  landing."75  Bruce  T.  Lundin,  an 
associate  director  of  NASA  Lewis,  would  chair  this  new  committee. 

Whether  Houbolt 's  letter,  written  nearly  a  week  before,  directly  caused 
Seamans  to  create  the  Lundin  Committee  is  not  certain.  But  the  letter 
surely  was  a  contributing  factor  as  two  pieces  of  circumstantial  evidence  ap- 
pear to  indicate.*  First,  in  explaining  why  a  new  task  force  was  necessary, 


JLi 

Houbolt  believes  that  Seamans  created  the  Lundin  Committee  solely  in  response  to  his  letter.  "The 
story  I  got  [from  somebody  else  at  NASA  headquarters]  was  that  my  letter  jolted  Seamans,  and  he  got  up 
at  five  o'clock  in  the  morning,  got  on  the  phone,  called  several  people  and  said,  'Be  at  my  office  at  seven 
o'clock'.  .  .  .  And  then  they  formed  the  Lundin  Committee."  No  documents  exist  to  back  up  Houbolt's 
version  of  the  story.  Based  on  what  Seamans  has  said  about  the  formation  of  the  Lundin  Committee, 
there  is  no  doubt  that  Houbolt's  letter  did  contribute  directly  to  its  establishment  but  perhaps  not  as 
exclusively  as  Houbolt  has  heard.  (Houbolt  interview  with  author,  Williarrisburg,  Va.,  24  Aug.  1989, 
transcript,  p.  31,  Langley  Historical  Archives.) 

250 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

Seamans  explained  to  his  director  of  Advanced  Research  Programs  (Ira  H. 
Abbott)  and  his  director  of  Launch  Vehicle  Programs  (Don  R.  Ostrander) 
that  the  Fleming  Committee  was  finding  it  necessary  "to  restrict  its  con- 
siderations to  a  limited  number  of  techniques  by  which  it  is  feasible  to  ac- 
complish the  mission  in  the  shortest  possible  time."  Consequently,  "nu- 
merous other  approaches" — and  Seamans  specifically  mentioned  the  use  of 
rendezvous — were  not  currently  being  assessed.  Second,  Seamans,  a  busy 
man,  wrote  to  Houbolt  on  2  June,  thanking  him  for  his  comments  and  re- 
assuring the  distressed  Langley  researcher  that  "the  problems  that  concern 
you  are  of  great  concern  to  the  whole  agency."  NASA  headquarters  had  just 
organized  "some  intensive  study  programs,"  Seamans  informed  him,  without 
mentioning  the  Fleming  or  Lundin  committees  by  name.  These  programs 
"will  provide  us  a  base  for  decisions."76 

Some  historians  have  said  that  Seamans  made  sure  that  Houbolt  was  on 
the  Lundin  Committee;  this  is  untrue.  Houbolt  was  not  an  official  member 
of  that  committee.  Laurence  K.  Loftin,  Jr.,  was  Langley 's  representative, 
although  he  apparently  did  not  attend  all  the  meetings.  Houbolt  did  meet 
with  and  talk  to  the  committee  several  times,  and  in  his  view,  was  "the  real 
Langley  representative"  because  Loftin  did  not  attend  as  regularly  as  he.77 

The  idea  behind  the  Lundin  Committee,  at  least  as  originally  conceived 
by  Seamans,  was  to  take  an  open-minded  look  into  the  alternative  "modes" 
for  getting  to  the  moon.  Primarily,  Seamans  wanted  the  committee  to  ex- 
amine those  options  involving  "mission  staging  by  rendezvous"  and  "alter- 
native Nova  vehicles."  In  the  committee's  initial  meeting,  however,  that 
original  objective  seems  to  have  been  seriously  compromised.  Larry  Loftin, 
who  attended  the  opening  meeting  in  early  June  1961,  remembers  that  Bob 
Seamans  came  in  the  first  day  and  "sort  of  gave  us  our  marching  orders." 
Then  Abe  Silverstein,  director  of  the  Office  of  Space  Flight  Programs  at 
NASA  headquarters,  came  in  to  address  the  men: 

Well,  look  fellas,  I  want  you  to  understand  something.  I've  been  right  most  of  my 
life  about  things,  and  if  you  guys  are  going  to  talk  about  rendezvous,  any  kind  of 
rendezvous,  as  a  way  of  going  to  the  moon,  forget  it.  I've  heard  all  those  schemes  and 
I  don't  want  to  hear  any  more  of  them,  because  we're  not  going  to  the  moon  using 
any  of  those  schemes. 

With  those  words  of  warning,  which  completely  violated  the  reason  for 
forming  the  committee  in  the  first  place,  Silverstein  "stomped  out  of  the 
room.''78 

To  its  credit,  the  Lundin  Committee  disregarded  Silverstein 's  admonition 
and  considered  a  broad  range  of  rendezvous  schemes.  With  a  complete 
analysis  of  the  rendezvous  problems  by  Houbolt  and  assorted  insights  from 
invited  analysts  both  from  inside  and  outside  NASA,  the  group  looked  into 
mission  profiles  involving  rendezvous  in  earth  orbit,  in  transit  to  the  moon, 
in  lunar  orbit  before  landing,  in  lunar  orbit  after  takeoff  from  the  moon,  and 
in  both  earth  and  lunar  orbit.  The  committee  even  considered  the  idea  of 

251 


Space/light  Revolution 

a  lunar-surface  rendezvous.  This  involved  launching  a  fuel  cache  and  a  few 
other  unmanned  components  of  a  return  spacecraft  to  the  -moon's  surface — a 
payload  of  some  5000  pounds — and  then  landing  astronauts  separately  in  a 
second  spacecraft  whose  fuel  supply  would  be  exhausted  just  reaching  the 
moon.  The  notion,  as  absurd  as  it  now  sounds,  was  for  the  landed  astronauts 
to  leave  their  craft  and  locate  the  previously  deposited  hardware  (homing 
beacons  previously  landed  as  part  of  the  unmanned  Surveyor  program  were 
to  make  pinpoint  landings  possible)  and  then  to  assemble  and  fuel  a  new 
spacecraft  for  the  return  trip  home.  The  spacecraft  would  be  checked  out 
by  television  monitoring  equipment  before  sending  men  from  earth  to  the 
landing  area  via  a  second  spacecraft. 

Houbolt  thought  this  was  "the  most  harebrained  idea"  he  had  ever  heard. 
In  the  committee's  final  "summary  rating"  of  the  comparative  value  of 
the  various  rendezvous  concepts,  however,  lunar-surface  rendezvous  finished 
only  slightly  lower  than  did  Houbolt 's  LOR.  One  anonymous  committee 
member  (most  likely  the  JPL  representative)  chose  lunar-surface  rendezvous 
as  his  first  choice.79 

As  Houbolt  remembers  bitterly,  the  Lundin  Committee  "turned  down 
LOR  cold."  In  the  final  rating  made  by  the  six  voting  committee  members 
(Loftin  voted,  but  Houbolt  did  not),  it  finished  a  distant  third — receiving 
no  first-place  votes  and  only  one  second-place  vote.  Coming  in  far  ahead  of 
LOR  were  two  low-earth-orbit  rendezvous  schemes,  the  first  one  utilizing  two 
to  three  Saturn  C-3  boosters  and  the  other  involving  a  Saturn  C-l  plus  the 
Nova.  Both  were  concepts  strongly  favored  by  NASA  Marshall  staff,  who  by 
this  time  had  grabbed  onto  the  idea  of  EOR  for  its  potential  technological 
applications  to  the  development  of  an  orbiting  space  station.80 

Houbolt  was  devastated  when  he  heard  the  results.  To  have  LOR  placed 
on  the  same  level  as  the  ridiculous  lunar-surface  rendezvous  was  especially 
insulting.  He  had  given  the  Lundin  Committee  his  full-blown  pitch  complete 
with  foldout  sheet  and  slides.  "They'd  say,  'That  sounds  pretty  good,  John,' 
but  then  the  next  morning  the  same  guys  would  come  up  and  say,  'John, 
that's  no  good;  we  don't  like  it  at  all.'"  For  Houbolt,  this  perverse  reaction 
was  hard  to  understand.81  Loftin  reflects  on  the  general  fear  and  pessimism 
about  LOR  that  ultimately  ruled  the  committee: 

We  thought  it  was  too  risky.  Remember  in  1961  we  hadn't  even  orbited  Glenn  yet. 
We  certainly  had  done  no  rendezvous  yet.  And  to  put  this  poor  bastard  out  there, 
separate  him  in  a  module,  let  him  go  down  to  the  surface,  and  then  fire  him  back 
up  and  expect  him  to  rendezvous.  He  didn't  get  a  second  chance;  it  had  to  be  dead 
right  the  first  time.  I  mean  that  just  seemed  like  a  bit  much. 

Loftin  and  the  others  believed — incorrectly — that  LOR  offered  no  possibility 
for  a  rescue  mission.  In  earth  orbit,  if  something  went  wrong,  NASA  still 
might  be  able  to  save  its  astronauts.  Loftin  felt  along  with  the  others  that 
the  idea  of  LOR  was  just  "kind  of  absurd."82  The  Lundin  Committee  could 

252 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

not  bring  itself  to  acknowledge  that  all  the  other  mission-mode  options 
entailed  greater  risks. 

As  discouraged  as  John  Houbolt  was  after  the  Lundin  Committee's 
recommendation,  the  situation  would  soon  become  worse.  On  20  June, 
10  days  after  the  Lundin  Committee  delivered  its  blow,  Bob  Seamans 
formed  yet  another  task  force.  This  one  was  chaired  by  his  assistant  director 
of  launch  vehicle  programs,  Donald  H.  Heaton.  Following  up  on  the 
summary  ratings  and  recommendations  of  the  Lundin  Committee,  Seamans 
asked  Heaton's  group  to  focus  on  EOR  and  to  establish  the  program 
plans  and  the  supporting  resources  needed  to  accomplish  the  manned 
lunar  landing  mission  using  rendezvous  techniques.83  Trying  to  stay  within 
those  guidelines,  Heaton  refused  to  let  Houbolt,  an  official  member  of  this 
committee,  mention  LOR. 

Houbolt  felt  he  was  caught  in  a  bizarre  trap  of  someone  else's  making. 
He  was  one  of  the  strongest  believers  in  rendezvous  in  the  country,  and  that 
meant  either  kind  of  rendezvous.  Just  days  before  the  Heaton  Committee 
was  formed,  he  had  returned  from  France,  where  he  had  given  a  well-received 
formal  presentation  on  EOR  and  LOR  at  an  international  spaceflight 
symposium.84  He  and  his  Langley  associates  had  done  the  analysis,  and 
they  knew  that  LOR  would  work  better  than  EOR  for  a  manned  lunar 
landing.  He  pleaded  with  Heaton  to  study  LOR  as  well  as  EOR.  Heaton 
simply  answered,  "We're  not  going  to  do  that,  John.  It's  not  in  our  charter." 
"If  you  feel  strongly  enough  about  it,"  Heaton  challenged,  "write  your  own 
lunar-orbit  report."85 

Houbolt  eventually  did  just  that.  As  for  Heaton's  own  report,  which  was 
published  in  late  August,  it  concluded  that  rendezvous — EOR,  that  is — 
"offers  the  earliest  possibility  for  a  successful  manned  lunar  landing."  In 
postulating  the  design  of  the  spacecraft  that  would  make  that  type  of  lunar 
mission,  however,  the  Heaton  Committee  previewed  a  baseline  configuration 
that  Houbolt  regarded  as  a  "beast."  It  involved  "some  five  different  pieces 
of  hardware  that  were  going  to  be  assembled  in  the  earth-orbit  rendezvous," 
Houbolt  remembers.  "It  was  a  great  big  long  cigar."  In  his  opinion,  such  an 
unwieldy  concept  "would  hurt  the  cause  of  rendezvous."  He  feared  NASA 
engineers,  especially  in  the  STG,  would  read  the  Heaton  report  and  say, 
"Well,  we  knew  it  all  the  time:  these  rendezvous  guys  are  nuts."86 

Or  they  were  being  driven  nuts.  For  many  NASA  engineers,  the  summer 
of  1961  was  the  busiest  summer  of  their  lives;  it  certainly  was  the  busiest 
of  John  Houbolt's.  "I  was  living  half  the  time  in  Washington,  half  the 
time  on  the  road,  dashing  back  and  forth."87  In  mid- July  he  was  to  be  in 
Washington  again,  to  give  a  talk  at  the  NASA/Industry  Apollo  Technical 
Conference.  This  was  an  important  meeting  that  was  to  include  about  300 
potential  Project  Apollo  contractors.  It  was  so  important  that  Langley 
management  in  association  with  the  STG,  in  the  tradition  of  N  AC  A/NASA 
annual  inspections,  was  holding  a  formal  rehearsal  of  all  its  presentations 
prior  to  the  conference. 

253 


Spaceflight  Revolution 


Direct  Flight 


Lunar  Rendezvous 


Relativ.                             VT(7       )        «***                               \>       J          Relative                                  V»   7 

p^babUity                            N.1-X          P^^'y                             N--/           probability                                \JL/ 

.98*    1  Launch 

38*    1  Launch 

.98      1  Launch  part  of  escape  payload 

.94*    2  Establish  coasting  orbit 

345    2  Establish  coasting  orbit 

.94  6    2  Establish  coasting  orbit 

-948    3  Earth  escape 
.94      4  Midcoune  correction 
.94      S  Brake  into  lunar  orbit 
.90      6  Descent  and  landing 
.96      7  Ascent 
.98      8  Lunar  escape 
.98      9  Midcouree  correction 
.99     10  Re-entry 
—    11  Touch  down 

34      3  Earthescape 
39      4  Midcoune  correction 
33      5  Brake  into  lunar  orbit 
.95      6  Descent  with  lander 
SB      7  Ascent  with  lander 
35      8  Rendezvous 
39      9  Lunar  escape 
39    10  Midcoune  correction 
39    11  Re-entry 

.98      3  Launch  remainder  of  escape  payload 
.94      4  Establish  parking  orbit 
.94      5  Accelerate  to  transfer  to  outer  orbit 
.90      6  Rendezvous  (and  fuel  transfer) 
34      7  Earthescape 
.94      8  Midcoune  correction 
.94      9  Brake  into  lunar  orbit 
.90    10  Descent  and  landing 
.98    11  Ascent 

—    12  Touchdown 

.98     12  Lunar  escape 

.98    13  Midcoune  correction 

.99    14  Re-entry 

—    15  Touch  down 

SJ 

at 

9fi       t                            Overall  probability 

Probability  for  equal 

aa 

.78 

.26    _«  poundage  on  pad 

In  making  his  pitch  to  various  NASA  committees  and  study  groups  in  1961  and  1962, 
Houbolt  used  this  viewgraph  comparing  the  propulsion  steps  involved  in  direct  flight, 
LOR,  and  EOR,  thereby  demonstrating  the  much  higher  probability  of  success  with 
LOR. 


Houbolt  was  to  give  his  talk  at  the  end  of  rehearsals  because  he  had 
another  NASA  meeting  earlier  that  day  in  Washington.  "I  was  to  rush  out 
to  the  airport  at  Washington  National,  get  on  the  airplane,  they  were  to  pick 
me  up  here  and  then  bring  me  to  where  they  were  having  the  rehearsals." 
However,  when  he  arrived  breathless  at  the  airport,  the  airplane  could  not 
take  off.  In  refueling  the  aircraft,  the  ground  crew  had  spilled  fuel  on  one 
of  the  tires  and  the  Federal  Aviation  Administration  (FAA)  would  not  let 
the  plane  take  off  until  the  tire  had  been  changed.  That  made  Houbolt  a 
little  late  and  the  STG  member  waiting  for  him  a  little  impatient.  "They 
dashed  me  back  to  the  conference  room,"  and  with  all  of  the  other  rehearsals 
finished,  "everybody  was  sort  of  twiddling  their  thumbs,"  complaining, 
"'Where  the  hell  is  this  Houbolt?'"88 

With  a  brief  apology,  Houbolt  moved  right  into  his  talk.  Until  the  end, 
he  purposefully  said  nothing  about  LOR;  he  spoke  only  about  rendezvous 
in  general.  Then  he  showed  three  or  four  final  slides.  "There  is  a  very 
interesting  possibility  that  rendezvous  offers,"  Houbolt  ventured,  feeling  like 
a  lawyer  who  was  trying  to  slip  in  evidence  that  he  knew  the  judge  would 


254 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

not  allow,  "and  that  is  how  to  go  to  the  moon  in  a  very  simplified  way."  He 
then  described  the  whole  LOR  concept.89 

People  listened  politely  and  thanked  him  when  he  had  finished.  "That's 
a  damn  good  paper,  John,"  offered  Langley  Associate  Director  Charles 
Donlan.  "But  throw  out  all  that  nonsense  on  lunar-orbit  rendezvous." 
Houbolt  remembers  that  Max  Faget  and  several  other  members  of  the  STG 
piped  in  with  the  same  advice.90 

The  Lundin  Committee  had  been  strike  one  against  Houbolt:  LOR  had 
been  turned  down  cold.  The  Heaton  Committee  had  been  strike  two:  LOR 
would  not  even  be  considered.  Houbolt's  rehearsal  talk  was  in  a  sense  a 
third  strike.  But  at  least  all  three  had  been  swinging  strikes,  so  to  speak. 
Houbolt  had  used  each  occasion  to  promote  LOR,  and  he  had  given  his  best 
effort  each  time.  Furthermore,  he  was  allowed  a  few  more  times  at  bat.  An 
inning  was  over,  but  the  game  was  not. 

Houbolt's  next  time  at  bat  came  quickly,  in  August  1961,  when  he 
met  with  the  Golovin  Committee,  which  was  yet  another  of  Bob  Seamans' 
ad  hoc  task  forces.  Established  on  7  July  1961,  this  joint  Large  Launch 
Vehicle  Planning  Group  was  co-chaired  by  Nicholas  E.  Golovin,  Seamans' 
special  technical  assistant,  and  Lawrence  L.  Kavanau  of  the  DOD.  This 
committee  was  to  recommend  not  only  a  booster  rocket  for  Project  Apollo 
but  also  other  launch  vehicle  configurations  that  would  meet  the  anticipated 
needs  of  NASA  and  the  DOD.91 

Nothing  in  the  committee's  charge,  which  was  to  concern  itself  only  with 
large  launch  vehicle  systems,  necessitated  an  inquiry  into  the  LOR  scheme; 
however,  Eldon  W.  Hall,  Harvey  Hall,  and  Milton  W.  Rosen  (all  of  the  Office 
of  Launch  Vehicle  Programs)  and  members  of  the  Golovin  Committee  asked 
that  the  LOR  concept  be  presented  for  their  consideration  in  the  form  of  a 
mission  plan.92  This  was  to  be  done  as  part  of  a  systematic  comparative 
evaluation  of  three  types  of  rendezvous  operations  (earth-orbit,  lunar-orbit, 
and  lunar-surface)  and  direct  ascent  for  manned  lunar  landing.  The  Golovin 
Committee  assigned  the  study  of  EOR  to  Marshall  Space  Flight  Center, 
lunar-surface  rendezvous  to  JPL,  and  LOR  to  Langley.  The  NASA  Office 
of  Launch  Vehicle  Programs  would  itself  provide  the  information  on  direct 
ascent.93 

This  commitment  to  a  comparative  evaluation  of  the  mission  modes,  in- 
cluding LOR,  constitutes  a  critical  turning  point  in  the  torturous  intellectual 
and  bureaucratic  process  by  which  NASA  eventually  decided  upon  a  mission 
mode  for  Project  Apollo.  The  Golovin  Committee  would  not  conclude  in 
favor  of  LOR.  Its  final  somewhat  vacillating  recommendation,  made  in  mid- 
October,  was  in  favor  of  a  hybrid  rendezvous  scheme  that  combined  aspects 
of  both  EOR  and  LOR.  However,  the  committee's  preference  was  clearly 
for  some  form  of  rendezvous.  Lunar-surface  rendezvous,  JPL's  pet  project, 
had  been  ruled  out,  and  direct  ascent  was  fading  from  the  realm  of  possibil- 
ity. The  engineering  calculations  showed  clearly  that  any  single  rocket  that 
had  to  carry  all  the  fuel  necessary  for  completing  the  entire  lunar  mission 

255 


Space/light  Revolution 


.COMPARATIVE  SIZES  OF  MANNED  PROJECT 


EOR 


LOR 
C-5 


C-l 

(SATURN) 


GEMINI 


(TITAN 


MERCURY 
(ATLAS) 

i! 


ENA  D) 


L-1630 


COMPARISON    OF   LANDER   SIZES 

DIRECT  LANDING 


13.4' 


LUNAR  FERRY  OF 
LUNAR  RENDEZVOUS 


21.2' 


LUNAR 

EXCURSION 

VEHICLE 


L-1629 


By  using  drawings  that  compared  the  sizes  of  rockets  (top)  and  lunar  landing  vehicles 
(bottom),  Houbolt  tried  to  convince  the  nonbelievers  that  LOR  was  the  only  way  to 
go  to  the  moon. 

256 


Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 

was  not  a  realistic  option — especially  if  the  mission  was  to  be  accomplished 
anytime  close  to  President  Kennedy's  deadline. 

For  Houbolt  and  the  other  LOR  advocates,  the  work  of  the  Golovin 
Committee  meant  the  first  meaningful  opportunity  to  demonstrate  the 
merits  of  LOR  in  a  full-blown  comparison  with  the  other  viable  options. 
This  consideration  was  the  opportunity  that  Houbolt  had  been  asking  for 
in  all  of  his  previously  unsuccessful  briefings.  When  he  appeared  before  the 
Golovin  Committee  in  August,  "they  were  damn  impressed."  They  asked 
him,  to  his  delight,  whether  the  STG  knew  all  about  LOR.  Golovin  turned 
to  Aleck  C.  Bond,  the  STG's  representative  on  the  committee,  and  asked 
him  to  go  back  to  Langley  and  "check  with  your  fellows  on  what  they're 
doing  about  this."  A  few  days  later,  Houbolt  was  back  in  front  of  the  STG 
talking  to  them  about  the  same  thing  that  they  had  told  him  not  to  talk 
about  just  the  month  before.94 

The  STG,  with  the  Shepard  and  Grissom  flights  at  least  behind  them 
and  the  Golovin  Committee  now  urging  them  to  study  rendezvous,  started 
to  reconsider.  Thus  far,  as  other  historians  have  noted,  the  STG  had  "seen 
little  merit  in  any  form  of  rendezvous  for  lunar  missions,"  but  reserved 
"its  greatest  disdain  for  the  lunar  orbit  version."  Now,  at  least,  some 
STG  engineers  were  showing  solid  interest.  In  early  September  1961, 
Jim  Chamberlin,  who  had  asked  for  Houbolt's  material  after  hearing  the 
proposals  for  MORAD  and  MALLIR  five  months  earlier,  talked  to  Gilruth 
about  an  LOR  plan  for  a  lunar  landing  program  and  for  a  preparatory 
three-flight  rendezvous  experiment,  both  of  which  sounded  similar  to  the 
ideas  Houbolt  had  been  promoting.  Although  Gilruth  was  not  convinced  of 
the  merits  of  such  a  scheme,  he  was  open  to  their  further  evaluation.95 

Chamberlin's  notion  derived  in  part  from  the  STG's  August  1961  pro- 
posal for  an  accelerated  circumlunar  program;  this  proposal  appeared  as  an 
appendix  to  its  "Preliminary  Project  Development  Plan  for  an  Advanced 
Manned  Space  Program  Utilizing  the  Mark  II  Two-Man  Spacecraft."  In 
essence,  the  larger  document  called  for  the  start  of  what  became  known 
as  Project  Gemini,  the  series  of  two-man  rendezvous  and  docking  mis- 
sions in  earth  orbit  that  NASA  successfully  carried  out  between  March 
1965  and  November  1966.  But  the  idea  for  Project  Gemini,  as  proposed 
by  Chamberlin  at  least,  must  also  have  had  some  important  connection  to 
Houbolt's  April  1961  MORAD  proposal.96 


A  Voice  in  the  Wilderness 

During  the  late  summer  and  early  fall  of  1961,  Houbolt  was  busily 
preparing  the  formal  report  that  the  Golovin  Committee  had  requested. 
Except  for  his  "admiral's  page,"  much  of  the  analysis  in  favor  of  LOR  was 
still  in  a  loose  form.  With  John  Bird,  Art  Vogeley,  Max  Kurbjun,  and  the 
other  rendezvous  people  at  Langley,  he  set  out  to  document  their  research 

257 


Space/light  Revolution 

findings  and  demonstrate  what  a  complete  manned  lunar  landing  mission 
using  LOR  would  entail.  The  result  was  an  impressive  -two-volume  report 
entitled  "Manned  Lunar-Landing  through  Use  of  Lunar-Orbit  Rendezvous." 
Published  by  NASA  Langley  on  31  October  1961,  this  report  promoted  what 
its  principal  author,  John  Houbolt,  called  a  "particularly  appealing  scheme" 
for  performing  the  manned  lunar  landing  mission.97 

This  extremely  thorough  document  might  seem  sufficient  even  for  a 
zealous  crusader  like  Houbolt,  but  it  was  not.  The  Heaton  Committee  had 
submitted  its  final  report  in  August  1961 — a  report  with  which  Houbolt 
fervently  disagreed.  Houbolt  took  committee  chair  Heaton  up  on  his  remark 
about  submitting  his  own  report. 

On  15  November  1961,  Houbolt  "fired  off"  a  nine-page  letter  to  Seamans 
with  two  different  editions  of  his  LOR  admiral's  page  attached  to  it  without 
ever  thinking  that  it  might  cost  him  his  job.  He  was  again  bypassing  proper 
channels,  a  bold  move  for  a  government  employee,  and  appealing  directly 
to  the  associate  administrator.  "Somewhat  as  a  voice  in  the  wilderness," 
Houbolt's  letter  opened,  "I  would  like  to  pass  on  a  few  thoughts  that  have 
been  of  deep  concern  to  me  over  recent  months."  Houbolt's  main  complaint 
was  about  the  bureaucratic  guidelines  that  had  made  it  impossible  for  the 
Heaton  Committee  to  consider  the  merits  of  LOR.  "Do  we  want  to  go  to 
the  moon  or  not?,  and,  if  so,  why  do  we  have  to  restrict  our  thinking  to  a 
certain  narrow  channel?"  He  asked:  "Why  is  Nova,  with  its  ponderous  size 
simply  just  accepted,  and  why  is  a  much  less  grandiose  scheme  involving 
rendezvous  ostracized  or  put  on  the  defensive?" 

"I  fully  realize  that  contacting  you  in  this  manner  is  somewhat  unortho- 
dox," Houbolt  admitted,  "but  the  issues  at  stake  are  crucial  enough  to  us  all 
that  an  unusual  course  is  warranted."  Houbolt  realized  that  Seamans  might 
feel  that  he  was  "dealing  with  a  crank."  "Do  not  be  afraid  of  this,"  Houbolt 
pleaded.  "The  thoughts  expressed  here  may  not  be  stated  in  as  diplomatic 
a  fashion  as  they  might  be,  or  as  I  would  normally  try  to  do,  but  this  is  by 
choice."  Most  important  was  that  Seamans  hear  his  heartfelt  ideas  directly 
and  "not  after  they  have  filtered  through  a  score  or  more  of  other  people, 
with  the  attendant  risk  [that]  they  may  not  even  reach  you."s 

It  took  two  weeks  for  Seamans  to  reply  to  Houbolt's  extraordinary  letter. 
When  he  did,  the  associate  administrator  agreed  that  "it  would  be  extremely 
harmful  to  our  organization  and  to  the  country  if  our  qualified  staff  were 
unduly  limited  by  restrictive  guidelines."  He  assured  Houbolt  that  NASA 
would  in  the  future  be  paying  more  attention  to  LOR.100 

Seamans  also  informed  him  that  he  had  passed  his  long  letter  with  its 
attachments  on  to  Brainerd  Holmes,  who  had  just  replaced  Abe  Silverstein 
as  head  of  the  Office  of  Manned  Space  Flight  (recently  renamed  Space  Flight 
Programs).  Unlike  Seamans,  who  apparently  was  not  overly  bothered  by  the 
letter  being  sent  out  of  formal  organizational  channels,  Holmes  "didn't  like 
it  at  all"  and  said  so  when  in  turn  he  passed  Houbolt's  letter  on  to  George 
Low,  his  director  of  spacecraft  and  flight  missions.  Low  was  more  forgiving. 

258 


Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 


Although  he  too  had  been  slow  to  ac- 
cept the  LOR  scheme,  NASA  leader 
George  M.  Low  eventually  became  a 
devout  believer  not  only  in  LOR  but 
also  in  the  essential  role  played  by 
Houbolt  in  its  adoption. 
L-70-1270 


Although  he  conceded  that  Houbolt  probably  should  have  followed  standard 
procedures,  he  found  the  basic  message  "relatively  sound."  He,  too,  felt  that 
"the  bug  approach"  might  yet  prove  to  be  "the  best  way  of  getting  to  the 
moon"  and  that  NASA  needed  to  give  it  as  much  attention  as  any  other 
alternative.  At  the  end  of  the  memo  to  Holmes  in  which  he  passed  on 
these  feelings,  Low  recommended  that  Houbolt  be  invited  to  Washington  to 
present  in  detail  Langley's  plan  for  a  manned  lunar  landing  via  LOR.  Low 
even  went  so  far  as  to  suggest  that  Houbolt  should  be  made  a  member  of 
Holmes's  staff.101 

That  never  happened,  but  another  person  who  joined  Holmes's  staff  at 
this  time,  Dr.  Joseph  F.  Shea,  came  to  play  a  major  role  in  supporting 
Houbolt 's  ideas  and  making  the  eventual  decision  in  favor  of  LOR.  Shea 
arrived  at  NASA  the  first  week  of  January  1962  as  Holmes's  deputy  director 
for  spaceflight  systems.  From  1956  to  1959  the  energetic  engineer  from 
the  Bronx  had  served  as  the  systems  engineer  at  Bell  Laboratories  for  a 
radio  guidance  project  involving  the  Titan  I  rocket.  In  1959  he  moved  to 
General  Motors,  where  he  ran  the  advanced  development  operation  for  its 
A.  C.  Sparkplug  Division.  His  major  achievement  while  in  this  job  was  to 
win  a  contract  for  the  development  of  an  inertial  guidance  system  for  the 
Titan  II.102 

With  NASA,  Joe  Shea  found  himself  thrust  into  the  job  of  sorting  out  the 
best  means  of  accomplishing  the  lunar  landing  mission.  During  his  first  days 

259 


Space/light  Revolution 

in  office,  Brainerd  Holmes  came  to  see  him,  with  his  copy  of  Houbolt's  letter 
in  hand.  Shea  perused  the  long  letter  and  was  taken  down  to  Seamans'  office 
where  Seamans  asked  him  if  he  thought  anything  of  value  could  be  found 
in  Houbolt's  message.  Having  received  an  unsure  response,  Seamans  then 
advised  the  young  systems  engineer  that  NASA  really  did  not  know  how  it 
was  going  to  go  to  the  moon.  Shea  answered  tactfully,  "I  was  beginning  to 
get  that  same  suspicion."103 

"Shea  didn't  know  much  about  what  was  going  on,"  John  Houbolt 
remembers,  but  quickly  he  became  informed.  Within  days  of  his  meeting 
with  Seamans  and  Holmes  about  the  Houbolt  letter,  Shea  was  at  Langley  for 
a  private  conversation  with  Houbolt  and  for  a  general  briefing  attended  by 
Langley  management  and  the  leadership  of  the  STG.  Going  into  the  meeting, 
if  Shea  had  a  preference  for  any  one  lunar  mission  mode,  it  was  a  weak  one  for 
EOR,  but  after  reading  Houbolt's  letter  to  Seamans  and  knowing  Seamans' 
sympathetic  reaction  to  it,  Shea  was  not  adverse  to  other  options.  Shea  was 
an  open-minded  man  who  "prided  himself  on  going  wherever  the  data  took 
him."104 

This  time  the  data  took  him  to  LOR.  When  Houbolt  finished  his  much- 
practiced  pitch,  the  receptive  Shea  admitted  that  the  analysis  looked  "pretty 
good"  to  him.  The  new  boy  on  the  block  of  manned  spaceflight  then  turned 
to  Gilruth,  Faget,  and  other  members  of  the  STG  and  asked  them  politely  if 
they,  too,  had  been  thinking  along  the  lines  of  LOR.  Having  gotten  the  word 
about  the  general  skepticism  to  Houbolt's  ideas,  Shea  expected  a  negative 
reaction.  He  did  not  receive  one.  Instead,  the  STG  leaders  responded  in 
a  mildly  positive  way  that  signified  to  Shea,  as  the  discussion  continued, 
that  "actually,  they  had  been  doing  some  more  thinking  about  lunar-orbit 
rendezvous  and,  as  a  matter  of  fact,  they  were  beginning  to  think  it  was  a 
good  idea."105 

Shea  returned  to  Washington  convinced  that  LOR  was  a  viable  option 
for  Apollo  and  that  the  next  step  was  for  NASA  to  contract  for  an  even 
more  detailed  study  of  its  potential.  On  1  March  1962,  eight  days  after 
astronaut  John  Glenn's  historic  three-orbit  flight  in  the  Mercury  spacecraft 
Friendship  7,  NASA  named  Chance  Vought  Corporation  as  the  contractor 
to  study  spacecraft  rendezvous.  The  firm  had  on  staff  one  of  the  original 
proponents  of  LOR,  Tom  Dolan.  At  Langley  on  29  March  1962,  a  group  of 
researchers  led  by  Houbolt  briefed  a  Chance  Vought  team  on  the  center's 
LOR  research  and  mission  plan.  On  2  and  3  April,  Shea  presented  LOR 
as  a  possible  mission  mode  for  Apollo  in  a  headquarters  meeting  that  was 
attended  by  representatives  of  all  the  NASA  centers.106  The  final  decision 
to  select  LOR  for  Apollo  was  about  to  be  made. 

The  LOR  Decision 

In  the  months  following  Houbolt's  second  letter  to  Seamans,  NASA  gave 
LOR  the  serious  consideration  that  Houbolt  had  long  been  crusading  for. 

260 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

To  the  surprise  of  many,  both  inside  and  outside  the  agency,  the  dark- 
horse  candidate  became  the  front  runner.  Several  factors  worked  in  LOR's 
favor.  First,  many  were  becoming  disenchanted  with  the  idea  of  direct  ascent 
because  of  the  time  and  money  required  to  develop  the  huge  Nova  rocket. 
Second,  technical  apprehension  was  growing  over  how  the  relatively  large 
spacecraft  demanded  even  by  EOR  would  be  able  to  maneuver  to  a  soft 
and  pinpoint  landing  on  the  moon.  As  Langley's  expert  on  the  dynamics 
of  rendezvous,  Art  Vogeley,  has  explained,  "The  business  of  eyeballing  that 
thing  down  to  the  moon  really  didn't  have  a  satisfactory  answer.  The  best 
thing  about  LOR  was  that  it  allowed  us  to  build  a  separate  vehicle  for 
landing."107 

The  first  major  group  to  break  camp  in  favor  of  LOR  was  Bob  Gilruth's 
STG,  which  during  the  critical  months  of  the  Apollo  mission  mode  debate 
was  harried  not  only  with  the  planning  for  the  first  Mercury  orbital  flight 
but  also  with  packing  and  leaving  for  its  new  home,  the  Manned  Spacecraft 
Center  in  Houston.  During  an  interview  in  the  late  1980s,  Houston's  Max 
Faget  recalled  the  details  of  how  the  STG  Manned  Spacecraft  Center  finally 
became  convinced  that  LOR  was  the  right  choice.  By  early  1962, 

we  found  ourselves  settling  into  a  program  that  was  not  easy  to  run,  because  so 
many  different  groups  were  involved.  In  particular,  we  were  concerned  about  the  big 
landing  rocket,  because  landing  on  the  moon  would,  of  course,  be  the  most  delicate 
part  of  the  mission.  The  landing  rocket's  engine,  which  would  be  controlled  by  the 
astronauts,  would  have  to  be  throttleable,  so  that  the  command-and-service  module 
could  hover,  and  move  this  way  and  that,  to  find  a  proper  place  to  touch  down. 

Obtaining  that  capability  meant  the  need  for  "a  really  intimate  interface, 
requiring  numerous  connections,  between  the  two  elements,"  as  well  as 
between  Houston  and  NASA  Lewis. 

Accordingly,  we  invented  a  new  proposal  for  our  own  and  von  Braun's  approach.  It 
involved  a  simpler  descent  engine,  called  the  lunar  crasher,  which  Lewis  would  do. 
It  wouldn't  be  throttleable,  so  the  interface  would  be  simpler,  and  it  would  take 
the  astronauts  down  to  a  thousand  feet  above  the  lunar  surface.  There  it  would 
be  jettisoned,  and  it  would  crash  onto  the  moon.  Then  there  would  be  a  smaller, 
throttleable  landing  stage  for  the  last  thousand  feet,  which  we  would  do,  so  that  we 
would  be  in  charge  of  both  sides  of  that  particular  interface. 

At  that  point,  however,  Faget  and  his  colleagues  in  Texas  "ran  into  a  real 
wall."108 

Initially  their  thinking  had  been  that  the  landing  would  be  done  auto- 
matically with  radar  and  instrument  control.  Then  the  astronauts,  along 
with  a  growing  number  of  NASA  engineers  (primarily  at  Langley),  began  to 
argue  that  the  astronauts  were  going  to  need  complete  control  during  the 
last  phases  of  landing  and  therefore  would  require  a  wide  range  of  visibil- 
ity from  the  descending  spacecraft.  How  to  provide  that  visibility  "with  a 


Space/light  Revolution 

landing  rocket  big  enough  to  get  the  command-and-service  module  down  to 
the  lunar  surface  and  wide  enough  to  keep  it  upright"  was  the  problem  that 
Houston  began  tackling  in  early  1962  and  found  very  quickly  they  could  not 
solve.  "We  toyed  with  various  concepts,"  Faget  remembers,  such  as  putting 
a  front  viewing  porch  on  the  outside  or  a  glass  bubble  on  top  of  the  CM 
similar  to  the  cockpit  of  a  helicopter.  But  all  the  redesigns  had  serious  flaws. 
For  example,  "the  porch  would  have  to  be  jettisoned  before  lift-off  from  the 
moon,  because  it  would  unbalance  the  spacecraft."  "It  was  a  mess,"  Faget 
admitted.  "No  one  had  a  winning  idea.  Lunar-orbit  rendezvous  was  the 
only  sensible  alternative." ] 

Houbolt's  role  in  the  STG's  eventual  "coming-around"  to  LOR  cannot  be 
described  without  upsetting  someone — or  at  least  questioning  the  accuracy 
of  someone's  memory.  Faget,  Gilruth,  and  others  associated  with  the 
Manned  Spacecraft  Center  believe  that  Houbolt's  activities  were  "useful" 
but  hardly  as  vital  as  many  others,  notably  Houbolt  himself,  believe  them 
to  be.  "John  Houbolt  just  assumed  that  he  had  to  go  to  the  very  top," 
Gilruth  has  explained,  "he  never  talked  to  me."  Gilruth  maintains  that  "the 
lunar  orbit  rendezvous  would  have  been  chosen  without  Houbolt's  somewhat 
frantic  efforts."  The  "real  work  of  convincing  the  officials  in  Washington  and 
Huntsville,"  he  says,  was  done  "by  the  spacecraft  group  in  Houston  during 
the  six  or  eight  months  following  President  Kennedy's  decision  to  fly  to 
the  moon."  Gilruth's  group  sold  the  concept,  first  to  Huntsville  and  then, 
together  with  von  Braun,  to  NASA  headquarters.  Houbolt's  out-of-channels 
letter  to  Seamans  was,  in  Gilruth's  opinion,  irrelevant.11 

Houbolt  calls  the  STG's  version  self-serving  "baloney."  He  talked  to 
Gilruth  or  his  people  many  times,  and  not  once  did  they  tell  him  that  they 
were  really  on  his  side.  If  just  one  time  Gilruth  or  some  other  influential 
officer  in  the  manned  space  program  had  said  to  him,  "You  can  stop  fighting. 
We  are  now  on  your  side;  and  we'll  take  it  from  here,"  then,  Houbolt  claims, 
he  would  have  been  satisfied.  But  they  never  said  that  to  him,  and  they 
certainly  did  not  say  it  "during  the  six  or  eight  months"  after  Kennedy's 
speech.  In  fact,  their  words  always  suggested  just  the  opposite.  Not  until 
early  1962,  after  prodding  from  Joe  Shea,  did  the  STG  give  any  indication 
that  it,  too,  was  interested  in  LOR.11 

Significantly,  the  outsiders  or  third  parties  to  the  question  of  Houbolt's 
role  in  influencing  the  STG's  position  on  the  mission  mode  for  Apollo  tend 
to  side  with  Houbolt.  Bob  Seamans  remembers  the  STG  showing  nothing 
but  disdain  for  LOR  during  1961. 112  George  Low  agrees.  To  the  best 
of  his  recollection,  "it  was  Houbolt's  letter  to  Seamans  that  brought  the 
Lunar  Orbit  Rendezvous  Mode  back  into  the  picture."  Only  after  that 
did  a  group  within  the  STG  under  Owen  Maynard  begin  to  study  LOR. 
"Based  on  Houbolt's  input"  and  on  the  results  of  the  systems  engineering 
studies  carried  out  at  the  behest  of  Joe  Shea's  Office  of  Manned  Space 
Flight  Systems,  "the  decision  was  finally  made"  about  the  lunar  landing 
mission  mode.  "Without  a  doubt,"  in  Low's  view,  the  letter  Houbolt  sent 

262 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

to  Seamans  in  November  1961  and  the  discussions  at  headquarters  that  it 
provoked  "were  the  start  of  bringing  LOR  into  Apollo."113 

One  final  piece  of  testimony  from  an  informed  third  party  supports 
the  importance  of  Houbolt's  role  in  convincing  the  STG  of  the  benefits 
of  LOR.  Starting  in  late  1961,  NACA  veteran  Axel  Mattson  served  as 
NASA  Langley's  technical  liaison  officer  at  the  Manned  Spacecraft  Center. 
Mattson,  who  was  responsible  for  coordinating  the  other  NASA  centers 
for  the  first  NASA  inspection,  maintained  a  small  office  at  the  Houston 
facility  for  the  timely  moving  of  technical  information  between  Langley  and 
Gilruth's  recently  removed  STG.  Mattson's  operation  was  not  high  profile, 
nor  was  it  supposed  to  be.  According  to  the  agreement  that  had  been  worked 
out  between  Gilruth  and  Langley  Director  Floyd  Thompson,  Mattson  was 
to  spend  most  of  his  time  with  the  engineers  who  were  working  on  Mercury 
problems.11 

In  early  1962,  sometime  after  the  Shea  briefing,  Langley  sent  Houbolt  to 
Houston.  The  purpose  of  his  visit  was,  in  Mattson's  words,  "to  get  the  STG 
people  really  to  agree  that  [LOR]  was  the  best  way  to  go  and  to  support 
it."  Mattson  took  him  to  practically  everyone  who  had  some  interest  in  the 
mission  mode  issue,  and  Houbolt  told  them  about  LOR  and  answered  all 
their  questions.  At  the  end  of  the  day,  Mattson  felt  that  "it  was  all  over." 
"We  had  the  support  of  the  Manned  Spacecraft  Center"  for  LOR.115 

Significantly,  on  6  February  1962,  Houbolt  and  former  Langley  engineer 
Charles  W.  Mathews  of  the  Manned  Spacecraft  Center  gave  a  joint  presenta- 
tion on  rendezvous  to  the  Manned  Space  Flight  Management  Council.  This 
council  was  a  special  body  formed  by  Brainerd  Holmes  in  December  1961 
to  identify  and  resolve  difficulties  in  the  manned  spaceflight  program  on  a 
month-to-month  basis.  In  their  presentation  the  two  engineers  compared  the 
merits  of  LOR  and  EOR  and  clearly  favored  LOR.  Gilruth  had  telephoned 
Houbolt  personally  to  ask  him  to  give  this  talk.  In  Houbolt's  memory,  the 
invitation  was  "the  first  concession"  that  Gilruth  had  ever  made  to  him 
regarding  LOR.116 

With  the  STG  now  firmly  behind  LOR,  its  adoption  became  a  contest 
between  the  Manned  Spacecraft  Center  in  Houston  and  the  Marshall  Space 
Flight  Center  in  Huntsville.  Marshall  was  a  bastion  of  EOR  supporters.  Von 
Braun's  people  recognized  two  things:  EOR  would  require  the  development 
of  advanced  versions  of  Marshall's  own  Saturn  booster,  and  the  selection  of 
EOR  for  the  lunar  landing  program  would  require  construction  of  a  platform 
in  earth  orbit  that  could  have  many  uses  other  than  for  Apollo.  For  this 
reason,  space  station  advocates — who  existed  in  droves  at  the  Alabama 
facility — were  enthusiastic  about  EOR.117  To  this  day,  many  of  them  feel 
that  EOR  would  have  had  the  best  long-term  results. 

But  von  Braun,  their  own  director,  would  disappoint  them.  During  the 
spring  of  1962,  the  transplanted  German  rocket  designer  made  the  decision 
to  throw  his  weight  behind  LOR.  He  surprised  his  staff  with  this  shocking 


263 


Space/light  Revolution 


L-66-3090 

Taking  charge  of  every  situation,  Wernher  von  Braun  (second  from  left)  entertains 
his  hosts  during  a  visit  to  Langley  in  April  1966.  To  the  far  right  stand  Floyd 
Thompson  and  Charles  Donlan. 


announcement  at  the  end  of  a  day-long  briefing  given  to  Joe  Shea  at  Marshall 
on  7  June  1962: 

We  at  the  Marshall  Space  Flight  Center  readily  admit  that  when  first  exposed  to  the 
proposal  of  the  Lunar  Orbit  Rendezvous  Mode  we  were  a  bit  skeptical — particularly 
of  the  aspect  of  having  the  astronauts  execute  a  complicated  rendezvous  maneuver 
at  a  distance  of  240,000  miles  from  the  earth  where  any  rescue  possibility  appeared 
remote.  In  the  meantime,  however,  we  have  spent  a  great  deal  of  time  and  effort 
studying  the  four  modes  [EOR,  LOR,  and  two  Direct  Ascent  modes,  one  involving 
the  Nova  and  the  other  a  Saturn  C-5],  and  we  have  come  to  the  conclusion  that  this 
particular  disadvantage  is  far  outweighed  by  [its]  advantages.  .  .  . 

We  understand  that  the  Manned  Spacecraft  Center  was  also  quite  skeptical  at 
first  when  John  Houbolt  advanced  the  proposal  of  the  Lunar  Orbit  Rendezvous  Mode, 
and  that  it  took  them  quite  a  while  to  substantiate  the  feasibility  of  the  method  and 
finally  endorse  it. 

Against  this  background  it  can,  therefore,  be  concluded  that  the  issue  of  'invented 
here'  versus  'not  invented  here'  does  not  apply  to  either  the  Manned  Spacecraft  Center 
or  the  Marshall  Space  Flight  Center;  that  both  Centers  have  actually  embraced  a 
scheme  suggested  by  a  third  source.  ...  I  consider  it  fortunate  indeed  for  the 
Manned  Lunar  Landing  Program  that  both  Centers,  after  much  soul  searching,  have 
come  to  identical  conclusions. 

264 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

The  persuasive  von  Braun  then  elaborated  on  "why  we  do  not  recommend" 
the  direct  ascent  and  EOR  modes,  and  "why  we  do  recommend  the  Lunar- 
Orbit  Rendezvous  Mode."118 

For  Marshall  employees  and  many  other  people  inside  NASA,  von  Braun's 
announcement  seemed  to  represent  a  type  of  closure,  that  is,  the  culmination 
of  a  sociopolitical  process  "when  a  consensus  emerges  that  a  problem  arising 
during  the  development  of  a  technology  has  been  solved."  In  this  case,  it 
was  a  very  undemocratic  form  of  closure,  coming  from  von  Braun  himself, 
with  very  little  support  from  his  own  people.  But  NASA,  of  course,  was 
not  a  democratic  organization.  For  closure  to  occur  and  LOR  to  become 
the  mission  mode  for  Apollo,  referendum  or  consensus  was  not  necessary;  it 
only  required  that  a  decision  be  made  and  supported  by  a  few  key  people: 
von  Braun,  Bob  Gilruth,  Bob  Seamans,  Administrator  James  Webb,  and 
President  Kennedy.11 

How  von  Braun  was  persuaded  is  a  historically  significant  matter.  Al- 
though some  questions  about  his  motives  remain  unanswered,  one  apparent 
factor  in  his  conversion  was  that  he  understood  the  necessity  of  moving 
ahead  with  the  program  if  NASA  was  to  meet  President  Kennedy's  dead- 
line. No  progress  was  possible  until  the  decision  about  the  mission  mode 
was  made.  Both  the  Manned  Spacecraft  Center  and  Langley's  John  Houbolt 
had  worked  to  convince  von  Braun  to  come  over  to  their  side.  In  April 
1962  Houbolt  sent  von  Braun  several  papers  prepared  at  Langley  on  a  lu- 
nar landing  mission  using  LOR,  including  the  published  two- volume  report. 
Von  Braun  had  requested  the  papers  personally  after  hearing  Houbolt 's  pre- 
sentation at  NASA  headquarters.  Von  Braun  not  only  passed  copies  of  the 
Langley  papers  to  Hermann  Koelle  in  Marshall's  Future  Projects  Office  but 
also,  after  making  his  unexpected  announcement  in  favor  of  LOR  to  the 
stunned  crowd  of  Marshall  employees  in  early  June,  reciprocated  by  send- 
ing Houbolt  a  copy  of  the  remarks  he  had  made.  This  was  a  noteworthy 
courtesy.  The  final  sentence  of  the  cover  letter  asked  Houbolt  to  "please 
treat  this  confidentially  since  no  final  decision  on  the  mode  has  yet  been 
made."120 

The  LOR  decision  was  finalized  in  the  following  weeks  when  the  two 
powerful  groups  of  converts  at  Houston  and  Huntsville,  along  with  the 
original  little  band  of  true  believers  at  Langley,  persuaded  key  officials  at 
NASA  headquarters,  notably  Administrator  James  Webb,  who  had  been 
holding  out  for  direct  ascent,  that  LOR  was  the  only  way  to  land  on  the 
moon  by  1969.  With  the  key  players  now  supporting  the  concept,  the  NASA 
Manned  Space  Flight  Management  Council  announced  on  22  June  1962  that 
it  favored  LOR.  On  11  July,  the  agency  announced  that  it  had  selected  the 
mode  for  Apollo.  Webb  made  the  announcement  even  though  President 
Kennedy's  science  adviser,  Dr.  Jerome  Wiesner,  remained  firmly  opposed  to 
LOR.12f 

On  the  day  that  NASA  made  the  public  announcement,  Houbolt  was  giv- 
ing a  paper  on  the  dynamic  response  of  airplanes  to  atmospheric  turbulence 

265 


Space/light  Revolution 


On  14  March  1969,  four  months  before 
the  first  lunar  landing,  Life  magazine  fea- 
tured the  LEM  on  its  cover  (right).  The 
magazine  proposed  a  cover  featuring  John 
Houbolt  (below)  but  did  not  use  it  because 
of  NASA 's  concern  for  giving  too  much 
credit  to  any  one  person  for  the  decision 
to  go  to  the  moon  via  LOR. 


L-66-6301 


266 


Enchanted  Rendezvous:  The  Lunar- Orbit  Rendezvous  Concept 

at  a  meeting  of  NATO's  Advisory  Group  for  Aeronautical  Research  and 
Development  (AGARD)  in  Paris.122  His  division  chief,  Isadore  E.  ("Ed") 
Garrick,  was  also  at  the  meeting.  A  talented  applied  mathematician  who 
had  been  working  at  Langley  since  the  1930s,  Garrick  had  witnessed  the 
evolution  of  his  assistant's  ideas  on  space  navigation  and  rendezvous.  He 
had  listened  sympathetically  to  all  of  Houbolt's  stories  about  the  terrible 
things  that  had  been  blocking  a  fair  hearing  of  LOR. 

While  at  the  AGARD  meeting  in  Paris,  Garrick  saw  a  little  blurb  in  the 
overseas  edition  of  the  New  York  Herald  Tribune  about  NASA's  decision 
to  use  LOR.  Garrick  showed  the  paper  to  Houbolt,  who  had  not  seen  it, 
shook  Houbolt's  hand,  and  said,  "Congratulations,  John.  They've  adopted 
your  scheme.  I  can  safely  say  I'm  shaking  hands  with  the  man  who  single- 
handedly  saved  the  government  $20  billion."123 

In  the  ensuing  years,  whenever  the  question  of  Houbolt's  importance 
for  the  LOR  decision  came  up  for  discussion,  Garrick  said  that  he  was 
"practically  certain  that  without  John  Houbolt's  persistence  it  would  have 
taken  several  more  years  for  LOR  to  have  been  adopted."  Although  "the 
decisions  of  many  other  people  were  essential  to  the  process"  and  although 
"there  is  no  controversy  that  Houbolt  had  help  from  others,  . . .  the  essential 
prime  mover,  moving  'heaven  and  earth'  to  get  the  concepts  across,  remains 
Houbolt  himself."124 


Postscript 

Whether  NASA's  choice  of  LOR  would  have  been  made  in  the  summer  of 
1962  or  at  any  later  time  without  the  research  information,  the  commitment, 
and  the  crusading  zeal  of  Houbolt  remains  a  matter  for  historical  conjecture. 
His  basic  contribution,  however,  and  that  of  his  Langley  associates  who  in 
their  more  quiet  ways  also  developed  and  advocated  LOR,  seem  now  to  be 
beyond  debate.  They  were  the  first  in  NASA  to  recognize  the  fundamental 
advantages  of  the  LOR  concept,  and  for  a  critical  period  in  the  early  1960s, 
they  were  also  the  only  ones  inside  the  agency  to  foster  and  fight  for  it.  The 
story  of  the  genesis  of  LOR  underscores  the  vital  role  occasionally  played  by 
the  unpopular  opinion.  It  testifies  to  the  essential  importance  of  the  single 
individual  contribution  even  within  the  context  of  a  large  organization  based 
on  teamwork.  And  it  demonstrates  the  importance  of  passionate  persistence 
in  the  face  of  strong  opposition  and  the  pressure  for  conformity. 

Thousands  of  factors  contributed  to  the  ultimate  success  of  the  Apollo 
lunar  landing  missions,  but  no  single  factor  was  more  essential  than  the 
concept  of  LOR.  Without  NASA's  adoption  of  this  stubbornly  held  minority 
opinion,  the  United  States  might  not  have  reached  the  moon  by  the  end  of 
the  decade  as  President  Kennedy  had  promised.  Without  LOR,  possibly  no 
one  even  now — near  the  beginning  of  the  twenty-first  century — would  have 
stepped  onto  the  moon. 

267 


Space/light  Revolution 

No  less  an  authority  than  George  Low  has  expressed  this  same  judgment. 
"It  is  my  opinion  to  this  day,"  Low  wrote  in  1982,  "that  had  the  Lunar  Orbit 
Rendezvous  Mode  not  been  chosen,  Apollo  would  not  have  succeeded."  All 
of  the  other  modes  "would  have  been  so  complex  technically,  that  there 
would  have  been  major  setbacks  in  the  program,  and  it  probably  would 
have  failed  along  the  way."  Low  has  also  gone  on  record  with  his  belief,  that 
without  "John  Houbolt's  persistence  in  calling  this  method  to  the  attention 
of  NASA's  decision  makers,"  and  "without  Houbolt's  letter  to  Seamans  (and 
the  work  that  backed  up  that  letter),"  NASA  "might  not  have  chosen  the 
Lunar  Orbit  Rendezvous  Mode."  Houbolt's  commitment  was  a  key  factor 
in  the  adoption  of  LOR  and  was  "a  major  contribution  to  the  success  of 
Apollo  and,  therefore,  to  the  Nation." 125 


At  4:17  p.m.  (EDT)  on  20  July  1969,  John  Houbolt,  now  a  senior  con- 
sultant with  the  innovative  Aeronautical  Research  Associates  of  Princeton, 
New  Jersey,  sat  inconspicuously  as  one  of  the  invited  guests  and  dignitaries 
in  the  viewing  room  of  Mission  Control  at  the  Manned  Spacecraft  Center  in 
Houston.  Like  so  many  others  around  the  world  at  that  moment,  he  listened 
in  wonder  to  the  deliberately  spoken  yet  wildly  dramatic  words  of  Apollo  1 1 
astronaut  Neil  Armstrong:  "Houston,  Tranquility  Base  here.  The  Eagle  has 
landed." 

Alternate  cheering  and  shushing  followed  that  precious  moment,  when 
Americans  landed  and  stepped  onto  the  moon  for  the  first  time.  Turn- 
ing from  his  seat,  NASA's  master  rocketeer,  Wernher  von  Braun,  found 
Houbolt's  eye  among  all  the  others,  gave  him  the  okay  sign,  and  said  to  him 
simply,  "John,  it  worked  beautifully."126 


268 


9 


Skipping  "The  Next  Logical  Step" 


The  reason  some  of  us  wanted  EOR  was  not  just  to  go 
to  the  moon  but  to  have  something  afterwards:  orbital 
operations,  a  space  station,  a  springboard.  LOR  was 
a  one-shot  deal,  very  limited,  very  inflexible. 

— Jesco  von  Puttkamer 
NASA  Marshall  engineer 

By  1969,  it  was  apparent  that  there  was  no  logical  se- 
quel to  the  lunar  landing,  and  that  the  agency  would 
have  to  redeploy  its  resources  in  a  radically  differ- 
ent direction.  Had  NASA  selected  earth- orbit  ren- 
dezvous instead,  the  lunar  landing  could  still  have 
been  achieved  and  NASA  would  have  had  at  least  a 
ten-year  start  on  deploying  an  orbiting  space  station, 
rather  than  waiting  until  1982  to  let  contracts  for  its 
design. 

—Hans  Mark  and  Arnold  Levine 
The  Management  of  Research  Institutions 


No  decision  in  NASA  history  had  a  greater  impact  on  the  course  of 
the  American  space  program  than  the  selection  of  LOR  as  the  mission 
mode  for  Apollo.  The  LOR  decision  led  to  a  total  of  six  successful  lunar 
landing  missions  by  1972,  thus  enabling  the  United  States  to  win  the  most 
important  leg  of  the  space  race.  Whether  the  United  States  reaped  the 
many  anticipated  advantages  of  winning  that  race,  given  the  critical  national 
and  international  problems  plaguing  the  country  during  the  Vietnam  era, 
is  another  matter  altogether.  The  LOR  decision,  however,  had  other 
ramifications  for  the  U.S.  space  program;  it  meant  that  the  country  would 
skip  the  well-laid  plans  for  a  manned  space  station. 

269 


Space/light  Revolution 

Excited  NASA  researchers  had  been  studying  space  station  concepts 
seriously  for  at  least  four  years  when  NASA  chose  the  LOR  mode  for  Apollo; 
in  truth,  many  researchers  had  been  planning  for  a  space  station  from 
the  moment  of  NASA's  beginning  as  an  organization.  Although  the  LOR 
decision  did  not  stop  all  space  station  planning,  it  decisively  changed  space 
station  studies  by  de-emphasizing  the  immediate  importance  of  earth-orbital 
capabilities.  Moreover,  the  goal  of  landing  humans  on  the  moon  by  the  end 
of  the  decade  became  all-consuming,  and  researchers  who  did  space  station 
work  in  the  wake  of  the  mission-mode  decision  had  to  compete  with  Apollo 
for  support.  After  Apollo,  the  situation  did  not  improve;  space  station 
advocates  then  had  to  justify  a  return  to  the  development  of  something 
that  the  country  had  once  decided  it  did  not  need.  NASA  Marshall  engineer 
Jesco  von  Puttkamer  explains  this  predicament: 

After  the  close  down  of  Apollo,  we  began  to  pay  the  price.  We  are  trying  to  fill 
that  gap,  which  we  jumped  over,  and  are  having  a  tough  time  with  a  convincing 
justification  to  do  it.  Sometimes  I  wish  we  had  done  EOR.  Then  we  would  probably 
have  a  space  station  already.  Then  we  wouldn't  have  to  go  back  and  rejustify 
something  that  looks  to  many  people  like  a  step  backwards.  And  in  a  certain  sense 
it  is.  We've  been  to  the  moon  already,  history  knows,  and  now  all  of  a  sudden  we're 
trying  to  fill  this  empty  space. 

This  was  a  major  psychological  and  political  obstacle  for  the  champions  of 
any  space  station  concept  to  overcome.  It  explains  why  now,  on  the  eve  of 
the  twenty-first  century,  "the  next  logical  step"  in  space  exploration  after 
orbiting  a  human  has  not  yet  been  taken. 

In  the  mid-1970s,  the  United  States  did  launch  an  orbital  space  station, 
Skylab.  The  technology  for  this  station  was  a  direct  outgrowth  of  the 
Apollo  Extension  System,  a  spin-off  of  the  LOR-determined  Apollo  program. 
Skylab,  as  successful  as  it  proved  to  be,  was  not  the  versatile  and  long- 
lasting  station  that  NASA  had  planned  since  the  late  1950s.  Designed 
to  satisfy  the  institutional  need  to  do  something  after  Apollo  and  to  keep 
the  NASA  team  together  long  enough  to  finish  the  lunar  landing  missions, 
Skylab  was  makeshift  and  temporary.  NASA's  space  station  engineers, 
in  fact,  deliberately  built  the  station  without  the  thrusters  necessary  to 
keep  it  in  orbit  for  any  significant  amount  of  time.  By  limiting  Skylab's 
"lifetime,"  they  hoped  to  ensure  the  construction  of  a  more  permanent  and 
sophisticated  station — one  more  in  keeping  with  their  original  plans.  When 
Skylab  came  down,  they  would  replace  it  with  the  station  they  had  always 
wanted— that  was  the  idea.  In  1979,  Skylab  did  fall  to  earth  and  made 
more  news  as  a  burning  hunk  of  metal  than  it  ever  did  as  an  operating 
space  laboratory.  The  public  feared  that  falling  pieces  of  the  spacecraft 
might  destroy  homes  or  kill  children  at  play  in  school  yards.  Most  of  the 
orbital  workshop  landed  in  Western  Australia,  and  none  of  it  did  any  serious 
damage.  Although  Skylab  came  down,  by  the  1990s,  NASA  still  had  not 

270 


Skipping  "The  Next  Logical  Step" 

been  able  to  replace  it  as  hoped.  Once  skipped  over,  "the  next  logical  step" 
proved  increasingly  difficult  to  justify.2 


"As  Inevitable  as  the  Rising  Sun" 

In  imagining  how  humans  would  voyage  to  the  moon  and  the  planets, 
all  rocket  pioneers  envisioned  the  value  of  a  staging  base  in  earth  orbit. 
The  Russian  theoretician  Konstantin  Tsiolkovskii  recommended  such  an 
outpost  in  1911  in  his  pioneering  Investigation  of  Universal  Space  by  Means 
of  Reactive  Devices,  and  the  German  scientist  Hermann  Oberth  suggested 
likewise  in  his  1923  book  Die  Rakete  zu  den  Planetenraumen.  Austria's 
Guido  von  Pirquet  envisioned  the  use  of  an  earth-orbit  station  in  his  series 
of  provocative  articles  on  "Interplanetary  Travel  Routes"  appearing  in  Die 
Rakete  ( The  Rocket) ,  which  was  published  by  the  German  Society  for  Space 
Travel  in  1928  and  1929.  Despite  these  early  ideas  for  a  station,  rocket 
enthusiasts  did  not  seriously  consider  building  one  until  several  years  after 
the  end  of  World  War  II  and  the  appearance  of  the  first  practical  jet  and 
rocket  engines.3 

One  of  the  first  to  offer  a  station  design  was  the  master  designer  of 
the  V-2  rocket,  Wernher  von  Braun.  In  1952,  having  quickly  acclimated 
himself  to  the  American  scene  and  recognizing  the  need  to  make  spaceflight 
a  respectable  topic  for  public  discussion,  von  Braun  wrote  an  article  for  a 
special  issue  of  the  popular  American  magazine  Colliers.  This  issue  was 
devoted  to  the  idea  of  space  exploration.  Von  Braun  called  his  contribution 
"Crossing  the  Last  Frontier"  and  made  its  focus  the  imaginative  design  of 
a  manned  space  station  in  permanent  earth  orbit.4 

In  the  article  von  Braun  wrote,  "Development  of  the  space  station  is  as 
inevitable  as  the  rising  sun."  "Man  has  already  poked  his  nose  into  space" 
with  sounding  rockets,  and  "he  is  not  likely  to  pull  it  back."  "Within  the 
next  10  to  15  years,"  von  Braun  predicted,  "the  earth  will  have  a  new 
companion  in  the  skies."  An  "artificial  moon,"  an  earth-orbiting  base  "from 
which  a  trip  to  the  moon  itself  will  be  just  a  step,"  will  be  "carried  into  space, 
piece  by  piece,  by  rocket  ships."  From  there,  the  human  civilization  of  deep 
space  would  begin.5 

The  space  station  conceived  by  von  Braun  was  no  crude  affair;  it  was  an 
elaborate  and  beautiful  object,  a  huge  250- foot- wide  wheel.  The  enormous 
torus  rotated  slowly  as  it  orbited  the  earth  to  provide  artificial  or  "synthetic 
gravity"  for  pressurized  living  spaces  situated  about  the  wheel's  center. 
Writer  Arthur  C.  Clarke  and  moviemaker  Stanley  Kubrick  would  borrow 
the  torus  design  for  their  exhilarating  (and  baffling)  1968  movie  epic  2001: 
A  Space  Odyssey.  In  the  film,  the  space  wheel  turns  majestically  to  the 
waltz  of  Johann  Strauss's  "The  Blue  Danube,"  while  a  space  shuttle  vehicle 
with  passengers  aboard  leisurely  approaches  the  station. 


271 


Space/light  Revolution 

The  hub  of  von  Braun's  wheel  served  as  a  stationary  zero-gravity  mod- 
ule for  earth  and  space  observations  with  an  assembly  of  equipment  and  in- 
struments useful  for  a  host  of  scientific  and  applied  industrial  experiments. 
On-board  apparatuses  would  include  "powerful  telescopes  attached  to  large 
optical  screens,  radarscopes,  and  cameras  to  keep  under  constant  inspec- 
tion every  ocean,  continent,  country,  and  city."  At  short  distances  from  the 
station,  there  would  be  unmanned  stationary  platforms  for  remotely  con- 
trolled telescopic  observation  of  the  heavens.  While  helping  to  uncover  the 
secrets  of  the  universe,  the  space  station  would  also  work  to  disclose  the  evil 
ambitions  of  humankind.  With  its  telescopic  and  camera  eyes,  von  Braun 
claimed,  the  station  would  make  it  virtually  "impossible  for  any  nation  to 
hide  warlike  preparations  for  any  length  of  time."  Such  would  be  the  novel 
and  unprecedented  benefits  of  a  permanent  manned  base  in  earth  orbit.  Von 
Braun  predicted  that  the  station  would  become  a  reality  in  a  few  decades.6 

At  NACA  Langley  in  the  1950s,  the  prospects  of  an  orbiting  space  station 
did  not  pass  unnoticed.  Several  researchers  speculated  about  the  technology 
that  would  be  needed  someday  to  develop  an  operational  space  station 
such  as  the  one  von  Braun  had  described.  Suddenly,  in  1958,  interest 
in  a  space  station  exploded.  While  the  ink  was  still  drying  on  the  Space 
Act,  preliminary  working  groups  concerned  with  space  station  concepts  and 
technology  came  alive  both  within  NASA  and  around  the  aerospace  industry. 
NASA's  intercenter  Goett  Committee  was  one  of  these  early  groups. 

At  the  first  meeting  of  the  Goett  Committee  on  25-26  May  1959,  each 
member  addressed  the  group  for  10  to  15  minutes  to  propose  ideas  for  the 
next  manned  spaceflight  objective  after  Project  Mercury.  Of  all  the  speakers 
at  the  meeting,  no  one  sounded  more  enthusiastic  about  the  potential  of 
an  orbiting  space  station  than  Langley  representative  Larry  Loftin.  In  his 
presentation  for  what  he  called  Project  AMIS,  or  Advanced  Man  In  Space, 
Loftin  recommended  that  "NASA  undertake  research  directed  toward  the 
following  type  of  system:  a  permanent  space  station  with  a  'transport 
satellite'  capable  of  rendezvous  with  the  space  station."  According  to 
Loftin,  the  space  station  should  possess  the  following  general  characteristics: 
It  should  be  "large  enough  to  accommodate  two  or  more  persons  for  an 
extended  period  of  time";  it  should  be  "stabilized  and  oriented  in  some 
prescribed  manner" ;  it  should  be  "capable  of  changing  its  orientation,  and 
perhaps  its  orbit,  under  control  of  the  crew";  and  it  should  be  able  to  attach 
to  the  transport  satellite  for  supply  and  change  of  personnel.  In  addition, 
Loftin  argued  that  the  satellite  transport  or  "rendezvous  machine"  should 
be  able  to  alter,  "through  appropriate  guidance  and  control  systems,  its 
initial  orbit  so  as  to  rendezvous  with  the  space  station."  It  should  possess  a 
navigation  system  "which  will  ensure  that  that  pilot  can  find  and  intercept 
the  space  station."  The  transport  vehicle  should  be  able  to  dock  with  the 
station  "in  such  a  way  as  to  permit  transfer  of  payloads  between  vehicles," 
and,  importantly,  it  should  be  able  to  return  from  space  and  land,  under 
control  of  the  pilot,  "at  a  preselected  spot  on  the  earth."7 

272 


Skipping  "The  Next  Logical  Step" 


L-62-8400 

One  of  Langley's  early  concepts  for  a  manned  space  station:  a  self-inflating  75-foot- 
diameter  rotating  hexagon. 

For  emergency  use,  Loftin  explained,  the  station  could  be  outfitted  with 
a  "space  parachute,"  some  sort  of  "flexible,  kite-like"  package  that  would 
deploy  to  make  it  possible  for  the  station's  occupants  to  survive  atmospheric 
reentry.  Otherwise,  the  space  transport  would  make  all  trips  to  and  from 
space.  As  for  what  this  shuttle-like  vehicle  might  be,  Loftin  indicated 
that  the  air  force's  X-20  Dyna-Soar  manned  boost-glider  vehicle,  "could  be 
modified  to  perform  the  desired  function."  (The  X-20  would  not  be  built; 
Secretary  of  Defense  Robert  S.  McNamara  canceled  the  multimillion-dollar, 
six-year-old  program  in  1963.)  As  an  "initial  step"  to  test  the  transport 
concept,  Loftin  suggested  that  a  "proximity  rendezvous  of  Dyna  Soar"  with 
some  orbiting  satellite  might  be  undertaken.8 


273 


Space/light  Revolution 

In  his  talk  Loftin  emphasized  the  many  uses  of  the  space  station.  It  could 
serve  as  "a  medical  laboratory  for  the  study  of  man  and  his  ability  to  func- 
tion on  long  space  missions."  In  the  station,  researchers  could  study  "the 
effects  of  space  environment  on  materials,  equipment,  and  powerplants." 
NASA  could  use  the  station  to  develop  new  stabilization,  orientation,  and 
navigational  techniques,  as  well  as  to  learn  how  to  accomplish  rendezvous  in 
space.  With  telescopes  and  cameras  on  board,  the  station  could  also  serve 
as  an  orbiting  astronomical  observatory  and  as  an  "earth  survey  vehicle" 
for  meteorological,  geographical,  and  military  reconnaissance.  The  minutes 
of  the  Goett  Committee  do  not  record  the  immediate  reaction  to  Loftin's 
AMIS  presentation  specifically,  but  several  members  of  the  steering  com- 
mittee did  come  away  from  their  two-day  meeting  in  Washington  with  a 
strong  feeling  that  a  manned  space  station  should  be  the  "target  project" 
after  Mercury.9 

At  NASA's  First  Anniversary  Inspection  a  few  months  later,  Loftin  told 
a  large  audience  at  one  of  the  major  stops  along  the  tour  that  NASA's  long- 
range  objectives  included  "manned  exploration  of  the  moon  and  planets 
and  the  provision  of  manned  earth  satellites  for  purposes  of  terrestrial  and 
astronomical  observation,"  and  perhaps  even  for  military  surveillance.  But 
"the  next  major  step  [author's  emphasis]  in  the  direction  of  accomplishing 
these  long-range  objectives  of  manned  space  exploration  and  use  would 
appear  to  involve  the  establishment  of  a  manned  orbiting  space  laboratory 
capable  of  supporting  two  or  more  men  in  space  for  a  period  of  several 
weeks."  NASA  Langley,  Loftin  told  the  crowd,  was  now  focusing  its  research 
"with  a  view  toward  providing  the  technological  background  necessary  to 
support  the  development  of  a  manned  orbiting  laboratory." 10 

Interestingly,  in  his  original  typed  comments  for  the  inspection,  Loftin 
had  written:  "I  would  like  to  stress  that  we  at  Langley  do  not  intend  to 
develop,  build,  or  contract  for  the  construction  of  such  a  vehicle."  The 
center's  goal,  according  to  Loftin,  in  keeping  with  its  conservative  NACA 
policy  not  to  design  aircraft,  was  to  "seek  out  and  solve  the  problems  which 
lie  in  the  way  of  the  development  of  such  a  vehicle  system."  Loftin,  however, 
had  crossed  through  this  first  line.  Perhaps  he  realized  that  the  times  were 
changing;  Langley  could  "develop,  build,  or  contract"  for  NASA's  space 
station.11  This  was  NASA  not  the  NACA,  after  all. 


The  First  Space  Station  Task  Force 

When  Larry  Loftin  spoke  to  the  Goett  Committee,  he  had  already 
helped  Floyd  Thompson  organize  15  of  the  center's  brightest  researchers 
into  the  Manned  Space  Laboratory  Research  Group.  Thompson  had  made 
a  surprising  choice  for  the  chair  of  the  space  station  committee  in  veteran 
aeronautical  engineer  Mark  R.  Nichols,  the  longtime  head  of  the  Full-Scale 
Research  Division.  Nichols,  a  dedicated  airplane  man,  had  little  interest 

274 


Skipping  "The  Next  Logical  Step" 


L-62-4088  L-62-4065 

Two  key  members  of  Langley's  early  space  station  research  were  Paul  R.  Hill  (left) 
and  Robert  Osborne  (right). 

in  making  the  transition  to  space.  As  mentioned  in  chapter  4,  Thompson 
made  the  appointment  as  an  example  to  the  many  other  airplane  buffs  at 
the  center.  Langley  research  was  still  a  team  effort,  and  the  team  was 
now  moving  beyond  the  atmosphere.  Aeronautics  staff  members  would  be 
expected  to  become  involved  in  space  projects.  No  one  should  expect  a 
deferment — not  even  the  head  of  a  division.12 

The  Manned  Space  Laboratory  Research  Group  consisted  of  six  subcom- 
mittees responsible  for  the  study  of  various  essential  aspects  of  space  station 
design  and  operations:  (1)  Design  and  Uses  of  the  Space  Station,  led  by 
Paul  R.  Hill  of  PARD;  (2)  Stabilization  and  Orientation,  led  by  the  brilliant 
and  mild-mannered  head  of  the  Guidance  and  Control  Branch,  W.  Hewitt 
Phillips;  (3)  Life  Support,  headed  by  A.  Wythe  Sinclair,  Jr.,  of  the  new 
Theoretical  Mechanics  Division;  (4)  Rendezvous  Analysis,  led  by  Houbolt, 
then  the  assistant  chief  of  the  Dynamic  Loads  Division;  (5)  Rendezvous  Ve- 
hicle, led  by  Eugene  S.  Love,  who  was  an  extremely  talented  hypersonics 
specialist  working  in  the  Aero-Physics  Division;  and  (6)  Power  Plant,  led 
by  Nichols  himself.*  According  to  handwritten  comments  by  Thompson 
on  the  rough  organization  chart  sketched  by  Loftin,  the  objective  of  the 
space  station  committee  was  to  "develop  technology  and  make  pitch  for  do- 
ing it."  The  goal  was  to  demonstrate  that  "this  is  possible  and  this  is  the 
way  we  can  do  it."  As  for  how  to  organize  and  manage  the  work  of  the 


Due  to  Nichols'  ambivalence  about  the  space  project,  Paul  R.  Hill  actually  took  over  much  of  the 


leadership  role  for  the  group. 


275 


Space/light  Revolution 

committee,  Thompson  said  only  to  "organize  like  WS  110,"  that  is,  similar 
to  the  support  of  the  development  of  Weapons  System  110,  the  air  force's 
experimental  B-70  strategic  bomber.  The  organization  would  be  informal 
so  that  it  could  cut  across  formal  divisional  lines,  but  its  work  would  receive 
the  highest  priority  in  the  shops.13 

Thompson  made  one  other  revealing  note  at  the  bottom  of  the  commit- 
tee's organization  chart:  "Plan  whole  organization  of  getting  man  to  moon." 
This  footnote  implies  that  in  Thompson's  mind  the  clear  and  accepted  ob- 
jective of  NASA's  manned  space  effort  following  Project  Mercury  was  to 
send  an  astronaut  to  the  moon  and  back.  The  way  to  achieve  that  objective 
was,  as  all  the  visionaries  of  space  exploration  had  articulated,  by  moving 
out  from  an  orbiting  relay  station.  Langley's  associate  director  was  asking 
his  in-house  committee  to  study  the  entire  enterprise  involved  not  only  in 
building  and  operating  a  space  station  but  also  in  using  it  as  a  launchpad 
for  the  eventual  manned  lunar  landing  mission.14 

Not  everyone  in  NASA  thought  that  the  space  station  should  be  the  tar- 
get project.  Dr.  Adolf  Busemann,  the  German  pioneer  of  the  swept  wing 
who  came  to  Langley  in  1947,  argued  that  the  space  environment  would  offer 
experimenters  no  vital  scientific  or  technological  knowledge  that  researchers 
with  some  ingenuity  could  not  acquire  on  earth.  But  with  the  exception  of 
Busemann  and  the  small  group  of  lunar  landing  advocates  mostly  clustered 
around  Clint  Brown  and  the  Theoretical  Mechanics  Division,  nearly  every- 
one else  at  Langley  in  the  summer  of  1959,  including  senior  management, 
threw  their  weight  behind  the  space  station.  Members  of  Nichols'  group 
immersed  themselves  in  a  centerwide  effort  to  define  and  answer  a  host 
of  major  questions  related  to  placing  and  operating  a  manned  laboratory 
in  earth  orbit.  Inquiries  and  suggestions  were  pouring  into  Langley  from 
the  aerospace  industry,  notably  from  the  Goodyear  Aircraft  Corporation, 
Chance  Vought  Astronautics,  and  the  Martin  Company,  whose  representa- 
tives had  heard  what  NASA  Langley  was  up  to  and  wanted  to  participate 
in  the  development  of  the  manned  station.15 

By  the  fall  of  1959,  the  work  of  the  Nichols  committee  had  progressed  to 
the  point  where  Loftin  could  make  a  simple  three-point  statement  of  pur- 
pose. Langley  would  (1)  "study  the  psychological  and  physiological  reactions 
of  man  in  a  space  environment  over  extended  periods  of  time,"  thereby  de- 
termining "the  capabilities  and  limitations  of  man  in  performing  long  space 
missions" ;  (2)  "provide  a  means  for  studying  materials,  structures,  control 
and  orientation  systems,  auxiliary  powerplants,  etc.,  in  a  true  space  environ- 
ment"; and  (3)  "study  means  of  communication,  orbit  control,  rendezvous," 
as  well  as  techniques  for  earth  and  astronomical  observations.16  In  sum- 
mary, Loftin  told  the  committee  that  Langley  was  primed  and  ready  to  take 
on  the  role  of  the  lead  center  in  all  NASA's  space  station  work — quite  an 
ambitious  undertaking  for  the  former  NACA  aeronautics  laboratory. 


276 


Skipping  "The  Next  Logical  Step" 
From  the  Inflatable  Torus  to  the  Rotating  Hexagon 

If  Langley  researchers  favored  any  particular  kind  of  space  station  as 
they  set  out  to  examine  the  feasibility  of  various  configurations  in  1960, 
their  preference  was  definitely  a  self- deploying  inflatable.  The  Langley 
space  station  office  had  eliminated  one-by-one  the  concepts  for  noninflatable 
configurations,  some  of  which  came  from  industry.  Notions  for  a  simple 
orbiting  "can,"  or  cylinder,  and  for  a  cylinder  attached  to  a  terminal  stage 
of  a  booster  were  rejected  as  dynamically  unstable;  they  had  a  tendency 
to  roll  at  the  slightest  disturbance.  A  version  of  Lockheed's  sophisticated 
elongated  modular  concept  was  turned  down  because  it  was  too  futuristic 
and  required  the  launch  of  several  boosters  to  place  all  the  components  into 
earth  orbit.  Proposals  for  hub-and-spoke  designs,  big  orbiting  Ferris  wheels 
in  space,  were  turned  down  because  of  the  Coriolis  effects.  Disturbances  of 
the  inner  ear,  such  as  nausea,  vertigo,  and  dizziness,  would  debilitate  crew 
members  moving  radially  in  any  system  that  was  rotating  too  rapidly. 

Langley's  space  station  team  had  sound  technical  reasons  for  doubting 
the  feasibility  of  these  proposals.  However,  the  team  possessed  a  strong 
institutional  bias  for  an  inflatable  station.  After  all,  the  inflatable  was 
developed  at  Langley.  The  concept  also  made  good  engineering  sense. 
Hundreds  of  pounds  of  propellant  were  required  to  put  one  pound  of  payload 
into  orbit.  Any  plan  that  involved  lightening  the  payload  meant  simplifying 
rocket  requirements.  Because  of  their  experience  with  the  Echo  balloon, 
Langley  engineers  also  knew  firsthand  that  a  folded  station  packed  snugly 
inside  a  rocket  would  be  protected  during  the  rough  ride  through  the 
atmosphere. 

The  first  idea  for  an  inflatable  station  was  the  Erectable  Torus  Manned 
Space  Laboratory.  A  Langley  space  station  team  led  by  Paul  Hill  and 
Emanuel  "Manny"  Schnitzer  developed  the  concept  with  the  help  of  the 
Goodyear  Aircraft  Corporation.  Their  idea  called  for  a  flat  inflatable  ring 
or  torus  24  feet  in  diameter,  or  about  one-quarter  the  size  of  the  Echo  1 
sphere.17 

The  inflatable  torus  had  several  major  selling  points.  It  was  "unitized," 
meaning  that  all  its  elements  were  part  of  a  single  structure  that  could  be 
carried  to  orbit  by  the  launch  of  one  booster,  just  as  was  the  case  with  the 
Echo  balloon.  NASA  would  simply  fold  the  station  into  a  compact  payload 
for  an  automatic  deployment  once  the  payload  had  reached  altitude.  The 
inner  volume  of  the  torus  could  be  given  a  gravity  capability  of  0  to  1  G. 
The  station  could  be  designed  for  both  natural  and  artificial  stability,  for 
rendezvous-dock-abort  capability,  and  for  variable-demand  power  supply. 
The  torus  could  also  have  regenerative  life-support  systems  for  a  six-person 
crew.  To  provide  their  space  station  with  electric  power,  Hill  and  Schnitzer 
pursued  the  possibility  of  using  a  solar  turboelectric  system,  which  used  an 
innovative  umbrella-like  solar  collector  then  under  development  by  TRW 
as  part  of  the  NASA-supported  Sunflower  Auxiliary  Power  Plant  Project. 

277 


Space/light  Revolution 


L-63-3879 

Langley  researcher  Rene  Berglund  (left)  used  this  figure  (right)  in  1962  to  illustrate 
some  of  the  earliest  space  station  configurations  investigated  at  the  center:  (a)  a 
large  cylinder,  (b)  a  smaller  cylinder  attached  to  a  terminal-stage  booster,  (c)  a 
boom  with  multiple  docking  ports  powered  by  a  nuclear  power  plant  at  one  end, 
(d)  a  spoke  configuration,  (e)  a  modified  spoke  configuration  with  vertical  rather 
than  horizontal  modules,  and  (f)  a  wheel  or  torus. 


By  April  1960,  Schnitzer  was  so  enthusiastic  about  the  inflatable  torus 
that  he  made  a  formal  presentation  on  the  design  to  a  national  meeting 
of  the  American  Rocket  Society.  A  revised  and  updated  version  of  his  talk 
appeared  as  the  feature  article  in  the  January  1961  issue  of  Astronautics 
Magazine.  (In  late  1962,  Schnitzer  moved  to  the  Office  of  Manned  Space 
Flight  at  NASA  headquarters,  where  he  would  remain  active  in  space  station 
R&D  and  promote  Langley's  work  in  the  field.)18 

In  the  months  following  Schnitzer's  presentation,  Langley  built  and  tested 
various  models  of  the  Erectable  Torus  Manned  Space  Laboratory,  including 
a  full-scale  research  model  constructed  by  Goodyear.  But  researchers  began 
to  suspect  that  the  design  was  lacking  in  certain  key  respects.  The  principal 
concern  was  the  same  one  that  had  plagued  the  promoters  of  Echo:  the 
danger  of  a  meteorite  puncturing  the  structure.  Goodyear  built  the  research 


278 


Skipping  "The  Next  Logical  Step" 


L-61-8027  L-61-8029 

Testing  indicated  that  the  inflatable  torus  could  be  packaged  around  the  hub  so  that 
it  occupied  only  2  percent  of  its  inflated  volume. 


L-61-8693 

Looking  like  a  huge  pneumatic  tire  sitting  on  a  giant  car  jack,  Langley  's  full-size  test 
model  of  its  24-foot  toroidal  space  station  receives  a  visit  from  NASA  Administrator 
James  Webb  in  December  1961.  Escorting  Webb  are  Floyd  Thompson  (far  left)  and 
T.  Melvin  Butler,  Langley 's  assistant  director  for  administration. 

279 


Spaceflight  Revolution 


L-62-312 

Langley  engineers  check  out  the  interior  of  the  inflatable  24- foot  space  station  in 
January  1962. 

model  out  of  a  lightweight  three-ply  nylon  cord  held  together  firmly  by 
a  sticky  rubber-like  material  known  as  butyl  elastomer.  Such  a  large 
rubberized  surface  would  certainly  be  vulnerable  during  a  meteoroid  shower. 
This  concern  proved  much  harder  to  dismiss  for  a  manned  station  than  for 
the  unpiloted  satellite.  In  addition,  while  the  condition  of  being  "dead  soft" 
was  seen  as  an  advantage  for  the  Echo  balloon,  it  was  a  disadvantage  for 
a  busy  manned  space  station.  Some  engineers  worried  that  if  the  flexible 
material  was  not  strong  enough,  crew  members  moving  around  vigorously 
in  the  space  station  might  somehow  propel  themselves  so  forcefully  from  one 
side  of  the  station  to  the  other  that  they  would  break  through  a  wall  and 
go  shooting  into  outer  space. 

A  more  serious  engineering  concern  arose  that  was  related  to  the  dynam- 
ics of  the  toroidal  structure.  When  arriving  crew  members  moved  equipment 
from  the  central  hub  to  a  working  area  at  the  outside  periphery  of  the  ring, 
or  when  a  ferry  vehicle  simply  impacted  with  the  station's  docking  port, 
Langley  researchers  believed  that  the  station  might  become  slightly  unsta- 
ble, thus  upsetting  its  precisely  established  orbit.  Less  strenuous  activities 
might  also  disturb  the  fragile  dynamics  of  the  torus.  Knowing  that  the  hu- 
man occupants  of  the  station  would  have  no  weight  but  would  still  have 
mass,  the  Langley  space  station  group  conducted  analytical  studies  using 

280 


Skipping  "The  Next  Logical  Step" 

analog  computers  to  calculate  the  effect  of  astronauts  moving  about  in  the 
station.  The  results  showed  that  the  mass  distribution  would  be  changed 
when  crew  members  just  walked  from  one  part  of  the  vehicle  to  another. 
This  change  produced  a  slight  oscillation,  or  what  the  researchers  called  a 
"wobble,"  of  the  entire  station. 

To  discover  whether  they  could  alleviate  this  wobble,  the  Langley  space 
station  group  decided  to  build  a  10- foot-diameter  elastically  scaled  model  of 
the  torus.  This  model  did  not  become  operational  until  the  summer  of  1961, 
however,  and  by  that  time  NASA  had  realized  that  it  must  either  develop 
a  more  rigid  inflatable  or  abandon  the  idea  of  an  inflatable  altogether.19 

While  still  in  pursuit  of  the  best  possible  inflatable  torus,  the  NASA 
Langley  space  station  group  did  explore  other  ideas.  Most  notably,  in  the 
summer  of  1961  it  entered  into  a  six-month  contract  with  North  American 
Aviation  for  a  detailed  feasibility  study  of  an  advanced  space  station 
concept.*  Developed  by  Langley  engineer  Rene  A.  Berglund,  the  design 
called  for  a  large  modular  manned  space  station,  which  although  essentially 
rigid  in  structure,  could  still  be  automatically  erected  in  space.  In  essence, 
Berglund 's  idea  was  to  put  together  a  series  of  six  rigid  modules  that  were 
connected  by  inflatable  spokes  or  passageways  to  a  central  nonrotating  hub. 
The  75-foot-diameter  structure  (initially  planners  thought  it  might  be  as 
large  as  100  feet)  would  be  assembled  entirely  on  the  ground,  packaged  into 
a  small  launch  configuration,  and  boosted  into  space  atop  a  Saturn  rocket. 
One  of  Berglund 's  prerequisites  for  the  design  was  that  it  provide  protection 
against  micrometeorites.  To  accomplish  this,  he  gave  the  rigid  sections  of 
the  rotating  hexagon  air-lock  doors  that  could  be  sealed  when  any  threat 
arose  to  the  integrity  of  the  interconnecting  inflatable  sections.20 

This  sophisticated  modular  assembly  was  to  rotate  slowly  in  space,  thus 
making  it  possible  for  its  occupants  to  enjoy  the  benefits  of  artificial  gravity, 
which  virtually  all  space  station  designers  at  the  time  believed  was  absolutely 
necessary  for  any  long-term  stay  in  space.  In  fact,  the  diameter  of  75 
feet  was  selected  because  it  provided  the  minimal  rotational  radius  needed 
to  generate  at  low  rotational  velocities  the  1  G  desired  for  the  station's 
living  areas.  Rotation  was  the  only  mechanism  known  at  the  time  for 
artificially  creating  gravity  conditions.  The  only  part  of  the  structure  that 
would  not  rotate  was  the  central  hub;  suspended  by  bearings,  the  hub 
would  turn  mechanically  in  the  opposite  direction  of  the  hexagon  at  just 
the  right  rate  to  cancel  all  the  effects  of  the  rotation.  Located  in  this 
nonrotating  center  of  the  space  station  would  be  a  laboratory  for  various 
experiments,  including  comparative  studies  of  the  effects  of  zero  and  artificial 
gravity.  The  nonrotating  hub  would  also  contain  the  dock  for  the  ferry 
vehicle.  Preliminary  experience  with  Langley's  earliest  rendezvous  and 
docking  simulators  indicated  that  a  trained  pilot  could  execute  a  docking 


North   American   had  been  studying  the  logistics  of  a  permanent   satellite   base  and   a  global 
surveillance  system  for  the  air  force,  and  the  physics  of  meteoroid  impact  for  NASA. 

281 


Space/light  Revolution 


L-62-3819 


With  a  10- foot- diameter  scale  model 
(above),  Langley  researchers  studied  the 
attitude  errors,  wobbling  motions,  and 
other  dynamic  characteristics  of  a  space 
station  spinning  in  space.  The  effects  of 
crew  motion  and  cargo  transfer  within 
the  station  were  simulated  by  an  electri- 
cally driven  mass  moving  around  a  track 
on  the  torus.  To  the  right,  a  Langley 
engineer  takes  a  walk  in  simulated  zero 
gravity  around  a  mock-up  of  a  full-scale, 
24- foot- diameter  space  station. 


L-64-10,099 


282 


Skipping  "The  Next  Logical  Step" 

maneuver  with  surprising  ease  as  long  as  the  station  docking  hub  was  fixed. 
If  the  hub  rotated  along  with  a  rotating  station,  the  maneuvering  operations 
would  have  to  be  much  more  complicated. 

As  engineers  from  North  American  and  Langley  probed  more  deeply  into 
the  possibilities  of  a  rotating  hexagon,  they  became  increasingly  confident 
that  they  were  on  the  right  track.  The  condition  that  the  station  be  self- 
deploying  or  self-erecting  (implying  some  means  of  mechanical  erection  or  a 
combination  of  mechanical  erection  with  inflation)  was  not  negotiable,  given 
the  economic  and  technological  benefits  of  being  able  to  deliver  the  space 
station  to  its  orbit  via  a  single  booster.  Early  on,  the  space  station  group 
talked  with  their  fellow  engineers  in  the  Scout  Project  Office  at  Langley 
about  using  a  Scout  booster  to  launch  the  station,  but  Scout  did  not  appear 
to  be  powerful  enough  to  carry  all  171,000  tons  of  the  rotating  hexagon 
to  orbit  altitude.  The  group  also  looked  into  using  a  Centaur,  a  liquid- 
fuel  booster  for  which  NASA  had  taken  over  the  responsibility  from  the 
DOD.  The  Centaur  promised  higher  thrust  and  bigger  pay  loads  for  lunar 
and  planetary  missions;  however,  Langley  learned  in  early  1961  that  the 
Centaur  was  "out  of  the  question"  because  "nothing  in  the  [high  priority] 
NASA  manned  space  programs  calls  for  it."  Furthermore,  the  Centaur  was 
not  yet  "man-rated,"  that  is,  approved  for  flights  with  astronauts  aboard, 
and  a  man  rating  was  "neither  expected  nor  anticipated."  Centaur  would 
prove  to  be  a  troublesome  launch  vehicle  even  for  its  specified  unmanned 
missions,  and  the  rocket  never  would  be  authorized  to  fly  humans.21 

Soon  space  station  advocates  turned  to  von  Braun's  Saturn.  With  its 
210,000-pound  payload  capacity,  an  advanced  Saturn  could  easily  lift  the 
171,000-pound  hexagon  into  orbit.  A  team  of  Langley  researchers  led  by 
Berglund  did  what  they  could  to  mate  their  space  station  to  the  top  stage 
of  a  Saturn.  Working  with  a  dynamic  scale  model,  they  refined  the  system 
of  mechanical  hinges  that  enabled  the  six  interconnected  modules  of  the 
hexagon  to  fold  into  one  compact  mass.  As  a  bonus,  the  hinges  also 
eliminated  the  need  for  fabric  connections  between  modules,  which  were 
more  vulnerable  to  damage.  Tests  demonstrated  that  the  arrangement  could 
be  carried  aloft  in  one  piece  with  the  three  retractable  spokes  stowed  safely 
inside  the  cavity  of  the  assimilated  module  cluster.  Once  orbit  was  achieved, 
a  series  of  screw-jack  actuators  located  at  the  joints  between  the  modules 
would  kick  in  to  deploy  the  folded  structure.  The  Langley  researchers  also 
made  sure  that  the  nonrotating  central  hub  of  their  hexagon  would  have  a 
port  that  could  accommodate  ferry  vehicles.  Such  vehicles  were  then  being 
proposed  for  the  Apollo  circumlunar  mission  and,  later,  for  a  lunar  landing 
via  EOR. 

The  estimated  cost  for  the  entire  space  station  project,  for  either  the 
erectable  torus  or  the  rotating  hexagon,  was  $100  million,  a  tidy  sum  upon 
which  Langley  and  NASA  headquarters  agreed.  This  figure  amounted  to 
the  lowest  cost  proposal  for  a  space  station  submitted  to  the  air  force  at 
its  space  station  conference  in  early  1961.  But,  as  George  Low  pointed  out 

283 


Space/light  Revolution 


HEXAGONAL    CONFIGURATION 


SOLAR -CELL   PANELS 


63  FT 


LIVING 
MODULE 


1 03  FT 


DOCKING 
FACILITY 


APOLLO -TYPE 
FERRY   VEHICLE 


FT- 


L-62-8730  L-62-8732 

North  American  selected  this  space  station  design  in  1962  for  final  systems  analysis 
(diagram  shown  at  top,  models  at  bottom,  left  and  right).  Incorporating  all  the 
advantages  of  a  wheel  configuration,  it  had  rigid  cylindrical  modules  arranged  in  a 
hexagonal  shape  with  three  rigid  telescoping  spokes.  This  configuration  eliminated 
the  need  for  exposed  flexible  fabric. 


284 


Skipping  "The  Next  Logical  Step" 

at  a  space  station  meeting  held  at  Langley  on  18  April  1961,  NASA  did 
not  have  the  money  for  a  space  station  follow-on  to  Project  Mercury;  what 
funds  NASA  expected  were  "only  enough  to  finish  Mercury  and  $29  million 
for  Apollo."22 

For  five  more  weeks,  until  President  Kennedy's  speech  on  25  May,  Apollo 
entailed  only  a  circumlunar  mission,  with  the  possibility  of  building  a  space 
station  as  a  by-product  of  the  earth-orbital  phase;  however,  as  George  Low 
observed,  NASA  had  not  guaranteed  that  such  a  phase  would  be  part  of 
Apollo.  Low  warned  the  assembled  space  station  advocates  that  the  chances 
were  high  that  Apollo  would  not  require  a  space  station  with  artificial 
gravity.  If  that  were  the  case,  NASA  would  have  neither  the  mandate  nor 
the  money  to  build  a  space  station  of  any  kind  for  some  time  to  come. 

Such  uncertainty  put  Langley  in  a  difficult  but  not  unfamiliar  situation. 
Some  sort  of  space  station  was  possible  for  the  Apollo  era,  and  as  long  as  that 
possibility  existed,  the  basic  technology  needed  for  a  station  had  to  be  ready, 
perhaps  at  short  notice.  To  assure  that  Langley  would  be  technologically 
prepared,  exploratory  research  had  to  be  ongoing. 

Larry  Loftin  made  this  point  clear  on  19  May  1961,  six  days  before 
President  Kennedy's  lunar  landing  speech,  in  his  testimony  to  the  U.S. 
House  Committee  on  Science  and  Astronautics,  chaired  by  Overton  Brooks 
(Democrat  from  Louisiana).  "We  have  not  been  developing  a  manned 
vehicle,"  Loftin  reassured  the  congressmen  and  their  staffs.  "We  have  been 
studying  what  we  would  consider  to  be  salient  or  pertinent  problems  which 
would  have  to  be  solved"  if  the  country  decided  that  a  station  was  needed. 
Loftin  described  in  some  detail  Langley's  manned  space  station  work.  "In 
order  to  try  to  fix  what  the  problem  areas  were,"  he  explained,  "it  was 
necessary  to  arrive  at  some  sort  of  a  concept  of  what  the  vehicle  might 
look  like."  He  then  passed  around  a  series  of  pictures  showing  Langley's 
concepts  for  both  the  inflatable  torus  and  the  rotating  hexagon,  expressing 
no  preference  for  either  design.*  After  reviewing  the  general  characteristics 
of  both  designs,  Loftin  summarized  Langley's  assessment  of  the  status  of 
the  space  station: 

So  far  as  we  know,  so  far  as  we  have  gone  at  the  present  time,  we  don't  see  that  there 
are  required  any  fundamental  scientific  breakthroughs  ...  in  order  to  design  one  of 
these  things.  However,  we  have  not  undertaken  at  the  Langley  Research  Center  a 
detailed  engineering  design  study.  If  such  a  study  were  undertaken,  you  might  run 
into  some  problems  that  we  haven't  been  smart  enough  to  think  about  that  are 
fundamental.  I  don't  know  if  you  would,  but  you  could. 

Moreover,  Loftin  concluded  his  testimony  with  a  caution,  "In  such  a  careful 
engineering  design,  this  is  a  long-term  proposition.  We  are  not  really  sure 


Two  representatives  of  the  Goodyear  Aircraft  Corporation,  the  primary  contractor  involved  in 
Langley's  study  of  the  inflatable  torus,  were  testifying  the  same  day  before  the  congressional  committee. 

285 


Spaceflight  Revolution 

when  you  got  all  done  whether  you  would  have  something  you  really  want 
or  not."23 

Whether  the  politicians  understood  Loftin's  essential  point  is  uncertain, 
for  they  had  a  difficult  time  even  fathoming  what  a  manned  space  station  was 
all  about  and  how  someday  it  might  be  used.  Chairman  Overton  Brooks,  for 
example,  asked,  "You  are  just  going  to  allow  that  [thing]  to  float  around  in 
space?"  When  asked  by  Minnesota  congressman  Joseph  E.  Karth  what  the 
"primary  function  of  this  so-called  space  station"  would  be,  Loftin  answered, 
"It  could  have  many  functions.  We  are  not  really  proposing  a  space  station. 
What  we  are  doing  here  is  saying  if  you  want  one,  we  would  like  to  look  into 
the  problems  of  how  you  might  make  it."  Encouraged  to  say  what  those 
functions  were,  Loftin  explained  how  the  experience  of  having  humans  in 
an  orbiting  space  station  would  be  helpful  and  perhaps  even  necessary  in 
preparing  for  long-distance  space  flights,  perhaps  even  for  the  two-week 
trip  from  the  earth  to  the  moon  and  back  that  the  United  States  was  now 
planning.  Certainly,  if  the  United  States  was  to  attempt  any  flights  to  places 
more  distant  than  the  moon,  Loftin  explained,  "it  would  be  desirable  to  have 
a  space  station  in  orbit  where  we  could  put  men,  materials,  different  kinds 
of  mechanisms.  We  could  put  them  up  there  for  weeks  at  a  time  and  see  if 
there  are  any  undesirable  effects  that  we  have  not  foreseen.  If  these  effects 
crop  up,  then  you  bring  the  man  back."  An  astronaut  already  on  course  to 
a  distant  planet  was  not  so  easily  retrieved.24 


Betwixt  and  Between 

Six  days  after  Loftin's  appearance  before  the  congressional  committee, 
President  Kennedy  stunned  NASA  with  his  lunar  landing  speech.  Apollo 
was  no  longer  a  manned  circumlunar  mission;  it  was  now  the  project 
for  landing  a  man  on  the  moon  by  1969.  In  one  extraordinary  political 
moment,  step  three  of  the  space  program  had  become  step  one.  For  14 
months  following  Kennedy's  speech,  NASA  debated  the  advantages  and 
disadvantages  of  various  mission  modes.  For  at  least  the  first  half  of  this 
period,  many  in  NASA  were  quite  sure  that  the  country  would  be  going 
to  the  moon  via  EOR.  In  this  mode,  the  lunar  spacecraft  would  actually 
be  assembled  from  components  put  into  orbit  by  two  or  more  Saturn 
launch  vehicles.  This  EOR  plan  would  therefore  involve  the  development 
of  certain  orbital  capabilities  and  hardware  that  might  easily  translate  into 
the  country's  first  space  station. 

With  this  possibility  in  mind,  Langley's  space  station  team  worked 
through  the  remainder  of  1961  and  into  1962  to  identify  and  explore  the 
essential  problems  facing  the  design  and  operation  of  a  space  station. 
The  thrust  of  the  center's  research  during  this  period  of  political  and 
institutional  limbo  for  the  space  station  was  divided  among  three  major 
areas:  (1)  dynamics  and  control,  or  how  to  control  a  rotating  structure  in 

286 


Skipping  "The  Next  Logical  Step" 

orbit;  (2)  on-board  power,  or  how  to  provide  electrical  power  as  well  as  store 
and  use  energy  in  the  space  station;  and  (3)  life  support,  or  how  to  ensure 
that  the  occupants  of  the  station  could  best  remain  healthy  and  vigorous 
during  (and  after)  long  sojourns  in  space. 

From  the  start,  almost  all  space  station  designers  presumed  the  need 
for  artificial  gravity.  From  this  presumption  came  the  notion  of  a  rotating 
structure,  be  it  a  rotating  cylinder,  torus,  or  hexagon,  or  of  a  centrifuge 
mechanism  within  a  nonrotating  structure  that  could  provide  a  force  that 
substituted  for  the  lack  of  gravity.  Whether  it  was  absolutely  necessary 
to  substitute  centrifugal  force  for  the  effects  of  gravity,  no  one  really  knew. 
Perhaps  a  human  in  space  would  need  1  G;  perhaps  as  little  as  0.25  G  would 
do.  One  thing  the  space  station  researchers  did  know  with  some  certainty 
was  that  they  needed  to  be  careful  about  this  matter  of  gravitational  effects. 
If  the  rotational  radius  was  too  small,  or  the  structure  rotated  at  too  high 
an  rpm,  the  astronauts  inside  would  suddenly  become  ill. 

The  rotation  had  to  be  controlled  precisely,  whatever  the  station's 
configuration.  Thus,  one  set  of  problems  that  Langley  researchers  attempted 
to  solve  concerned  a  spinning  space  station's  inherent  vulnerability  to 
disturbances  in  dynamic  stability;  this  included  compensating  for  the  wobble 
motions  that  might  occur  when  crew  members  moved  about  inside  the 
station  or  when  ferry  vehicles  pushed  up  against  the  outside  structure  during 
docking. 

The  Langley  space  station  group  found  a  solution  for  attitude  control 
using  a  system  of  four  pulse  jets.*  These  small  pulse  jets  could  be  mounted 
at  90-degree  intervals  around  the  outside  rim  of  the  station  to  reorient  the 
station  when  necessary.  Then,  to  dampen  the  wobbles  caused  by  crew 
movements  and  other  disturbances  in  mass  distribution,  Langley  found  that 
a  spinning  flywheel  could  be  rotated  to  produce  the  necessary  countervailing 
torques;  the  same  flywheel  could  produce  the  torque  required  to  keep  the 
station  rotating  around  its  predetermined  axis.  If  the  flywheel  failed  to 
steady  the  wobbles,  the  pulse  jets  could  be  fired  as  a  backup.  In  late 
1961  and  early  1962,  Langley  researchers  subjected  full-scale  models  of 
both  the  rotating  hexagon  and  the  inflatable  torus  (the  torus  was  then  still 
being  considered)  to  systematic  tests  involving  these  experimental  control 
mechanisms.25 

Langley  researchers  found  little  reason  to  disagree  about  what  was  needed 
for  the  dynamic  control  of  the  space  station;  however,  bitter  arguments 
arose  over  the  power  source  for  the  station.  Two  main  energy  sources  were 
considered:  solar  and  nuclear.  (A  third  possibility,  involving  the  use  of 
chemical  energy  from  a  regenerative  fuel  cell,  was  considered  briefly  but  was 
summarily  dismissed  as  "futuristic"  and  "unfeasible.")  To  many  at  Langley, 


A  pulse  jet  is  a  simple  jet  engine,  which  does  not  involve  a  compressor,  in  which  combustion  takes 
place  intermittently  and  produces  thrust.  In  this  case,  10  pounds  of  thrust  per  pulse  jet  is  produced  by 
a  series  of  explosions. 

287 


Spaceflight  Revolution 

the  obvious  choice  was  solar.  The  sun's  energy  was  abundant  and  available. 
If  solar  power  was  used,  the  space  station  would  not  have  to  carry  the  weight 
of  its  own  fuel  into  orbit;  photovoltaic  or  solar  cells  (which  existed  in  1960 
but  not  in  a  very  advanced  form)  would  simply  convert  the  sunlight  into  the 
electrical  energy  needed  to  run  the  space  station. 

Outspoken  critics,  however,  argued  that  the  technology  did  not  yet  exist 
for  a  solar-powered  system  that  could  sustain  a  spacecraft  over  long  missions. 
With  the  rotation  necessary  for  artificial  gravity,  situating  and  realigning  the 
solar  panels  so  that  they  would  always  be  facing  the  sun  became  problematic. 
Depending  on  the  station's  configuration,  some  solar  panels  would  be  shaded 
from  the  sun  most  of  the  time.  Solar  panels,  especially  large  ones,  would 
also  have  an  undesirable  orbital  drag  effect. 

Some  argued  that  the  better  choice  was  nuclear  power.  The  problems 
of  shielding  living  areas  from  the  reactor's  radiation  and  radioactive  waste 
would  have  to  be  solved,  of  course,  because  humans  would  be  on  board, 
but  once  these  problems  were  resolved,  a  small  nuclear  reactor  could  safely 
supply  enough  power  (10  to  50  kilowatts)  to  sustain  the  operation  of  a  station 
for  a  year  or  more.  Yet  engineers  were  not  able  to  overcome  the  major 
logistical  and  safety  problems  of  the  proposed  reactor  systems.  Particularly 
bothersome  was  the  problem  of  replacing  an  operational  reactor  should 
it  fail.  Even  the  shielding  problem  proved  more  difficult  to  handle  than 
imagined.  In  later  space  station  designs,  engineers  tried  to  bypass  the 
shielding  problem  by  employing  a  large  shadow  shield  and  a  long  boom 
to  separate  the  reactor  from  the  habitation  areas,  but  the  boom  required 
such  a  major  reconfiguration  of  the  proposed  space  station  structure  that 
the  idea  had  to  be  abandoned. 

Some  researchers  rejected  both  solar  and  conventional  nuclear  systems 
and  advocated  a  radioisotope  system.  In  this  arrangement  a  radioactive 
element  such  as  uranium  238  or  polonium  210  would  emit  energy  over  a  long 
period  and  at  a  specific  and  known  rate.  This  power  system  was  based  on 
the  so-called  Brayton  cycle  (also  called  the  "Joule  cycle" ) ,  which  was  a  well- 
known  thermodynamic  cycle  named  after  American  engineer  and  inventor 
George  B.  Brayton  (b.  1873).  The  Brayton  cycle  consisted  of  an  isentropic 
compression  of  a  working  substance,  in  this  case  a  radioactive  isotope, 
the  addition  of  heat  at  a  constant  pressure,  an  isentropic  expansion  to  an 
ambient  pressure,  and,  finally,  the  production  of  an  exhaust.  Such  a  system 
required  minimum  shielding  and  did  not  require  booms  or  large  panels. 
The  availability  of  high-quality  waste  heat  could  also  be  used  in  thermal 
control  and  in  the  life-support  system,  thereby  reducing  the  overall  power 
system  requirements.  On  the  other  hand,  the  isotope  Brayton  cycle  power 
system  did  require  internal  rotating  machinery  that  still  needed  considerable 
development.  It  would  also  require  an  increased  radiator  area  as  well  as 
doors  on  the  skirt  of  the  radiator  that  could  open  to  allow  the  isotope  to 
radiate  directly  into  space  when  the  power  system  was  not  functioning.  Even 


288 


Skipping  "The  Next  Logical  Step" 

nuclear  enthusiasts  had  to  admit  that  this  machinery  and  auxiliary  hardware 
would  probably  not  be  available  for  at  least  10  years.26 

In  trying  to  choose  between  the  various  options  for  the  on-board  power 
supply,  the  "power  plant"  subcommittee  of  Langley's  Manned  Space  Labo- 
ratory Research  Group  reviewed  several  pertinent  R&D  programs  involving 
solar  and  nuclear  power  plants  then  under  development  by  NASA,  the  air 
force,  and  the  Atomic  Energy  Commission;  however,  after  this  review,  the 
subcommittee  was  still  undecided  about  the  best  power  source.  In  a  feasibil- 
ity study  of  the  rotating  hexagon  conducted  by  North  American  Aviation, 
solar  power  was  the  favored  source.  According  to  the  company's  proposed 
design,  a  group  of  solar  cells  and  associated  electrical  batteries  could  be 
mounted  successfully  on  the  six  main  modules  as  well  as  on  the  hub  of  the 
space  station.  When  Langley's  power  plant  subcommittee  evaluated  the  so- 
lar modular  system,  they  judged  it  to  be  the  most  feasible  in  the  near  term 
because  the  system  did  not  require  the  development  of  much  technology  but 
was  still  adequate  to  meet  the  projected  station's  electric  power  needs.27 

This  evaluation  only  temporarily  resolved  the  controversy  about  which 
type  of  power  plant  to  incorporate  in  the  study  configurations.  Several 
Langley  researchers  who  favored  a  small  on-board  nuclear  reactor  (and  who 
were  to  be  closely  associated  with  subsequent  space  station  planning  at  the 
center)  were  never  convinced  by  the  arguments  in  favor  of  solar  power.  This 
small  group,  whenever  the  opportunity  arose,  would  argue  that  energy  from 
a  naturally  decaying  radioactive  isotope  ultimately  offered  the  best  means  of 
powering  a  space  station.  However,  this  group  never  could  overcome  the  fear 
that  many  researchers  had  about  a  nuclear  accident,  no  matter  how  remote 
that  possibility  might  be.  If  the  small  canister  carrying  the  radioactive 
isotope  ever  happened  to  crash  into  the  earth,  because  of  a  launch  failure, 
for  instance,  the  results  of  the  contamination  could  be  catastrophic.28 

The  issue  of  the  power  supply  was  critical  to  the  design  of  the  space 
station  because  of  the  "human  factor."  As  everyone  involved  with  space 
station  research  understood,  the  greatest  single  draw  on  the  power  supply 
would  be  the  systems  necessary  to  keep  the  crew  inside  the  space  station 
alive  and  in  good  physical  and  emotional  condition.  In  fact,  the  human 
factor  was  central  to  all  the  elements  of  space  station  design:  the  gravity 
and  energy  requirements,  the  sources  of  wobble,  the  number  and  sizes  of 
modules  and  ferry  vehicles,  the  number  and  length  of  missions,  and  the  types 
of  internal  furnishings  and  accommodations.  Human  occupancy  established 
the  central  parameters  for  the  entire  research  and  design  process.  The  job 
of  the  Langley  space  station  group  was  not  to  build  the  actual  hardware 
that  would  sustain  human  life  in  space;  rather,  it  was  to  "evaluate  and 
originate  basic  concepts  of  life  support  systems."  This  evaluation  was  to 
include  exploration  of  a  range  of  prototypes  to  generate  the  technological 
knowledge  that  could  form  the  basis  for  an  "optimum-system  concept."29 

The  essential  requirements  for  a  human  life-support  system  aboard  a 
long-duration  spacecraft  in  earth  orbit  were  not  hard  to  determine.  The 

289 


Space/light  Revolution 

system  had  to  be  lightweight  and  very  dependable,  and  it  had  to  consume  as 
little  energy  as  possible.  It  would  have  to  provide  oxygen  for  breathing;  food 
for  eating;  accommodations  for  sleeping,  exercising,  washing,  and  taking 
care  of  other  bodily  functions;  and  it  would  have  to  somehow  eliminate  or 
recycle  human  and  other  waste  products. 

Either  through  in-house  research  or  by  contracting  out  to  industry,  all  of 
these  basic  matters  of  life  support  and  many  others  were  thoroughly  studied 
by  the  Langley  space  station  group  in  1961  and  1962.  Several  contractors 
became  specialists  in  the  development  of  experimental  mechanisms  for 
collecting,  treating,  reclaiming,  and  disposing  of  solid  and  liquid  wastes. 
For  its  rotating  hexagon,  North  American  invented  a  method  for  carbon 
dioxide  removal  involving  a  regenerative  molecular  sieve.  Small  silica  gel 
beds  removed  water  vapor  from  the  air  and  passed  it  into  the  molecular  sieve, 
which  then  either  vented  the  absorbed  water  and  exhaled  carbon  dioxide  to 
the  outside  or  shunted  it  to  an  oxygen  regeneration  system.30 

None  of  the  solutions  proposed  during  this  period,  however,  were  com- 
pletely satisfactory.  What  Langley  researchers  wanted  for  their  optimum 
space  station  was  a  totally  closed  water-oxygen  system — one  that  did  not 
have  to  be  resupplied  from  the  ground.  In  the  early  years,  many  problems 
associated  with  such  a  closed  life-support  system  appeared  relatively  easy  to 
solve,  but  they  proved  troublesome.  This  was  especially  true  for  the  water 
recovery  and  recycling  system  in  which  the  astronauts'  urine  was  to  be  con- 
verted into  drinking  water.  In  the  early  years,  researchers  tried  such  things 
as  simply  blowing  air  over  the  liquid  waste,  controlling  the  odor  by  using 
a  bactericide,  and  evaporating  the  water  on  a  cold  plate.  Unfortunately,  a 
huge  amount  of  power  was  needed  to  do  that,  and  it  was  more  power  than 
any  space  station  could  afford.  The  astronauts'  natural  aversion  to  drinking 
water  made  in  this  manner  also  posed  a  problem.  Psychological  studies, 
however,  showed  that  thirst  would  quickly  overcome  disgust.  Today,  after 
more  than  30  years  of  space  station  research,  effective  technology  for  such  a 
closed  water  recycling  system  still  does  not  exist.31 

Langley  researchers  went  to  great  lengths  to  discover  the  unknowns  of  life 
in  a  spinning  spacecraft.  One  fun-loving  group  made  a  trip  to  the  amusement 
park  at  Buckroe  Beach  near  the  mouth  of  the  Chesapeake  Bay  to  ride  the 
carousel.  They  took  a  bunch  of  tennis  balls  with  them  to  throw  back  and 
forth  while  sitting  atop  their  colorfully  painted  wooden  ponies.  The  man 
attending  the  carousel  soon  threw  the  researchers  off  the  ride  because  they 
were  making  children  sick.  But  even  this  information  was  instructive  about 
Coriolis  effects  on  astronaut  equilibrium  and  hand-eye  coordination.32 

Of  course,  the  Langley  researchers  also  carried  out  many  less  frivolous  and 
more  systematic  simulations  of  human  performance  in  space.  To  investigate 
how  the  effects  of  rotation  might  conceivably  hamper  astronaut  performance, 
the  space  station  group  put  several  volunteers,  including  a  few  Langley  test 
pilots,  into  simulators  that  mimicked  the  rotations  of  a  space  station.  Some 
of  these  volunteers  managed  to  stay  in  the  simulator  for  several  hours  before 

290 


Skipping  "The  Next  Logical  Step" 

asking  (or  in  some  cases,  demanding)  to  be  let  out.  Data  from  other  man- 
in-space  simulations,  some  of  them  done  to  garner  real-time  data  about  how 
crews  would  do  during  7-day  and  14-day  missions  to  the  moon,  also  shed 
light  on  what  to  expect  inside  a  space  station.33 

Overall,  the  early  findings  about  the  ability  of  humans  to  adapt  to  life  in 
space  were  quite  reassuring.  Simple  adjustments  in  sleep  and  work  schedules 
alleviated  astronaut  fatigue  and  boredom.  An  on-board  exercise  program 
would  forestall  marked  deterioration  in  muscle  tone  and  other  physiological 
functions  at  zero  gravity.  Most  importantly,  a  weeklong  confinement  of  a 
three-person  crew  within  the  close  quarters  of  a  module  had  no  detrimental 
effect  on  performance,  nor  did  it  trigger  psychological  stress.  In  short, 
the  Langley  simulations  of  1961  and  1962  reinforced  a  growing  body  of 
evidence  that  humans  could  indeed  live  successfully  in  space,  and  could 
remain  physically  and  mentally  healthy  and  able  to  carry  out  complex  tasks 
for  extended  periods. 

Other  critical  matters,  however,  still  demanded  study.  To  see  if  a  com- 
fortable "shirt-sleeve"  working  environment  could  be  provided  for  astronauts 
inside  the  space  station,  Langley  researchers  worked  with  a  thermal  vac- 
uum chamber  in  which  they  put  small,  scale  models  of  their  inflatable  torus 
and  rotating  hexagon  designs.  Built  for  Langley  by  Grumman,  this  cham- 
ber employed  an  arc  that  served  as  the  "sun"  and  smaller  electric  heaters 
that  served  as  analogues  for  heat-producing  humans.  After  several  weeks  of 
tests  with  this  thermal  chamber,  researchers  found  that  the  North  American 
hexagon  design,  because  of  its  insulated,  protective  walls,  was  superior  to 
the  torus.34 

Protecting  the  occupants  of  the  space  station  from  the  heat  of  the  sun 
was  one  thing;  protecting  them  from  meteorite  showers  was  still  another. 
Into  the  early  1960s,  according  to  a  Langley  study,  NASA  still  faced  "severe 
uncertainties  regarding  the  basic  structure  of  manned  space  stations."  How 
should  the  walls  of  such  a  structure  be  built,  and  out  of  what  materials? 
They  had  to  be  light  because  of  launch-weight  considerations  and  built  of 
a  material  that  would  help  in  the  control  of  internal  temperatures.  The 
walls  also  had  to  provide  dependable  and  long-term  protection  from  major 
meteoroid  penetrations;  some  small  chinks  and  dents  in  the  sides  of  a  space 
station  might  cause  no  trouble,  but  big  hits,  especially  in  the  case  of  an 
inflatable  torus,  could  prove  disastrous.  Thus,  structures  experts  at  both 
Langley  and  Ames  looked  for  the  type  of  wall  structure  that  offered  the 
greatest  protection  for  the  least  weight.  They  turned  to  a  sandwiched 
structure  with  an  inner  and  an  outer  wall.  Developed  by  North  American 
for  the  rotating  hexagon,  the  outer  wall  was  a  "meteoroid  bumper"  made 
of  aluminum,  backed  by  a  polyurethane  plastic  filler  that  overlay  a  bonded 
aluminum  honeycomb  sandwich.  Such  a  wall  seemed  to  meet  the  design 
criteria,  but  no  one  could  be  sure  because  the  actual  velocities  of  meteoroid 
impacts  were  impossible  to  simulate  in  any  ground  facility.  The  only  thing 
to  do  was  to  make  further  studies.  For  the  inner  wall,  Langley's  space 

291 


Space/light  Revolution 

station  office  looked  into  the  efficacy  of  nylon-neoprene,  dacron-silicone, 
saran,  Mylar  (E.  I.  du  Pont  de  Nemours  &  Co.,  Inc.),  polypropylene, 
Teflon  (E.  I.  du  Pont  de  Nemours  &  Co.,  Inc.),  and  other  flexible  and 
heat-absorbing  materials.  These  materials  could  not  be  toxic  or  leak  gases 
(especially  oxygen),  and  they  had  to  be  able  to  withstand  a  hard  vacuum, 
electromagnetic  and  particle  radiation,  and  temperatures  ranging  from  -50° 
to  150°F.35 

At  a  symposium  held  at  Langley  in  late  July  1962,  the  Langley  staff 
assembled  in  the  large  auditorium  in  the  center's  main  activities  building 
to  present  summary  progress  reports  on  their  exploratory  space  station 
research.  By  the  time  of  this  symposium,  Langley's  space  station  researchers 
had  arrived  at  four  key  conclusions: 

(1)  The  rotating  hexagon  was  superior  to  the  inflatable  torus;   a  15- 
foot  scale  model  of  the  North  American  design  had  been  undergoing  tests 
at  Langley  for  several  months,  whereas  studies  of  the  torus  had  virtually 
ceased.  6 

(2)  Although  the  hexagon  was  preferable  to  the  torus,   the  Langley 
researchers  knew  that  they  had  not  yet  discovered  the  optimum  design  and 
were  committed  to  carrying  out   "the  conceptual  design  of  several  space 
stations  in  order  to  uncover  the  problem  areas  in  such  vehicles."37 

(3)  A  flight  program,  something  akin  to  a  Project  Shotput,  was  needed 
to  extend  space  station  research.    The  space  environment  was  difficult  to 
impossible  to  simulate  in  a  ground  facility,  thus  making  tests  of  station 
materials  impossible  as  well.  As  early  as  May  1961,  members  of  the  Langley 
space  station  office  had  proposed  using  a  Scout  rocket  to  test  the  deployment 
of  a  10-foot  version  of  the  inflatable  torus  at  an  orbit  of  220  miles;  however, 
the  idea  for  the  test  had  gone  nowhere.38 

(4)  Whatever  R&D  was  to  be  done  on  space  stations  in  the  future, 
the  researchers  wanted  their  work  to  be  guided  by  the  broad  objectives 
of  learning  how  to  live  in  space,  of  making  the  station  a  place  for  scientific 
research,  and  of  finding  ways  to  make  the  station  "a  suitable  facility  for 
learning  some  of  the  fundamental  operations  necessary  for  launching  space 
missions  from  orbit."    Moreover,  they  wanted  their  space  station  efforts  to 
be  better  integrated  with  the  overall  NASA  effort.39 

Langley  researchers  regarded  a  manned  space  station  as  more  than  a 
jumping-off  point  for  Apollo  or  for  some  other  specific  mission.  They 
thought  of  it  as  a  versatile  laboratory  in  space,  a  Langley  research  operation 
that  happened  to  be  located  hundreds  of  miles  above  the  earth  rather  than  in 
Tidewater  Virginia.  Just  as  Langley  had  always  explored  the  basic  problems 
of  flight  with  a  view  to  their  practical  solution,  the  ultimate  use  of  a  space 
station  was  "for  continuing  to  advance  the  technology  of  space  flight."40  The 
objective  was  long-term,  not  just  immediate. 

How  the  space  station  would  fare  without  any  direct  application  to 
the  Apollo  lunar  landing  program  was  a  question  that  loomed  over  the 
researchers  at  the  symposium.  With  an  expensive  Apollo  program  in 

292 


Skipping  "The  Next  Logical  Step" 

progress  and  LOR  the  chosen  strategy,  Washington's  support  for  an  earth- 
orbiting  space  station  might  quickly  plummet,  no  matter  what  Langley 
scientists  and  engineers  had  to  say  about  the  potential  benefits  of  space 
station  operations.  If  the  space  station  was  to  be  built  in  the  near  future, 
Langley  would  have  to  quickly  reconcile  the  objectives  of  the  station  with 
those  of  the  Apollo  mission. 

Manned  Orbital  Research  Laboratory 

In  the  months  following  the  in-house  symposium,  Langley  management 
initiated  a  revised  program  of  space  station  studies  that  would  better  dove- 
tail with  the  Apollo  lunar  landing  program.  In  late  1962,  this  determination 
brought  forth  a  more  focused  space  station  effort — one  that  proved  to  be 
qualitatively  quite  different  from  the  broader  conceptual  studies  that  had 
given  birth  to  the  inflatable  torus  and  the  rotating  hexagon.  As  a  result 
of  this  concentrated  effort,  Langley  researchers  in  early  1963  conceived  a 
smaller  and  more  economical  space  station  that  would  complement  and  make 
maximum  use  of  the  technological  systems  being  developed  for  Apollo.  They 
called  it  the  Manned  Orbiting  Research  Laboratory,  or  MORL,  for  short. 

The  original  MORL  concept  evolved  within  Langley's  space  station 
group.  The  idea  was  for  a  "minimum  size  laboratory  to  conduct  a  na- 
tional experimental  program  of  biomedical,  scientific,  and  engineering  ex- 
periments," with  the  laboratory  to  be  specifically  designed  for  launch  in  one 
piece  atop  a  Saturn  I  or  IB.  The  goal  of  the  MORL  program  was  to  have  one 
crew  member  stay  in  space  for  one  year,  with  three  other  crew  members  on 
board  for  shorter  periods  on  a  rotating  schedule.  Langley  wanted  to  achieve 
this  goal  in  1965  or  1966,  a  few  years  before  the  anticipated  first  manned 
Apollo  flight,  and  accomplish  it  in  unison  with  Project  Gemini,  the  NASA 
program  that  bridged  Mercury  and  Apollo,  whose  basic  purpose  was  to  re- 
solve the  problems  of  rendezvous  and  docking  and  of  long-duration  manned 
spaceflights  necessary  for  a  successful  lunar  landing  via  LOR.  The  MORL 
would  be  launched  unmanned  by  a  Saturn  booster  into  a  circular  orbit  from 
Cape  Canaveral,  and  after  a  short  checkout  period,  two  crew  members  in  a 
Titan-mounted  Gemini  spacecraft  (then  under  development)  would  "ascend 
to  the  laboratory's  orbit  and  complete  a  rendezvous  and  docking  maneu- 
ver." A  few  weeks  later,  two  more  crew  members  would  join  the  laboratory 
by  the  same  method,  completing  the  four-person  crew.  One  new  astronaut 
would  enter  the  laboratory  at  each  crew  change,  thus  providing  a  check  on 
the  cumulative  effects  of  weightlessness  on  the  total  capability  of  the  crew. 
Three  of  the  astronauts  would  occupy  the  space  station  for  only  parts  of 
a  year;  only  one  crew  member  would  try  to  complete  a  full  year's  mission. 
Every  90  days  or  less,  an  unmanned  resupply  spacecraft  launched  by  an 
Atlas- Agena  combination  would  be  orbited  and  brought  by  radio  control  to 
a  rendezvous  with  one  of  the  laboratory's  multiple  docking  ports.  These 
ports  would  not  only  provide  the  means  for  the  crew  rotations  and  any 

293 


Space/light  Revolution 

emergency  evacuations  but  also  would  serve  as  attachment  sites  for  cargo 
and  experiment  modules.41 

By  the  spring  of  1963,  Langley  management  judged  the  MORL  design 
ready  for  industry  evaluation.  A  contractor  was  to  look  for  ways  of 
improving  and  refining  the  concept  into  what  engineers  called  a  "baseline 
system,"  that  is,  a  detailed  plan  for  an  optimum  MORL  configuration.  In 
late  April,  Langley  asked  the  aerospace  industry  to  submit  brief  proposals 
by  14  May  for  a  contract  study  of  "Manned  Orbital  Research  Laboratory 
Systems"  capable  of  sustaining  such  a  rotating  four-person  crew  in  space 
for  one  year.  The  Request  for  Proposals  outlined  an  industry  competition 
in  two  phases:  Phase  I  was  to  be  an  open  competition  for  "comparative 
study  of  several  alternative  ways  to  obtain  the  orbital  laboratory  which  is 
envisioned";  Phase  II  was  to  be  a  closed  contest  between  the  two  winners 
of  the  first  competition,  for  "preliminary  design  studies."  If  all  progressed 
well  and  NASA  approved  continued  work  on  MORL,  Langley  might  propose 
a  follow-on  to  the  second  phase  (Phase  II-A)  in  which  "a  single  contractor 
would  be  requested  to  synthesize  into  a  mature  concept"  the  design  study 
that  had  been  judged  by  NASA  as  the  most  feasible  and  to  furnish  a  baseline 
configuration  for  a  complete  orbital  laboratory  system.  Yet  another  phase 
(Phase  II-B)  might  involve  a  final  design  stage,  including  test  mock-ups  of 
the  laboratory  and  resupply  spacecraft.42 

Phase  I,  the  open  competition,  lasted  only  until  mid-June  1963,  when 
Langley  Director  Floyd  Thompson  announced  that  from  the  11  proposals 
received,  he  had  selected  those  from  Boeing  and  Douglas  as  the  winners. 
An  11-member  in-house  MORL  Technology  Steering  Committee,  chaired  by 
Paul  R.  Hill  of  the  Applied  Materials  and  Physics  Division  space  station 
office,  had  helped  Thompson  with  the  selection.  At  the  same  time  that 
Thompson  established  this  ad  hoc  steering  committee,  on  6  June,  he  also 
created  a  small  MORL  Studies  Office,  which  comprised  originally  only  six 
members  and  was  to  report  directly  to  the  director's  office.  Thompson  chose 
someone  new  to  space  station  research  to  head  the  new  office.  William 
N.  Gardner,  formerly  head  of  the  Flight  Physics  Branch  of  the  Applied 
Materials  and  Physics  Division,  and  his  six-person  staff  were  to  implement 
the  management  of  the  study  contracts  soon  to  be  awarded  to  Boeing  and 
Douglas.  Thompson  also  formalized  the  many  R&D  efforts  relating  to  a 
space  station  that  had  popped  up  inside  the  Applied  Materials  and  Physics 
Division.  He  did  this  on  10  June  by  establishing  a  new  19-member  Space 
Station  Research  Group,  with  Robert  S.  Osborne,  a  veteran  of  the  center's 
previous  space  station  office,  in  charge.43 

Langley's  revised  space  station  effort  had  not  progressed  without  a  hitch. 
Earlier  in  1963,  still  in  the  immediate  wake  of  the  LOR  decision,  NASA 
headquarters  had  threatened  the  cancellation  of  all  the  MORL  research 
at  the  research  center.  To  have  it  reinstated  even  on  a  provisional  basis, 
Langley  Associate  Director  Charles  J.  Donlan,  who  from  the  start  had  lent 
strong  support  to  space  station  research  at  Langley,  traveled  to  Washington 

294 


Skipping  "The  Next  Logical  Step" 

with  some  of  the  most  articulate  members  of  the  center's  space  station  group 
for  several  meetings  with  old  friends  and  other  NASA  officials.  Donlan  had 
always  been  a  strong  supporter  of  Langley's  space  station  research,  and 
together  with  associates  he  argued  that  a  manned  space  station  was  still 
"the  next  logical  step,"  after  Apollo,  and  was  thus  likely  to  be  a  central 
part  of  the  agency's  post- Apollo  planetary  exploration.  Donlan  pointed  out 
that  a  manned  orbital  laboratory  offered  perhaps  the  only  way  of  making 
many  necessary  studies  such  as  the  effects  of  weightlessness. 

Eventually,  the  lobbying  paid  off.  In  the  spring  of  1963,  Bob  Seamans 
issued  MORL  a  reprieve,  thus  arranging  for  the  authorization  Langley 
needed  to  proceed  with  the  first  phase  of  the  industry  competition.  At 
the  start,  that  was  the  only  permission  Langley  had.  When  Phase  I  started, 
NASA  headquarters  had  not  yet  approved  Phase  II  and  had  certainly  not 
given  the  go-ahead  for  any  follow-on  phases.44 

Some  of  the  ground  rules  for  Phase  I  of  the  MORL  competition  conformed 
closely  to  the  general  specifications  of  the  rotating  hexagon,  but  others 
reflected  some  significant  changes  in  the  way  Langley  was  now  thinking 
about  space  stations.  The  major  shift  in  thinking  was  the  realization,  gained 
by  American  and  Soviet  experiences  with  manned  spaceflight,  that  humans 
could  in  fact  function  quite  well  in  zero  gravity,  at  least  for  several  orbits, 
without  serious  ill  effects.  If  a  few  days  of  weightlessness  did  not  debilitate 
an  astronaut,  the  same  would  most  likely  hold  true  for  a  couple  of  weeks. 
Further  experimentation  certainly  had  to  be  done  to  determine  exactly  how 
long  humans  could  perform  in  zero  gravity,  but  as  reflected  in  MORL  ground 
rules,  researchers  were  growing  confident  that  a  person  might  be  able  to 
perform  well  in  space  for  as  long  as  a  year.  When  Langley  asked  industry 
in  April  1963  to  design  MORL  with  zero  gravity  as  the  primary  operating 
mode,  it  was  abandoning  once  and  for  all  the  long-held  notion  that  a  space 
station  must  continually  rotate  to  provide  artificial  gravity.45 

Douglas  and  Boeing  took  Phase  II  of  the  competition  seriously,  each 
assembling  its  MORL  personnel  into  a  team  situated  at  a  single  plant  (Santa 
Monica  for  Douglas  and  Seattle  for  Boeing).  Douglas  had  shown  interest 
in  a  space  station  for  some  time;  in  1958,  the  company  had  won  a  $10,000 
first  prize  in  a  contest  for  a  design  of  "A  Home  in  Space,"  which  had  been 
sponsored  by  the  London  Daily  Mail.4Q  Douglas  had  also  been  a  serious 
bidder  for  the  NASA  contract  for  a  six-month  study  of  Berglund's  rotating 
hexagon  concept,  which  NASA  had  awarded  to  North  American  in  the 
summer  of  1961. 47  Boeing,  on  the  other  hand,  was  something  of  a  newcomer 
to  the  field  of  space  exploration.  However,  as  the  reader  shall  learn  in  more 
detail  in  the  next  chapter,  the  well-known  airplane  manufacturer  was  at 
this  time  completing  a  solid  performance  in  the  Bomarc  missile  program 
and  was  keen  to  be  involved  with  the  civilian  space  program.  Not  only  did 
Boeing  want  the  space  station  study  contract,  it  also  wanted  to  become  the 
prime  contractor  for  the  ambitious  Lunar  Orbiter  project,  the  unmanned 


295 


Spaceflight  Revolution 


Douglas  engineers  incorporated  the 
idea  of  a  two-person  centrifuge  into 
their  winning  MORL  baseline  con- 
figuration proposal  in  1963  (right). 
The  centrifuge  (bottom,  second  cut- 
away from  left)  was  to  serve  as  a 
possible  remedial  or  therapeutic  de- 
vice for  enhancing  the  astronauts' 
tolerance  to  weightless  conditions  and 
for  preconditioning  crew  members 
for  the  stresses  of  reentry. 


L-63-1132 


296 


Skipping  "The  Next  Logical  Step" 

photographic  mission  to  the  moon  which  NASA  was  planning  in  order  to 
select  the  best  landing  sites  for  Apollo. 

A  NASA  "technical  assessment  team"  consisting  of  43  engineers  (36 
of  them  from  Langley)  and  organized  into  four  review  panels  ("Major 
Systems  Configuration  and  Integration,"  "Subsystems  Configuration  and 
Integration,"  "Operations,"  and  "Management  and  Planning")  looked  very 
carefully  and  fairly  at  both  MORL  studies  in  late  September  1963  before 
recommending  the  Douglas  study  to  the  Langley  director.48  Perhaps  NASA 
was  reluctant  to  give  a  company  inexperienced  in  space  exploration  the 
responsibility  for  doing  two  big  new  jobs  at  one  time.  (Boeing  had  just  been 
awarded  the  contract  for  Lunar  Orbiter.)  More  likely,  however,  the  Douglas 
proposal  was  simply  superior.  Members  of  the  MORL  Studies  Office  at 
Langley  had  spent  many  hours  at  the  plants  of  the  contractors  assessing 
their  space  station  work,  and  they  knew  firsthand  the  capabilities  of  their 
assembled  teams. 

For  the  next  two  months,  NASA  Langley  negotiated  with  Douglas  (and 
with  NASA  headquarters,  where  the  approval  for  Phase  II-A  was  still 
uncertain)  over  the  details  of  what  would  come  next:  a  six-to-nine-month 
study  at  the  end  of  which  Douglas  would  furnish  a  baseline  system  that 
would  be  so  detailed  and  fully  documented  that  a  final  design  could  be 
prepared  from  it,  if  NASA  so  chose.  By  mid-December  all  the  parties 
involved  reached  an  agreement,  and  on  20  December,  as  a  nice  little 
Christmas  present,  NASA  awarded  Douglas  a  Phase  II-A  nine-month  study 
contract  worth  just  over  $1.4  million  to  refine  its  winning  MORL  concept.49 

The  baseline  configuration  fleshed  out  by  the  Douglas  engineers  between 
December  1963  and  August  1964  proved,  not  surprisingly,  to  be  a  mixture  of 
old  and  new  ideas.  As  had  been  the  case  with  Langley's  former  pet  concept, 
the  Berglund/North  American  rotating  hexagon,  Douglas's  baseline  facility 
would  be  carried  into  orbit  as  a  unit  aboard  a  Saturn  launch  vehicle.  As  had 
been  proposed  for  the  hexagon,  the  first  generation  MORL  would  be  powered 
by  solar  cells,  but  with  either  a  nuclear  reactor  or  isotope  Brayton  cycle 
system  phased  in  at  an  early  date.  The  same  life-support  systems  for  meeting 
the  physical  needs  of  a  small  crew  in  a  shirt-sleeve  working  environment 
would  also  be  part  of  MORL.  As  before,  many  of  MORL's  design  features, 
such  as  separate  zero-gravity  and  artificial- gravity  operational  modes,  would 
in  effect  serve  as  experiments  that  would  yield  data  applicable  to  other 
manned  space  programs. 

Because  Langley's  thinking  changed  about  what  was  best  for  a  space 
station  in  the  age  of  Apollo,  Douglas's  baseline  system  for  MORL  involved 
key  differences  from  all  the  previous  space  station  concepts.  Unlike  the 
earlier  configurations,  which  had  unitized  structures,  MORL  would  consist  of 
a  series  of  discrete  modules.  The  modular  approach  would  promote  greater 
flexibility  of  function:  MORL  could  grow  with  evolving  space  technology 
and  over  time  serve  multiple  purposes  for  a  varied  constituency,  including 
perhaps  the  DOD.  The  DOD  had  been  carrying  out  its  own  manned  space 

297 


Space/light  Revolution 


MORL 
&  LOGISTICS  SYSTEM 


According  to  the  briefing  manual  submitted  by  Douglas  to  NASA  Langley  in  August 
1964,  MORL  "provides  a  flexible,  expandable  facility  developed  in  a  manner  similar 
to  current  submarine  concepts  that  permit  redundancy  of  life-support  equipment  and 
evacuation  from  one  compartment  to  another.  "As  shown  in  this  illustration  from 
the  manual,  MORL  was  to  be  launched  by  an  Apollo  Saturn  SIB. 

station  R&D  since  the  Military  Test  Space  Station  (MTSS)  project  of  the 
late  1950s  and  was  currently  involved  in  a  study  of  what  it  called  the 
National  Orbiting  Space  Station  (NOSS).50 

Besides  benefiting  the  military,  the  MORL  could  serve  the  progress 
of  science,  in  general.    This  was  a  mission  capability  that  had  not  been 

298 


Skipping  "The  Next  Logical  Step" 


L-64-841 

Two  Langley  engineers  test  an  experimental  air  lock  between  an  arriving  spacecraft 
and  a  space  station  portal  in  January  1964- 


especially  emphasized  in  the  earlier  space  station  studies.  When  consider- 
ing the  inflatable  torus  and  rotating  hexagon,  Langley  researchers  and  their 
contractors  had  envisioned  only  a  limited  role  for  general  scientific  exper- 
imentation aboard  the  station,  but  the  Douglas  engineers  were  beginning 
to  see  the  MORL  as  a  facility  for  research  covering  the  spectrum  of  scien- 
tific disciplines.  In  addition  to  carrying  one  or  more  astronomical  telescopes 
(a  capability  that  proponents  of  a  space  station  had  in  fact  been  pushing 
from  the  start),  the  MORL  could  be  designed  to  have  a  self-contained  mod- 
ule for  biological  studies  involving  animals,  plants,  and  bacteria.  Such  re- 
search had  potential  applications  not  only  in  basic  life  sciences  research  but 
also  in  medicine  and  pharmaceuticals.  For  geologists,  oceanographers,  and 
meteorologists,  Douglas  provided  a  specialized  nine-lens  camera  system  for 
multiband  spectral  reconnaissance  of  earth  features  and  weather  systems.  A 
special  radar  system  could  be  placed  on  board  to  garner  the  data  necessary 
for  large-scale  topographical  mapping. 

This  was  not  all  that  the  MORL  could  provide.  The  orbital  station  would 
also  be  the  ideal  place  to  study  subsystems  for  interplanetary  vehicles  and 
their  propulsion  systems,  technologies  that  could  not  be  tested  adequately 
on  the  ground.  Douglas's  integrated  plan  even  included  using  the  MORL 
in  lunar  orbit  to  provide  surface  observation  and  mapping,  landing  site 
selection,  and  LEM  support.  With  such  capabilities,  NASA  might  not  need 

299 


Space/light  Revolution 


L-64-4636 

William  N.  Gardner,  head  of  the  MORL  Studies  Office,  explains  the  interior  design 
of  the  space  station  at  the  1964  NASA  inspection. 


the  unmanned  Lunar  Orbiter  program.  If  equipped  with  a  state-of-the-art 
landing  stage,  the  MORL  could  land  on  the  moon  and  become  a  long-term 
base  for  exploration.  MORL  could  serve  as  the  jumping  off  point  for  a 
manned  mission  to  Mars  and  as  a  module  of  a  planetary-mission  vehicle 
in  which  a  crew  would  investigate  the  physical  environment  and  assess  the 
habitability  of  a  selected  planet.51 

In  fact,  as  Douglas  touted  it,  there  was  little  that  the  MORL  system 
could  not  do,  if  NASA  wanted  it  done.  Thus,  while  trying  to  stay  within  the 
political  and  economic  framework  of  Apollo,  the  proponents  of  the  MORL 
were  actually  demonstrating  how  a  versatile  space  station  could  greatly 
expand  U.S.  capabilities  in  space  and  make  new  exploration  possible.  The 
MORL  would  have  spin-off  studies  in  areas  such  as  biology,  medicine,  and 
possibly  industrial  manufacturing,  which  would  ultimately  benefit  all  sectors 
of  society.  The  lunar  landing  program,  by  itself,  would  make  few  of  those 
things  possible.  But  in  1964  that  was  a  point  that  neither  the  Langley  space 
station  advocates  nor  their  counterparts  in  industry  dared  to  make,  given 
the  national  commitment  to  Apollo. 


300 


Skipping  "The  Next  Logical  Step" 
Keeping  the  "R"  Alive 

All  in  all,  Langley  was  happy  with  the  baseline  system  that  Douglas 
submitted  to  NASA  in  August  1964  and  was  interested  in  moving  to 
Phase  II-B  in  which  full-scale  mock-ups  of  the  laboratory  would  be  tested  in 
preparation  for  a  final  MORL  design.  By  1964,  however,  MORL  was  facing 
stiff  competition  from  other  space  station  concepts,  not  to  mention  space 
projects  proposed  by  other  NASA  centers. 

As  Phase  II-A  of  the  MORL  began  in  early  1964,  the  Office  of  Manned 
Space  Flight  at  NASA  headquarters  was  considering  what  to  do  next  with 
several  other  space  station  designs.  Most  of  these  ideas  came  from  either 
Houston  or  Huntsville.  The  most  ambitious  of  these  schemes  called  for 
a  Large  Orbiting  Research  Laboratory  (LORL),  a  huge  structure  to  be 
launched  unmanned  by  a  Saturn  V,  with  a  volume  more  than  seven  times 
that  of  MORL  (67,300  cubic  feet  compared  with  MORL's  9000)  and  a  weight 
more  than  10  times  greater  (74,600  pounds  versus  6800).  According  to  the 
plan,  LORL  would  be  capable  of  holding  a  24-person  crew  for  five  years;  as 
such,  it  was  the  "Cadillac"  of  NASA's  space  station  concepts  at  the  time. 
At  the  other  end  of  the  design  spectrum  was  a  "Volkswagon"  version  known 
as  "Apollo  X."  This  space  station  (only  600  cubic  feet  in  volume)  was  based 
entirely  on  Apollo  technology;  a  modified  Apollo  command  module  would 
be  used  as  a  small  orbital  workshop.  Manned  from  time  of  launch,  Apollo 
X  would  thus  be  a  small  "limited-life"  laboratory  serving  a  crew  of  two  for 
30  to  120  days.  Between  these  two  extremes  were  various  designs  for  some 
type  of  Apollo  Orbital  Research  Laboratory  (AORL),  a  medium-size  station 
(5600  cubic  feet  in  volume)  that  would  have  an  extended  life  of  two  years, 
with  a  crew  of  three  to  six.52 

Besides  the  NASA  concepts,  military  space  station  ideas  also  had  to  be 
considered.  Interagency  agreements  had  been  made  related  to  the  Gemini 
program  requiring  that  all  planning  for  manned  earth-orbital  missions  and 
supporting  technology  be  coordinated  between  NASA  and  the  DOD.  As 
mentioned  earlier,  the  DOD,  particularly  the  air  force,  was  busy  conducting 
its  own  space  station  studies.  By  late  1963,  experts  in  the  DOD  were  keenly 
interested  in  the  potential  military  applications  of  MORL  or  of  a  revised 
MORL  design  for  an  air  force  Gemini-based  Manned  Orbiting  Laboratory 
(MOL)  in  which  NASA's  research  component  was  left  out.  Even  before  the 
Phase  II  contract  was  awarded  to  Douglas,  managers  in  the  OART  at  NASA 
headquarters  had  been  referring  not  to  MORL  studies,  but  simply  to  MOL, 
which  Secretary  of  Defense  McNamara  and  NASA  Administrator  Webb  were 
coming  to  see  as  a  way  of  combining  DOD  and  NASA  first-generation  space 
station  objectives.53 

Surprisingly,  Langley  researchers  seem  to  have  accepted  the  shift  from 
MORL  to  MOL  without  complaint.  In  the  minutes  of  the  28  October  1963 
meeting  of  the  Langley  MORL  Technology  Steering  Committee,  secretary 


301 


Spaceflight  Revolution 


L-65-7433 

The  MORL-Saturn  IB  launch  combination  undergoes  aerodynamic  testing  in  the 
8-Foot  Transonic  Tunnel  in  October  1965. 

John  R.  Dawson  noted  with  emphasis  that  MORL  was  being  "redesignated 
MOL"  and  that  "MOL  Phase  IIA  was  now  planned  so  as  to  fit  with  DOD 
coordination  requirements."  In  all  the  committee  minutes  following  that 
meeting,  Dawson  always  referred  to  "MOL  Phase  II"  rather  than  to  MORL 
Phase  II.54  So,  too,  would  the  Langley  press  release  of  2  December  1963 
refer  to  MOL  Phase  II.  (In  this  release,  NASA  announced  that  Douglas 
had  won  the  second  round  of  competition  over  Boeing  for  a  "follow-on 
study  contract  for  refinement  and  evaluation  of  a  NASA  Manned  Orbital 
Laboratory  concept.")  Somehow  the  "R"  in  the  space  station  plan  was 
being  erased  as  Langley  tried  to  justify  a  space  station  as  part  of  the  Apollo- 
driven  national  space  program. 

At  Langley,  researchers  did  what  they  could  to  keep  the  spirit  of  MORL 
alive.  Looking  beyond  the  industry  study  contracts,  Langley  engineers  and 
managers  invested  thousands  of  hours  in  MORL/MOL  research.  In  1963 
and  1964,  basic  studies  continued  on  a  broad  front. 

In  the  area  of  life-support  systems,  Langley  researchers  tried  to  stay 
particularly  active.  As  outlined  earlier  in  this  chapter,  Langley  early  on  had 
taken  the  lead  among  the  NASA  centers  in  this  vital  field.  In  1963,  Langley 
researchers  wanted  to  extend  their  efforts  with  a  fully  operational  prototype 
of  a  space  station  life-support  system.  In  such  a  prototype,  they  could 

302 


Skipping  "The  Next  Logical  Step" 


Langley's  Otto  Trout  suggested  as  early  as 
1963  that  zero-gravity  activities  could  be 
simulated  by  immersing  astronauts  in  a 
large  tank  of  water.  Years  later,  Marshall 
Space  Flight  Center  turned  Trout 's  abortive 
idea  into  a  major  component  of  NASA 's 
astronaut  training  program. 
L-66-7850 


test  all  the  integrated  mechanisms  for  water  management  and  sanitation, 
oxygen  regeneration,  ridding  the  system  of  waste  heat  and  gases,  and  all 
other  required  functions.  They  wanted  a  ground  test  facility — a  wind  tunnel 
of  sorts  but  one  equipped  for  the  physiology  of  humans  rather  than  for  the 
physics  of  air  molecules. 

Langley  explored  several  options  before  it  found  the  right  facility  to  meet 
these  research  needs.  Engineers  working  with  the  MORL  Studies  Office  built 
a  small  life-support  test  tank  but  found  that  the  device  could  not  be  used 
in  manned  tests  because  of  safety  concerns.  As  William  Gardner,  head  of 
the  MORL  Studies  Office  remembers,  "Essentially,  the  medical  profession 
killed  it.  The  medical  experts  who  came  in  as  consultants  would  not  endorse 
anything  we  were  doing.  We  couldn't  get  the  medical  people  to  say  it  would 
be  safe  to  do  the  tests."55  Another  bold  idea  came  from  Otto  Trout,  an 
ingenious  engineer  working  in  the  Space  Systems  Division.  Trout  suggested 
that  the  space  station  group  simulate  zero  gravity  by  immersing  test  subjects 
in  a  tank  of  water  for  long  periods.  Robert  Osborne  curtly  dismissed  Trout's 
novel  idea,  a  hasty  decision  that  Osborne  came  to  regret  when  engineers  at 
Marshall  Space  Flight  Center  took  up  the  idea  and  turned  it  into  a  major 
component  of  NASA's  astronaut  training  program.56 

Osborne  and  others  did  not  want  Gardner's  little  tank  or  Trout's  big  tank 
of  water  but  sought  an  enclosed,  self-sustaining  life-support  system  in  which 
four  human  subjects  could  live  for  as  long  as  six  months.  A  rivalry  existed 

303 


Space/light  Revolution 


The  $2.3  million  ILSS  arrives  at  Langley 
by  barge  (right)  from  its  manufacturer, 
the  Convair  Division  of  General  Dynam- 
ics, in  August  1965.  Below  is  the  home 
of  the  huge  30-ton  life-support  tank  in 
Building  1250.  Test  subjects  occupied 
this  facility  for  as  long  as  28  days  at  a 
time. 


L-65-6054 


I 


304 


Skipping  "The  Next  Logical  Step" 

between  Gardner's  MORL  group  and  Osborne's  life-support  studies  group. 
In  this  case  the  Osborne  group  won.  In  late  June  1963,  he  and  his  colleagues 
got  their  wish  when  NASA  awarded  a  contract  to  the  Astronautics  Division 
of  the  General  Dynamics  Corporation  for  the  design  and  construction  of  an 
Integrative  Life  Support  System  (ILSS).  Funded  by  the  OART's  director  of 
Biotechnology  and  Human  Research,  this  facility  was  to  be  built  by  General 
Dynamics  at  its  plant  in  San  Diego  and  shipped  to  Langley  at  a  total  cost 
of  $2.3  million.57 

Two  years  passed  before  General  Dynamics  finished  the  ILSS  unit.  The 
unique  structure  stood  18  feet  tall,  weighed  30  tons,  and  was  housed  in  a 
cylindrical  tank  18  feet  in  diameter.  When  the  big  chamber  arrived  by  barge 
at  the  dock  of  Langley  AFB  in  August  1965,  a  more  curious  structure  had 
not  been  delivered  since  the  85-ton  pressure  shell  for  the  laboratory's  historic 
Variable-Density  Tunnel  had  arrived  atop  a  railcar  from  the  Newport  News 
Shipyard  and  Dry  Dock  Co.  in  February  1922. 

The  ILSS  did  not  prove  to  be  the  landmark  facility  that  the  Variable- 
Density  Tunnel  became,  but  it  did  contribute  significant  data.  In  the  years 
following  its  long-anticipated  arrival,  manned  and  unmanned  tests  in  the 
big  test  chamber  provided  a  wealth  of  new  information  about  how  various 
life-support  systems  would  work  individually  and  together.  The  longest 
human  occupancy  experiment  lasted  28  days.  The  ILSS  test  program 
even  included  microbiological  experiments  on  possible  toxic  contaminants 
in  space.  Langley  management  heartily  supported  the  ILSS  program,  thus 
allowing  it  to  encompass  the  efforts  of  dozens  of  Langley  staff  members  in 
the  Space  Systems  and  Instrument  Research  divisions.  Associate  Director 
Charles  Donlan  even  worked  personally  on  some  aspects  of  the  project.58 

By  the  time  ILSS  came  on-line  at  Langley  in  August  1965,  however, 
NASA  knew  that  its  space  station  research  must,  out  of  political  and 
economic  necessity,  become  more  sharply  defined.  With  costs  for  Gemini  and 
Apollo  rapidly  outstripping  early  estimates  and  the  nation  in  an  increasingly 
expensive  war  in  Vietnam,  the  space  agency  realized  that  if  any  manned 
orbiting  facility  was  to  obtain  funding  and  become  a  reality,  it  would  have 
to  be  a  part  of  Apollo. 


Understanding  Why  and  Why  Not 

The  economical  Apollo  Extension  System  became  NASA's  surrogate 
choice  for  its  first  orbiting  space  station.  This  crushed  Langley  researchers' 
dreams  for  MORL.  Instead  of  a  versatile  laboratory  with  an  extended 
life  of  five  years  in  which  all  sorts  of  experiments  could  be  done,  NASA 
would  settle,  at  least  for  the  time  being,  for  a  small  space  station  with 
a  limited  life.  This  station  would  be  launched  as  soon  as  possible  after 
Apollo  astronauts  set  foot  on  the  moon.  For  the  Apollo  Extension  System, 
NASA  headquarters  asked  Langley  researchers  to  devise  potential  mission 

305 


Spaceflight  Revolution 

experiments,  tempting  them  with  the  responsibility  of  acting  as  principal 
investigators.  Osborne's  panel  on  space  station  experiments  responded 
by  collecting  experiments  in  11  categories  ranging  from  regenerative  life- 
support  systems  to  extravehicular  activities,  horizon  sensing,  and  radiation 
effects.59 

Two  years  would  pass  before  a  new  president,  Richard  Nixon,  and  the 
Congress  extinguished  what  remained  of  Langley's  hopes  for  a  multifaceted 
U.S.  space  program.  In  1967  the  Apollo  Extension  System  became  the 
Apollo  Applications  Program.  NASA  headquarters  called  upon  Langley, 
Houston,  and  Marshall  to  carry  out  independent  studies  to  "identify  the 
most  desirable  Agency  program  for  the  Saturn  workshop,"  noting  "the 
constraints  of  projected  funding  limitation."  The  outcome  at  Langley 
was  one  of  the  research  center's  last  major  contributions  to  space  station 
development:  an  "Intermediate  Orbital  Workshop  System  Study"  issued  by 
the  MORL  Studies  Office  on  28  June  1968.60 

The  concluding  remarks  of  this  1968  in-house  report  encapsulate  the 
years  of  hard  work  and  intellectual  energy  Langley  designers  and  researchers 
had  devoted  to  the  idea  of  a  U.S.  civilian  space  station.  The  report  described 
a  versatile  facility  that  "should  be  and  can  be  inherently  capable  of  growth 
into  the  ultimate  space  station  which  will  provide  broad  capability  manned 
systems."  True  to  the  original  Langley  vision,  it  called  for  a  two-phase 
program  that  would  begin  with  a  manned  orbiting  workshop,  followed  by  a 
space  station  similar  to  MORL.  The  report  emphasized  that  "definition  of  a 
real  manned  experiment  program  and  supporting  requirements  is  mandatory 
to  the  true  understanding  of  spacecraft  system  needs  and  total  flight  system 
scope."61 

Not  even  the  economical  first  phase  came  to  pass  as  conceived.  In 
late  1968,  a  spending-weary  Congress  slashed  the  budget  of  the  Apollo 
Applications  Program  to  one-third  of  the  NASA  request.  A  down-scaled 
concept,  the  Skylab  orbital  workshop,  would  be  launched  in  May  1973, 
carrying  with  it  an  experiment  package  developed  by  Langley  researchers. 
By  that  time,  however,  with  personnel  reductions  and  program  shifts 
resulting  from  severe  budget  cuts  within  NASA,  Langley  was  largely  out 
of  the  space  station  business.  When  so-called  Phase  B  Definition  Studies 
for  NASA's  space  station  program  began  in  1969,  they  were  managed  by 
the  Marshall  and  Johnson  centers.62 

Bigger  ideas  were  stealing  the  thunder  from  the  Langley  concept.  At 
Houston  in  1968,  engineers  were  working  on  plans  for  a  huge  "Space 
Base"  weighing  a  million  pounds,  with  room  for  thousands  of  pounds  of 
experiments,  and  a  crew  of  75  to  100  people.  According  to  the  plan,  the 
Space  Base  would  provide  .1  G  by  spinning  at  3.5  rpm  at  the  240-foot 
radius  of  the  living  module  and  would  operate  "on  a  permanent  basis  to 
take  advantage  of  the  economics  of  size,  centralization,  and  permanency." 
The  base  would  be  constructed  in  an  orbital  buildup  of  hardware  delivered 
by  no  less  than  three  Saturn  launches.63 

306 


Skipping  "The  Next  Logical  Step" 

Although  everyone  recognized  that  this  large  space  station  would  have 
to  come  after  the  Apollo  Applications  workshop,  Houston's  grandiose 
idea  nonetheless  had  "a  significant  effect  on  agency  planning" — and  one 
that  in  the  end  did  not  help  the  ultimate  cause  of  the  space  station 
program.64  When  Phase  B  Definition  began  in  late  1969,  with  major  con- 
tracts awarded  to  McDonnell  Douglas  and  North  American  Rockwell,  the 
notion  of  a  large  station  held  sway.  The  contractors  were  asked  to  explore 
the  feasibility  of  a  smaller  but  still  rather  large  station,  33  feet  in  diameter, 
to  be  launched  by  a  Saturn  V  and  manned  initially  by  a  crew  of  12.  The 
NASA/industry  space  station  teams  were  to  do  this  "in  concert  with  stud- 
ies of  future  large  space  bases,"  involving  crews  of  100  people  or  more,  as 
well  as  with  manned  missions  to  Mars.  Crews  for  some  of  these  space  base 
concepts  exceeded  100  and  included  plans  for  an  advanced  logistics  system, 
which  was  soon  to  be  named  the  "Space  Shuttle."65 

In  1971,  when  the  decision  was  made  to  go  forward  with  the  development 
of  the  manned  Space  Shuttle,  NASA  redirected  its  space  station  contractors 
to  consider  a  modular  design,  with  the  modules  to  be  placed  in  orbit,  not 
by  Saturns,  but  by  a  totally  reusable  shuttle.  The  purpose  of  Phase  B  from 
that  point  on,  into  1972,  was  to  define  the  modular  concepts.  A  large  space 
station  with  the  Space  Shuttle  to  assemble  and  service  it  was  now  "the 
next  logical  step"  in  NASA's  manned  space  program  following  the  Apollo 
Applications  Program  and  Skylab. 

Politics  and  budget  pressures,  however,  once  again  meant  a  missed  step. 
The  nation  had  neither  the  will  nor  the  money  for  NASA's  entire  mission 
package.  As  Howard  McCurdy  points  out  in  his  1990  analysis  The  Space 
Station  Decision:  Incremental  Politics  and  Technological  Choice,  NASA 
officials,  having  failed  to  win  the  support  of  the  Nixon  administration  for 
their  internal  long-range  plan,  decided  to  shift  their  strategy.  "Rather  than 
seek  a  comprehensive,  Apollo-style  commitment,  they  decided  to  pursue  the 
steps  in  their  plan  one  by  one."66  NASA  would  ask  first  for  an  economical 
Space  Transportation  System,  the  Shuttle,  then  they  would  ask  for  the  space 
station.  The  result,  after  Nixon  accepted  NASA's  compromise,  meant  that 
"the  next  logical  step"  would  be  skipped  once  again. 

Lost  in  Space? 

The  majority  of  Langley  researchers  involved  in  the  pioneering  space 
station  studies  of  the  early  1960s  believe  that  the  decision  not  to  develop 
and  deploy  the  MORL  was  a  major  national  mistake.  As  many  of  them 
have  asserted  in  retrospect,  the  Soviet's  tremendously  successful  MIR  space 
station  of  the  1980s  (the  spacious  follow-on  to  the  more  primitive  Salyuts 
first  launched  in  1971)  "would  prove  to  be  almost  exactly  like  what  MORL 
would  have  been."67  W.  Ray  Hook,  a  member  of  Langley 's  MORL  Studies 
Office,  expresses  the  general  sentiment  of  Langley  researchers: 


307 


Space/light  Revolution 


Skipping  over  a  space  station  for  a  second  time 
left  William  N.  Gardner,  head  of  Langley's 
MORL  office,  with  a  bitter  taste  for  his  pi- 
oneering work  of  the  1960s  and  a  judgment 
that  NASA,  unlike  NACA,  was  too  much  the 
creature  of  presidential  projects — or  the  lack  of 

them.  ^___^^^^_______ 

L-66-7229 


Our  goal  was  to  get  one  man  in  space  for  one  year.  That  was  the  simple  objective. 
Of  course,  it  has  since  gotten  a  lot  more  complicated.  I  have  often  thought  that  if 
we'd  stuck  with  that  simple-minded  objective,  we  would  have,  thirty  years  later,  one 

fi8 

man  in  space  for  one  year,  which  we  don't. 

If  the  modular  MORL  had  been  ready  for  deployment  on  the  heels  of  Skylab, 
as  MIR  was  ready  to  go  after  the  Salyuts,  the  United  States  like  the  Soviets 
would  have  amassed  countless  man-hours  in  space  and  conducted  numerous 
useful  experiments.  If  the  country  had  supported  MORL,  it  might  have  been 
easier  to  design  and  justify  Space  Station  Freedom,  and  instead  of  being  in 
the  present  position  of  considering  the  purchase  of  a  MIR  from  the  former 
Soviet  Union  and  proceeding  toward  an  international  space  station,  Alpha, 
the  United  States  might  today  be  operating  its  own  station. 

The  most  bitter  among  Langley  space  station  enthusiasts  feel  that  the 
decisions  regarding  the  station  were  not  only  mistakes  but  also  symptoms  of 
a  basic  flaw  in  NASA's  organizational  character.  "It  finally  dawned  on 
me,"  explains  William  Gardner,  the  head  of  the  MORL  Studies  Office, 
"that  NASA  wasn't  intended  to  be  a  real  federal  agency."  NASA  did 
not  enjoy  a  long-term  goal  like  the  former  NACA — an  agency  designed 
"to  supervise  and  direct  the  scientific  study  of  the  problems  of  flight  with 
a  view  to  their  practical  solution,"  or  even  like  the  FAA,  whose  job  was 
to  make  air  travel  effective  and  safe.  "NASA  was  just  a  project  of  the 
presidential  administration,"  and  under  Presidents  Kennedy,  Johnson,  and 
Nixon,  the  project  was  "just  to  put  a  man  on  the  moon."  "We  do  spectacular 
things  when  the  Administration  wants  spectacular  things  done,"  Gardner 
challenges,  and  when  it  does  not,  "we  don't  really  have  .a  mandate."69 


308 


Skipping  "The  Next  Logical  Step" 

But  Langley  researchers,  out  of  technological  conviction  or  political 
naivete,  or  both,  kept  working  on  the  space  station  as  if  they  had  such 
a  mandate.  The  effort  was  not  in  vain.  In  persisting  with  their  design  and 
redesign  of  space  stations  and  seeking  to  understand  how  humans  could  live 
and  work  in  space,  they  contributed  to  the  successes  of  Apollo,  Skylab,  and 
the  Space  Shuttle,  and  they  laid  a  solid  foundation  upon  which  to  build 
when  NASA  in  the  early  1980s  created  a  new  Space  Station  Task  Force 
and  once  again  began  examining  the  program  options  for  "the  next  logical 
step" — a  step  that  may  in  its  own  time  be  skipped  over  for  something  else. 
In  the  wake  of  the  spaceflight  revolution,  it  would  take  all  the  running  space 
station  researchers  could  do,  just  to  keep  in  the  same  place. 


309 


10 


To  Behold  the  Moon: 
The  Lunar  Orbiter  Project 


But,  Cliff,  you  said  we  weren't  going  to  improvise  like 
this  .  .  . 

But,  listen  to  what  I  say  now!  We  've  worked  out  the 
numbers.  It's  worth  the  risk. 

— Conversation  between  Boeing  engineer 
project  manager  Robert  J.  Helberg 
and  Clifford  H.  Nelson,  head  of 
Langley's  Lunar  Orbiter  Project  Office, 
concerning  a  change  in  the  mission 
plan  for  Lunar  Orbiter  I. 


The  bold  plan  for  an  Apollo  mission  based  on  LOR  held  the  promise 
of  landing  on  the  moon  by  1969,  but  it  presented  many  daunting  technical 
difficulties.  Before  NASA  could  dare  attempt  any  type  of  lunar  landing, 
it  had  to  learn  a  great  deal  more  about  the  destination.  Although  no  one 
believed  that  the  moon  was  made  of  green  cheese,  some  lunar  theories  of 
the  early  1960s  seemed  equally  fantastic.  One  theory  suggested  that  the 
moon  was  covered  by  a  layer  of  dust  perhaps  50  feet  thick.  If  this  were  true, 
no  spacecraft  would  be  able  to  safely  land  on  or  take  off  from  the  lunar 
surface.  Another  theory  claimed  that  the  moon's  dust  was  not  nearly  so 
thick  but  that  it  possessed  an  electrostatic  charge  that  would  cause  it  to 
stick  to  the  windows  of  the  lunar  landing  vehicle,  thus  making  it  impossible 
for  the  astronauts  to  see  out  as  they  landed.  Cornell  University  astronomer 
Thomas  Gold  warned  that  the  moon  might  even  be  composed  of  a  spongy 
material  that  would  crumble  upon  impact.1 

At  Langley,  Dr.  Leonard  Roberts,  a  British  mathematician  in  Clint 
Brown's  Theoretical  Mechanics  Division,  pondered  the  riddle  of  the  lunar 

311 


Spaceflight  Revolution 

surface  and  drew  an  equally  pessimistic  conclusion.  Roberts  speculated 
that  because  the  moon  was  millions  of  years  old  and  had  been  constantly 
bombarded  without  the  protection  of  an  atmosphere,  its  surface  was  most 
likely  so  soft  that  any  vehicle  attempting  to  land  on  it  would  sink  and  be 
buried  as  if  it  had  landed  in  quicksand.  After  the  president's  commitment 
to  a  manned  lunar  landing  in  1961,  Roberts  began  an  extensive  three-year 
research  program  to  show  just  what  would  happen  if  an  exhaust  rocket 
blasted  into  a  surface  of  very  thick  powdered  sand.  His  analysis  indicated 
that  an  incoming  rocket  would  throw  up  a  mountain  of  sand,  thus  creating 
a  big  rim  all  the  way  around  the  outside  of  the  landed  spacecraft.  Once 
the  spacecraft  settled,  this  huge  bordering  volume  of  sand  would  collapse, 
completely  engulf  the  spacecraft,  and  kill  its  occupants.2 

Telescopes  revealed  little  about  the  nature  of  the  lunar  surface.  Not  even 
the  latest,  most  powerful  optical  instruments  could  see  through  the  earth's 
atmosphere  well  enough  to  resolve  the  moon's  detailed  surface  features. 
Even  an  object  the  size  of  a  football  stadium  would  not  show  up  on  a 
telescopic  photograph,  and  enlarging  the  photograph  would  only  increase 
the  blur.  To  separate  fact  from  fiction  and  obtain  the  necessary  information 
about  the  craters,  crevices,  and  jagged  rocks  on  the  lunar  surface,  NASA 
would  have  to  send  out  automated  probes  to  take  a  closer  look. 

The  first  of  these  probes  took  off  for  the  moon  in  January  1962  as  part 
of  a  NASA  project  known  as  Ranger.  A  small  800-pound  spacecraft  was 
to  make  a  "hard  landing,"  crashing  to  its  destruction  on  the  moon.  Before 
Ranger  crashed,  however,  its  on-board  multiple  television  camera  payload 
was  to  send  back  close  views  of  the  surface — views  far  more  detailed  than 
any  captured  by  a  telescope.  Sadly,  the  first  six  Ranger  probes  were  not 
successful.  Malfunctions  of  the  booster  or  failures  of  the  launch-vehicle 
guidance  system  plagued  the  first  three  attempts;  malfunctions  of  the 
spacecraft  itself  hampered  the  fourth  and  fifth  probes;  and  the  primary 
experiment  could  not  take  place  during  the  sixth  Ranger  attempt  because 
the  television  equipment  would  not  transmit.  Although  these  incomplete 
missions  did  provide  some  extremely  valuable  high- resolution  photographs, 
as  well  as  some  significant  data  on  the  performance  of  Ranger's  systems, 
in  total  the  highly  publicized  record  of  failures  embarrassed  NASA  and 
demoralized  the  Ranger  project  managers  at  JPL.  Fortunately,  the  last 
three  Ranger  flights  in  1964  and  1965  were  successful.  These  flights  showed 
that  a  lunar  landing  was  possible,  but  the  site  would  have  to  be  carefully 
chosen  to  avoid  craters  and  big  boulders.3 

JPL  managed  a  follow-on  project  to  Ranger  known  as  Surveyor.  Despite 
failures  and  serious  schedule  delays,  between  May  1966  and  January  1968, 
six  Surveyor  spacecraft  made  successful  soft  landings  at  predetermined 
points  on  the  lunar  surface.  From  the  touchdown  dynamics,  surface- 
bearing  strength  measurements,  and  eye-level  television  scanning  of  the 
local  surface  conditions,  NASA  learned  that  the  moon  could  easily  support 
the  impact  and  the  weight  of  a  small  lander.  Originally,  NASA  also 

312 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

planned  for  (and  Congress  had  authorized)  a  second  type  of  Surveyor 
spacecraft,  which  instead  of  making  a  soft  landing  on  the  moon,  was  to 
be  equipped  for  high-resolution  stereoscopic  film  photography  of  the  moon's 
surface  from  lunar  orbit  and  for  instrumented  measurements  of  the  lunar 
environment.  However,  this  second  Surveyor  or  "Surveyor  Orbiter"  did  not 
materialize.  The  staff  and  facilities  of  JPL  were  already  overburdened  with 
the  responsibilities  for  Ranger  and  "Surveyor  Lander";  they  simply  could 
not  take  on  another  major  spaceflight  project.4 

In  1963,  NASA  scrapped  its  plans  for  a  Surveyor  Orbiter  and  turned 
its  attention  to  a  lunar  orbiter  project  that  would  not  use  the  Surveyor 
spacecraft  system  or  the  Surveyor  launch  vehicle,  Centaur.  Lunar  Orbiter 
would  have  a  new  spacecraft  and  use  the  Atlas- Agena  D  to  launch  it  into 
space.  Unlike  the  preceding  unmanned  lunar  probes,  which  were  originally 
designed  for  general  scientific  study,  Lunar  Orbiter  was  conceived  after  a 
manned  lunar  landing  became  a  national  commitment.  The  project  goal 
from  the  start  was  to  support  the  Apollo  mission.  Specifically,  Lunar  Orbiter 
was  designed  to  provide  information  on  the  lunar  surface  conditions  most 
relevant  to  a  spacecraft  landing.  This  meant,  among  other  things,  that 
its  camera  had  to  be  sensitive  enough  to  capture  subtle  slopes  and  minor 
protuberances  and  depressions  over  a  broad  area  of  the  moon's  front  side. 
As  an  early  working  group  on  the  requirements  of  the  lunar  photographic 
mission  had  determined,  Lunar  Orbiter  had  to  allow  the  identification  of  45- 
meter  objects  over  the  entire  facing  surface  of  the  moon,  4.5-meter  objects 
in  the  "Apollo  zone  of  interest,"  and  1.2-meter  objects  in  all  the  proposed 
landing  areas.5 

Five  Lunar  Orbiter  missions  took  place.  The  first  launch  occurred  in 
August  1966  within  two  months  of  the  initial  target  date.  The  next  four 
Lunar  Orbiters  were  launched  on  schedule;  the  final  mission  was  completed 
in  August  1967,  barely  a  year  after  the  first  launch.  NASA  had  planned  five 
flights  because  mission  reliability  studies  had  indicated  that  five  might  be 
necessary  to  achieve  even  one  success.  However,  all  five  Lunar  Orbiters  were 
successful,  and  the  prime  objective  of  the  project,  which  was  to  photograph 
in  detail  all  the  proposed  landing  sites,  was  met  in  three  missions.  This 
meant  that  the  last  two  flights  could  be  devoted  to  photographic  exploration 
of  the  rest  of  the  lunar  surface  for  more  general  scientific  purposes.  The  final 
cost  of  the  program  was  not  slight:  it  totaled  $163  million,  which  was  more 
than  twice  the  original  estimate  of  $77  million.  That  increase,  however, 
compares  favorably  with  the  escalation  in  the  price  of  similar  projects,  such 
as  Surveyor,  which  had  an  estimated  cost  of  $125  million  and  a  final  cost  of 
$469  million. 

In  retrospect,  Lunar  Orbiter  must  be,  and  rightfully  has  been,  regarded 
as  an  unqualified  success.  For  the  people  and  institutions  responsible, 
the  project  proved  to  be  an  overwhelmingly  positive  learning  experience 
on  which  greater  capabilities  and  ambitions  were  built.  For  both  the 
prime  contractor,  the  Boeing  Company,  a  world  leader  in  the  building  of 

313 


Space/light  Revolution 


L-66-1571 

The  most  successful  of  the  pre-Apollo  probes,  Lunar  Orbiter  mapped  the  equatorial 
regions  of  the  moon  and  gave  NASA  the  data  it  needed  to  pinpoint  ideal  landing 
spots. 


314 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

airplanes,  and  the  project  manager,  Langley  Research  Center,  a  premier 
aeronautics  laboratory,  involvement  in  Lunar  Orbiter  was  a  turning  point. 
The  successful  execution  of  a  risky  enterprise  became  proof  positive  that 
they  were  more  than  capable  of  moving  into  the  new  world  of  deep  space. 
For  many  observers  as  well  as  for  the  people  who  worked  on  the  project, 
Lunar  Orbiter  quickly  became  a  model  of  how  to  handle  a  program  of  space 
exploration.  Its  successful  progress  demonstrated  how  a  clear  and  discrete 
objective,  strong  leadership,  and  positive  person-to-person  communication 
skills  can  keep  a  project  on  track  from  start  to  finish.6 

Many  people  inside  the  American  space  science  community  believed  that 
neither  Boeing  nor  Langley  was  capable  of  managing  a  project  like  Lunar 
Orbiter  or  of  supporting  the  integration  of  first-rate  scientific  experiments 
and  space  missions.  After  NASA  headquarters  announced  in  the  summer 
of  1963  that  Langley  would  manage  Lunar  Orbiter,  more  than  one  space 
scientist  was  upset.  Dr.  Harold  C.  Urey,  a  prominent  scientist  from  the 
University  of  California  at  San  Diego,  wrote  a  letter  to  Administrator  James 
Webb  asking  him,  "How  in  the  world  could  the  Langley  Research  Center, 
which  is  nothing  more  than  a  bunch  of  plumbers,  manage  this  scientific 
program  to  the  moon?"7 

Urey's  questioning  of  Langley's  competency  was  part  of  an  unfolding 
debate  over  the  proper  place  of  general  scientific  objectives  within  NASA's 
spaceflight  programs.  The  U.S.  astrophysics  community  and  Dr.  Homer  E. 
Newell 's  Office  of  Space  Sciences  at  NASA  headquarters  wanted  "quality 
science"  experiments  incorporated  into  every  space  mission,  but  this  caused 
problems.  Once  the  commitment  had  been  made  to  a  lunar  landing  mission, 
NASA  had  to  decide  which  was  more  important:  gathering  broad  scientific 
information  or  obtaining  data  required  for  accomplishing  the  lunar  landing 
mission.  Ideally,  both  goals  could  be  incorporated  in  a  project  without  one 
compromising  the  other,  but  when  that  seemed  impossible,  one  of  the  two 
had  to  be  given  priority.  The  requirements  of  the  manned  mission  usually 
won  out.  For  Ranger  and  Surveyor,  projects  involving  dozens  of  outside 
scientists  and  the  large  and  sophisticated  Space  Science  Division  at  JPL, 
that  meant  that  some  of  the  experiments  would  turn  out  to  be  less  extensive 
than  the  space  scientists  wanted.8  For  Lunar  Orbiter,  a  project  involving 
only  a  few  astrogeologists  at  the  U.S.  Geological  Survey  and  a  very  few  space 
scientists  at  Langley,  it  meant,  ironically,  that  the  primary  goal  of  serving 
Apollo  would  be  achieved  so  quickly  that  general  scientific  objectives  could 
be  included  in  its  last  two  missions. 


The  "Moonball"  Experiment 

Langley  management  had  entered  the  fray  between  science  and  project 
engineering  during  the  planning  for  Project  Ranger.  At  the  first  Senior 
Council  meeting  of  the  Office  of  Space  Sciences  (soon  to  be  renamed 

315 


Space/light  Revolution 

the  Office  of  Space  Sciences  and  Applications  [OSS A])  held  at  NASA 
headquarters  on  7  June  1962,  Langley  Associate  Director  Charles  Donlan 
had  questioned  the  priority  of  a  scientific  agenda  for  the  agency's  proposed 
unmanned  lunar  probes  because  a  national  commitment  had  since  been 
made  to  a  manned  lunar  landing.  The  initial  requirements  for  the  probes 
had  been  set  long  before  Kennedy's  announcement,  and  therefore,  Donlan 
felt  NASA  needed  to  rethink  them.  Based  on  his  experience  at  Langley  and 
with  Gilruth's  STG,  Donlan  knew  that  the  space  science  people  could  be 
"rather  unbending"  about  adjusting  experiments  to  obtain  "scientific  data 
which  would  assist  the  manned  program."  What  needed  to  be  done  now, 
he  felt,  was  to  turn  the  attention  of  the  scientists  to  exploration  that  would 
have  more  direct  applications  to  the  Apollo  lunar  landing  program.9 

Donlan  was  distressed  specifically  by  the  Office  of  Space  Sciences'  recent 
rejection  of  a  lunar  surface  experiment  proposed  by  a  penetrometer  feasi- 
bility study  group  at  Langley.  This  small  group,  consisting  of  half  a  dozen 
people  from  the  Dynamic  Loads  and  Instrument  Research  divisions,  had 
devised  a  spherical  projectile,  dubbed  "Moonball,"  that  was  equipped  with 
accelerometers  capable  of  transmitting  acceleration  versus  time  signatures 
during  impact  with  the  lunar  surface.  With  these  data,  researchers  could 
determine  the  hardness,  texture,  and  load-bearing  strength  of  possible  lunar 
landing  sites.  The  group  recommended  that  Moonball  be  flown  as  part  of 
the  follow-on  to  Ranger.10 

A  successful  landing  of  an  intact  payload  required  that  the  landing  loads 
not  exceed  the  structural  capabilities  of  the  vehicle  and  that  the  vehicle 
make  its  landing  in  some  tenable  position  so  it  could  take  off  again.  Both  of 
these  requirements  demanded  a  knowledge  of  basic  physical  properties  of  the 
surface  material,  particularly  data  demonstrating  its  hardness  or  resistance 
to  penetration.  In  the  early  1960s,  these  properties  were  still  unknown,  and 
the  Langley  penetrometer  feasibility  study  group  wanted  to  identify  them. 
Without  the  information,  any  design  of  Apollo's  lunar  lander  would  have  to 
be  based  on  assumed  surface  characteristics.11 

In  the  opinion  of  the  Langley  penetrometer  group,  its  lunar  surface 
hardness  experiment  would  be  of  "general  scientific  interest,"  but  it  would, 
more  importantly,  provide  "timely  engineering  information  important  to 
the  design  of  the  Apollo  manned  lunar  landing  vehicle."12  Experts  at  JPL, 
however,  questioned  whether  surface  hardness  was  an  important  criterion 
for  any  experiment  and  argued  that  "the  determination  of  the  terrain  was 
more  important,  particularly  for  a  horizontal  landing."13  In  the  end,  the 
Office  of  Space  Sciences  rejected  the  Langley  idea  in  favor  of  making  further 
seismometer  experiments,  which  might  tell  scientists  something  basic  about 
the  origins  of  the  moon  and  its  astrogeological  history.* 


Later  in  Apollo  planning,  engineers  at  the  Manned  Spacecraft  Center  in  Houston  thought  that 
deployment  of  a  penetrometer  from  the  LEM  during  its  final  approach  to  landing  would  prove  useful.  The 
penetrometer  would  "sound"  the  anticipated  target  and  thereby  determine  whether  surface  conditions 

316 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-67-8553 

Associate  Director   Charles   J.    Donlan  understood  that  the   requirements   of  the 
manned  lunar  landing  took  priority  over  pure  science  experiments. 


For  engineer  Donlan,  representing  a  research  organization  like  Langley 
dominated  by  engineers  and  by  their  quest  for  practical  solutions  to  applied 
problems,  this  rejection  seemed  a  mistake.  The  issue  came  down  to  what 
NASA  needed  to  know  now.  That  might  have  been  science  before  Kennedy's 
commitment,  but  it  definitely  was  not  science  after  it.  In  Donlan's  view, 
Langley's  rejected  approach  to  lunar  impact  studies  had  been  the  correct 
one.  The  consensus  at  the  first  Senior  Council  meeting,  however,  was  that 
"pure  science  experiments  will  be  able  to  provide  the  engineering  answers 
for  Project  Apollo." 14 

Over  the  next  few  years,  the  engineering  requirements  for  Apollo  would 
win  out  almost  totally.  As  historian  R.  Cargill  Hall  explains  in  his  story 
of  Project  Ranger,  a  "melding"  of  interests  occurred  between  the  Office 


were  conducive  to  landing.  Should  surface  conditions  prove  unsatisfactory,  the  LEM  could  be  flown  to 
another  spot  or  the  landing  could  be  aborted.  In  the  end,  NASA  deemed  the  experiment  unnecessary. 
What  the  Surveyor  missions  found  out  about  the  nature  of  the  lunar  soil  (that  it  resembled  basalt  and 
had  the  consistency  of  damp  sand)  made  NASA  so  confident  about  the  hardness  of  the  surface  that  it 
decided  this  penetrometer  experiment  could  be  deleted.  For  more  information,  see  Ivan  D.  Ertel  and 
Roland  W.  Newkirk,  The  Apollo  Spacecraft:  A  Chronology,  vol.  4,  NASA  SP-4009  (Washington,  1978), 
p.  24. 

317 


Space/light  Revolution 


L-63-1777 


L-67-4028 

Langley  gathered  information  specifically  for  the  accomplishment  of  Apollo.  Top,  a 
Langley  engineer  monitors  the  structural  dynamics  of  a  simulated  lunar  landing  in 
early  1963.  Bottom,  Lunar  Orbiter  III  maps  a  potential  Apollo  landing  site.  The 
large  crater  is  Kepler,  which  is  30  miles  across. 

318 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

of  Space  Sciences  and  the  Office  of  Manned  Space  Flight  followed  by  a 
virtually  complete  subordination  of  the  scientific  priorities  originally  built 
into  the  unmanned  projects.  Those  priorities,  as  important  as  they  were, 
"quite  simply  did  not  rate"  with  Apollo  in  importance.15 


Initiating  Lunar  Orbiter 

The  sensitive  camera  eyes  of  the  Lunar  Orbiter  spacecraft  carried  out  a 
vital  reconnaissance  mission  in  support  of  the  Apollo  program.  Although 
NASA  designed  the  project  to  provide  scientists  with  quantitative  infor- 
mation about  the  moon's  gravitational  field  and  the  dangers  of  micro- 
meteorites  and  solar  radiation  in  the  vicinity  of  the  lunar  environment,  the 
primary  objective  of  Lunar  Orbiter  was  to  fly  over  and  photograph  the  best 
landing  sites  for  the  Apollo  spacecraft.  NASA  suspected  that  it  might  have 
enough  information  about  the  lunar  terrain  to  land  astronauts  safely  with- 
out the  detailed  photographic  mosaics  of  the  lunar  surface  compiled  from 
the  orbiter  flights,  but  certainly  landing  sites  could  be  pinpointed  more  ac- 
curately with  the  help  of  high-resolution  photographic  maps.  Lunar  Orbiter 
would  even  help  to  train  the  astronauts  for  visual  recognition  of  the  lunar 
topography  and  for  last-second  maneuvering  above  it  before  touchdown. 

Langley  had  never  managed  a  deep-space  flight  project  before,  and 
Director  Floyd  Thompson  was  not  sure  that  he  wanted  to  take  on  the  burden 
of  responsibility  when  Oran  Nicks,  the  young  director  of  lunar  and  planetary 
programs  in  Homer  Newell's  Office  of  Space  Sciences,  came  to  him  with  the 
idea  early  in  1963.  Along  with  Newell's  deputy,  Edgar  M.  Cortright,  Nicks 
was  the  driving  force  behind  the  orbiter  mission  at  NASA  headquarters. 
Cortright,  however,  first  favored  giving  the  project  to  JPL  and  using 
Surveyor  Orbiter  and  the  Hughes  Aircraft  Company,  which  was  the  prime 
contractor  for  Surveyor  Lander.  Nicks  disagreed  with  this  plan  and  worked 
to  persuade  Cortright  and  others  that  he  was  right.  In  Nicks'  judgment,  JPL 
had  more  than  it  could  handle  with  Ranger  and  Surveyor  Lander  and  should 
not  have  anything  else  "put  on  its  plate,"  certainly  not  anything  as  large 
as  the  Lunar  Orbiter  project.  NASA  Langley,  on  the  other  hand,  besides 
having  a  reputation  for  being  able  to  handle  a  variety  of  aerospace  tasks,  had 
just  lost  the  STG  to  Houston  and  so,  Nicks  thought,  would  be  eager  to  take 
on  the  new  challenge  of  a  lunar  orbiter  project.  Nicks  worked  to  persuade 
Cortright  that  distributing  responsibilities  and  operational  programs  among 
the  NASA  field  centers  would  be  "a  prudent  management  decision."  NASA 
needed  balance  among  its  research  centers.  To  ensure  NASA's  future  in 
space,  headquarters  must  assign  to  all  its  centers  challenging  endeavors  that 
would  stimulate  the  development  of  "new  and  varied  capabilities."16 


319 


Space/light  Revolution 

Cortright  was  persuaded  and  gave  Nicks  permission  to  approach  Floyd 
Thompson.*  This  Nicks  did  on  2  January  1963,  during  a  Senior  Council 
meeting  of  the  Office  of  Space  Sciences  at  Cape  Canaveral.  Nicks  asked 
Thompson  whether  Langley  "would  be  willing  to  study  the  feasibility  of 
undertaking  a  lunar  photography  experiment,"  and  Thompson  answered 
cautiously  that  he  would  ask  his  staff  to  consider  the  idea.17 

The  historical  record  does  not  tell  us  much  about  Thompson's  personal 
thoughts  regarding  taking  on  Lunar  Orbiter.  But  one  can  infer  from  the 
evidence  that  Thompson  had  mixed  feelings,  not  unlike  those  he  experienced 
about  supporting  the  STG.  The  Langley  director  would  not  only  give  Nicks 
a  less  than  straightforward  answer  to  his  question  but  also  would  think 
about  the  offer  long  and  hard  before  committing  the  center.  Thompson 
invited  several  trusted  staff  members  to  share  their  feelings  about  assuming 
responsibility  for  the  project.  For  instance,  he  went  to  Clint  Brown,  by 
then  one  of  his  three  assistant  directors  for  research,  and  asked  him  what 
he  thought  Langley  should  do.  Brown  told  him  emphatically  that  he  did 
not  think  Langley  should  take  on  Lunar  Orbiter.  An  automated  deep-space 
project  would  be  difficult  to  manage  successfully.  The  Lunar  Orbiter  would 
be  completely  different  from  the  Ranger  and  Surveyor  spacecraft  and  being 
a  new  design,  would  no  doubt  encounter  many  unforeseen  problems.  Even 
if  it  were  done  to  everyone's  satisfaction — and  the  proposed  schedule  for  the 
first  launches  sounded  extremely  tight — Langley  would  probably  handicap 
its  functional  research  divisions  to  give  the  project  all  the  support  that  it 
would  need.  Projects  devoured  resources.  Langley  staff  had  learned  this 
firsthand  from  its  experience  with  the  STG.  Most  of  the  work  for  Lunar 
Orbiter  would  rest  in  the  management  of  contracts  at  industrial  plants  and 
in  the  direction  of  launch  and  mission  control  operations  at  Cape  Canaveral 
and  Pasadena.  Brown,  for  one,  did  not  want  to  be  involved.18 

But  Thompson  decided,  in  what  Brown  now  calls  his  director's  "greater 
wisdom,"  that  the  center  should  accept  the  job  of  managing  the  project. 
Some  researchers  in  Brown's  own  division  had  been  proposing  a  Langley- 
directed  photographic  mission  to  the  moon  for  some  time,  and  Thompson, 
too,  was  excited  by  the  prospect.19  Furthermore,  the  revamped  Lunar 
Orbiter  was  not  going  to  be  a  space  mission  seeking  general  scientific 
knowledge  about  the  moon.  It  was  going  to  be  a  mission  directly  in 
support  of  Apollo,  and  this  meant  that  engineering  requirements  would  be 
primary.  Langley  staff  preferred  that  practical  orientation;  their  past  work 
often  resembled  projects  on  a  smaller  scale.  Whether  the  "greater  wisdom" 
stemmed  from  Thompson's  own  powers  of  judgment  is  still  not  certain. 
Some  informed  Langley  veterans,  notably  Brown,  feel  that  Thompson  must 


Edgar  Cortright  and  Oran  Nicks  would  come  to  have  more  than  a  passing  familiarity  with  the 
capabilities  of  Langley  Research  Center.  In  1968,  NASA  would  name  Cortright  to  succeed  Thompson 
as  the  center's  director.  Shortly  thereafter,  Cortright  named  Nicks  as  his  deputy  director.  Both  men 
then  stayed  at  the  center  into  the  mid-1970s. 

320 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

have  also  received  some  strongly  stated  directive  from  NASA  headquarters 
that  said  Langley  had  no  choice  but  to  take  on  the  project. 

Whatever  was  the  case  in  the  beginning,  Langley  management  soon 
welcomed  Lunar  Orbiter.  It  was  a  chance  to  prove  that  they  could  manage 
a  major  undertaking.  Floyd  Thompson  personally  oversaw  many  aspects 
of  the  project  and  for  more  than  four  years  did  whatever  he  could  to  make 
sure  that  Langley's  functional  divisions  supported  it  fully.  Through  most  of 
this  period,  he  would  meet  every  Wednesday  morning  with  the  top  people  in 
the  project  office  to  hear  about  the  progress  of  their  work  and  offer  his  own 
ideas.  As  one  staff  member  recalls,  "I  enjoyed  these  meetings  thoroughly. 
[Thompson  was]  the  most  outstanding  guy  I've  ever  met,  a  tremendously 
smart  man  who  knew  what  to  do  and  when  to  do  it."20 

Throughout  the  early  months  of  1963,  Langley  worked  with  its  counter- 
parts at  NASA  headquarters  to  establish  a  solid  and  cooperative  working 
relationship  for  Lunar  Orbiter.  The  center  began  to  draw  up  preliminary 
specifications  for  a  lightweight  orbiter  spacecraft  and  for  the  vehicle  that 
would  launch  it  (already  thought  to  be  the  Atlas- Agena  D).  While  Langley 
personnel  were  busy  with  that,  TRW's  Space  Technologies  Laboratories 
(STL)  of  Redondo  Beach,  California,  was  conducting  a  parallel  study  of 
a  lunar  orbiter  photographic  spacecraft  under  contract  to  NASA  head- 
quarters. Representatives  from  STL  reported  on  this  work  at  meetings  at 
Langley  on  25  February  and  5  March  1963.  Langley  researchers  reviewed  the 
contractor's  assessment  and  found  that  STL's  estimates  of  the  chances  for 
mission  success  closely  matched  their  own.  If  five  missions  were  attempted, 
the  probability  of  achieving  one  success  was  93  percent.  The  probability  of 
achieving  tv/o  was  81  percent.  Both  studies  confirmed  that  a  lunar  orbiter 
system  using  existing  hardware  would  be  able  to  photograph  a  landed  Sur- 
veyor and  would  thus  be  able  to  verify  the  conditions  of  that  possible  Apollo 
landing  site.  The  independent  findings  concluded  that  the  Lunar  Orbiter 
project  could  be  done  successfully  and  should  be  done  quickly  because  its 
contribution  to  the  Apollo  program  would  be  great.21 

Project  Management 

With  the  exception  of  its  involvement  in  the  X-series  research  airplane 
programs  at  Muroc,  Langley  had  not  managed  a  major  project  during  the 
period  of  the  NACA.  As  a  NASA  center,  Langley  would  have  to  learn 
to  manage  projects  that  involved  contractors,  subcontractors,  other  NASA 
facilities,  and  headquarters — a  tall  order  for  an  organization  used  to  doing  all 
its  work  in-house  with  little  outside  interference.  Only  three  major  projects 
were  assigned  to  Langley  in  the  early  1960s:  Scout,  in  1960;  Fire,  in  1961; 
and  Lunar  Orbiter,  in  1963.  Project  Mercury  and  Little  Joe,  although 
heavily  supported  by  Langley,  had  been  managed  by  the  independent  STG, 
and  Project  Echo,  although  managed  by  Langley  for  a  while,  eventually  was 
given  to  Goddard  to  oversee. 

321 


Space/light  Revolution 

To  prepare  for  Lunar  Orbiter  in  early  1963,  Langley  management  re- 
viewed what  the  center  had  done  to  initiate  the  already  operating  Scout  and 
Fire  projects.  It  also  tried  to  learn  from  JPL  about  inaugurating  paperwork 
for,  and  subsequent  management  of,  Projects  Ranger  and  Surveyor.  After 
these  reviews,  Langley  felt  ready  to  prepare  the  formal  documents  required 
by  NASA  for  the  start-up  of  the  project.22 

As  Langley  prepared  for  Lunar  Orbiter,  NASA's  policies  and  procedures 
for  project  management  were  changing.  In  October  1962,  spurred  on  by  its 
new  top  man,  James  Webb,  the  agency  had  begun  to  implement  a  series 
of  structural  changes  in  its  overall  organization.  These  were  designed  to 
improve  relations  between  headquarters  and  the  field  centers,  an  area  of 
fundamental  concern.  Instead  of  managing  the  field  centers  through  the 
Office  of  Programs,  as  had  been  the  case,  NASA  was  moving  them  under 
the  command  of  the  headquarters  program  directors.  For  Langley,  this 
meant  direct  lines  of  communication  with  the  OART  and  the  OSSA.  By 
the  end  of  1963,  a  new  organizational  framework  was  in  place  that  allowed 
for  more  effective  management  of  NASA  projects. 

In  early  March  1963,  as  part  of  Webb's  reform,  NASA  headquarters  is- 
sued an  updated  version  of  General  Management  Instruction  4-1-1.  This 
revised  document  established  formal  guidelines  for  the  planning  and  man- 
agement of  a  project.  Every  project  was  supposed  to  pass  through  four 
preliminary  stages:  (1)  Project  Initiation,  (2)  Project  Approval,  (3)  Project 
Implementation,  and  (4)  Organization  for  Project  Management.23  Each  step 
required  the  submission  of  a  formal  document  for  headquarters'  approval. 

From  the  beginning,  everyone  involved  with  Lunar  Orbiter  realized  that 
it  had  to  be  a  fast-track  project.  In  order  to  help  Apollo,  everything  about 
it  had  to  be  initiated  quickly  and  without  too  much  concern  about  the  letter 
of  the  law  in  the  written  procedures.  Consequently,  although  no  step  was 
to  be  taken  without  first  securing  approval  for  the  preceding  step,  Langley 
initiated  the  paperwork  for  all  four  project  stages  at  the  same  time.  This 
same  no-time-to-lose  attitude  ruled  the  schedule  for  project  development. 
All  aspects  had  to  be  developed  concurrently.  Launch  facilities  had  to 
be  planned  at  the  same  time  that  the  design  of  the  spacecraft  started. 
The  photographic,  micrometeoroid,  and  selenodetic  experiments  had  to  be 
prepared  even  before  the  mission  operations  plan  was  complete.  Everything 
proceeded  in  parallel:  the  development  of  the  spacecraft,  the  mission 
design,  the  operational  plan  and  preparation  of  ground  equipment,  the  crea- 
tion of  computer  programs,  as  well  as  a  testing  plan.  About  this  parallel 
development,  Donald  H.  Ward,  a  key  member  of  Langley's  Lunar  Orbiter 
project  team,  remarked,  "Sometimes  this  causes  undoing  some  mistakes, 
but  it  gets  to  the  end  product  a  lot  faster  than  a  serial  operation  where 
you  design  the  spacecraft  and  then  the  facilities  to  support  it."24  Using  the 
all-at-once  approach,  Langley  put  Lunar  Orbiter  in  orbit  around  the  moon 
only  27  months  after  signing  with  the  contractor. 


322 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


Israel  Taback  (left)  and  Clifford  H. 
Nelson  (right),  head  of  LOPO, 
ponder  the  intricacies  of  the  space- 
craft design. 


L-64-7909 


On  11  September  1963,  Director  Floyd  Thompson  formally  established 
the  Lunar  Orbiter  Project  Office  (LOPO)  at  Langley,  a  lean  organization 
of  just  a  few  people  who  had  been  at  work  on  Lunar  Orbiter  since  May. 
Thompson  named  Clifford  H.  Nelson  as  the  project  manager.  An  NACA 
veteran  and  head  of  the  Measurements  Research  Branch  of  IRD,  Nelson  was 
an  extremely  bright  engineer.  He  had  served  as  project  engineer  on  several 
flight  research  programs,  and  Thompson  believed  that  he  showed  great 
promise  as  a  technical  manager.  He  worked  well  with  others,  and  Thompson 
knew  that  skill  in  interpersonal  relations  would  be  essential  in  managing 
Lunar  Orbiter  because  so  much  of  the  work  would  entail  interacting  with 
contractors. 

To  help  Nelson,  Thompson  originally  reassigned  eight  people  to  LOPO: 
engineers  Israel  Taback,  Robert  Girouard,  William  I.  Watson,  Gerald 
Brewer,  John  B.  Graham,  Edmund  A.  Brummer,  financial  accountant 
Robert  Fairburn,  and  secretary  Anna  Plott.  This  group  was  far  smaller 
than  the  staff  of  100  originally  estimated  for  this  office.  The  most  important 
technical  minds  brought  in  to  participate  came  from  either  IRD  or  from  the 
Applied  Materials  and  Physics  Division,  which  was  the  old  PARD.  Taback 
was  the  experienced  and  sage  head  of  the  Navigation  and  Guidance  Branch 
of  IRD;  Brummer,  an  expert  in  telemetry,  also  came  from  IRD;  and  two 

323 


Spaceflight  Revolution 

new  Langley  men,  Graham  and  Watson,  were  brought  in  to  look  over  the 
integration  of  mission  operations  and  spacecraft  assembly  for  the  project.  A 
little  later  IRD's  talented  Bill  Boyer  also  joined  the  group  as  flight  operations 
manager,  as  did  the  outstanding  mission  analyst  Norman  L.  Crabill,  who 
had  just  finished  working  on  Project  Echo.  All  four  of  the  NACA  veterans 
were  serving  as  branch  heads  at  the  time  of  their  assignment  to  LOPO.  This 
is  significant  given  that  individuals  at  that  level  of  authority  and  experience 
are  often  too  entrenched  and  concerned  about  further  career  development 
to  take  a  temporary  assignment  on  a  high-risk  project.  The  LOPO  staff  set 
up  an  office  in  a  room  in  the  large  16- Foot  Transonic  Tunnel  building  in  the 
Langley  West  Area. 

When  writing  the  Request  for  Proposals,  Nelson,  Taback,  and  the  others 
involved  could  only  afford  the  time  necessary  to  prepare  a  brief  document, 
merely  a  few  pages  long,  that  sketched  out  some  of  the  detailed  requirements. 
As  Israel  Taback  remembers,  even  before  the  project  office  was  established, 
he  and  a  few  fellow  members  of  what  would  become  LOPO  had  already 
talked  extensively  with  the  potential  contractors.  Taback  explains,  "Our 
idea  was  that  they  would  be  coming  back  to  us  [with  details].  So  it  wasn't 
like  we  were  going  out  cold,  with  a  brand  new  program."25 

Langley  did  need  to  provide  one  critical  detail  in  the  request:  the 
means  for  stabilizing  the  spacecraft  in  lunar  orbit.  Taback  recalls  that  an 
"enormous  difference"  arose  between  Langley  and  NASA  headquarters  over 
this  issue.  The  argument  was  about  whether  the  Request  for  Proposals 
should  require  that  the  contractors  produce  a  rotating  satellite  known  as 
a  "spinner."  The  staff  of  the  OSSA  preferred  a  spinner  based  on  STL's 
previous  study  of  Lunar  Orbiter  requirements.  However,  Langley's  Lunar 
Orbiter  staff  doubted  the  wisdom  of  specifying  the  means  of  stabilization 
in  the  Request  for  Proposals.  They  wished  to  keep  the  door  open  to  other, 
perhaps  better,  ways  of  stabilizing  the  vehicle  for  photography. 

The  goal  of  the  project,  after  all,  was  to  take  the  best  possible  high- 
resolution  pictures  of  the  moon's  surface.  To  do  that,  NASA  needed  to 
create  the  best  possible  orbital  platform  for  the  spacecraft's  sophisticated 
camera  equipment,  whatever  that  turned  out  to  be.  From  their  preliminary 
analysis  and  conversations  about  mission  requirements,  Taback,  Nelson,  and 
others  in  LOPO  felt  that  taking  these  pictures  from  a  three-axis  (yaw,  pitch, 
and  roll),  attitude-stabilized  device  would  be  easier  than  taking  them  from 
a  spinner.  A  spinner  would  cause  distortions  of  the  image  because  of  the 
rotation  of  the  vehicle.  Langley's  John  F.  Newcomb  of  the  Aero-Space 
Mechanics  Division  (and  eventual  member  of  LOPO)  had  calculated  that 
this  distortion  would  destroy  the  resolution  and  thus  seriously  compromise 
the  overall  quality  of  the  pictures.  This  was  a  compromise  that  the  people 
at  Langley  quickly  decided  they  could  not  live  with.  Thus,  for  sound 
technical  reasons,  Langley  insisted  that  the  design  of  the  orbiter  be  kept 
an  open  matter  and  not  be  specified  in  the  Request  for  Proposals.  Even 
if  Langley's  engineers  were  wrong  and  a  properly  designed  spinner  would 

324 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

be  most  effective,  the  sensible  approach  was  to  entertain  all  the  ideas  the 
aerospace  industry  could  come  up  with  before  choosing  a  design.26 

For  several  weeks  in  the  summer  of  1963,  headquarters  tried  to  resist  the 
Langley  position.  Preliminary  studies  by  both  STL  for  the  OSSA  and  by 
Bell  Communications  (BellComm)  for  the  Office  of  Manned  Space  Flight 
indicated  that  a  rotating  spacecraft  using  a  spin-scan  film  camera  similar  to 
the  one  developed  by  the  Rand  Corporation  in  1958  for  an  air  force  satellite 
reconnaissance  system  ("spy  in  the  sky")  would  work  well  for  Lunar  Orbiter. 
Such  a  spinner  would  be  less  complicated  and  less  costly  than  the  three-axis- 
stabilized  spacecraft  preferred  by  Langley.27 

But  Langley  staff  would  not  cave  in  on  an  issue  so  fundamental  to  the 
project's  success.  Eventually  Newell,  Cortright,  Nicks,  and  Scherer  in  the 
OSSA  offered  a  compromise  that  Langley  could  accept:  the  Request  for 
Proposals  could  state  that  "if  bidders  could  offer  approaches  which  differed 
from  the  established  specifications  but  which  would  result  in  substantial 
gains  in  the  probability  of  mission  success,  reliability,  schedule,  and  econ- 
omy," then  NASA  most  certainly  invited  them  to  submit  those  alternatives. 
The  request  would  also  emphasize  that  NASA  wanted  a  lunar  orbiter  that 
was  built  from  as  much  off-the-shelf  hardware  as  possible.  The  development 
of  many  new  technological  systems  would  require  time  that  Langley  did  not 
have.28 

Langley  and  headquarters  had  other  differences  of  opinion  about  the 
request.  For  example,  a  serious  problem  arose  over  the  nature  of  the  con- 
tract. Langley's  chief  procurement  officer,  Sherwood  Butler,  took  the  con- 
servative position  that  a  traditional  cost-plus-a-fixed-fee  contract  would  be 
best  in  a  project  in  which  several  unknown  development  problems  were 
bound  to  arise.  With  this  kind  of  contract,  NASA  would  pay  the  contractor 
for  all  actual  costs  plus  a  sum  of  money  fixed  by  the  contract  negotiations 
as  a  reasonable  profit. 

NASA  headquarters,  on  the  other  hand,  felt  that  some  attractive  financial 
incentives  should  be  built  into  the  contract.  Although  unusual  up  to  this 
point  in  NASA  history,  headquarters  believed  that  an  incentives  contract 
would  be  best  for  Lunar  Orbiter.  Such  a  contract  would  assure  that 
the  contractor  would  do  everything  possible  to  solve  all  the  problems 
encountered  and  make  sure  that  the  project  worked.  The  incentives  could  be 
written  up  in  such  a  way  that  if,  for  instance,  the  contractor  lost  money  on 
any  one  Lunar  Orbiter  mission,  the  loss  could  be  recouped  with  a  handsome 
profit  on  the  other  missions.  The  efficacy  of  a  cost-plus-incentives  contract 
rested  in  the  solid  premise  that  nothing  motivated  a  contractor  more  than 
making  money.  NASA  headquarters  apparently  understood  this  better  than 
Langley's  procurement  officer  who  wanted  to  keep  tight  fiscal  control  over 
the  project  and  did  not  want  to  do  the  hairsplitting  that  often  came  with 
evaluating  whether  the  incentive  clauses  had  been  met.29 

On  the  matter  of  incentives,  Langley's  LOPO  engineers  sided  against 
their  own  man  and  with  NASA  headquarters.  They,  too,  thought  that 

325 


Space/light  Revolution 

incentives  were  the  best  way  to  do  business  with  a  contractor — as  well 
as  the  best  way  to  illustrate  the  urgency  that  NASA  attached  to  Lunar 
Orbiter.30  The  only  thing  that  bothered  them  was  the  vagueness  of  the 
incentives  being  discussed.  When  Director  Floyd  Thompson  understood 
that  his  engineers  really  wanted  to  take  the  side  of  headquarters  on  this 
issue,  he  quickly  concurred.  He  insisted  only  on  three  things:  the  incentives 
had  to  be  based  on  clear  stipulations  tied  to  cost,  delivery,  and  performance, 
with  penalties  for  deadline  overruns;  the  contract  had  to  be  fully  negotiated 
and  signed  before  Langley  started  working  with  any  contractor  (in  other 
words,  work  could  not  start  under  a  letter  of  intent);  and  all  bidding  had 
to  be  competitive.  Thompson  worried  that  the  OSSA  might  be  biased  in 
favor  of  STL  as  the  prime  contractor  because  of  STL's  prior  study  of  the 

Q  1 

requirements  of  lunar  orbiter  systems.  L 

In  mid- August  1963,  with  these  problems  worked  out  with  headquarters, 
Langley  finalized  the  Request  for  Proposals  and  associated  Statement  of 
Work,  which  outlined  specifications,  and  delivered  both  to  Captain  Lee 
R.  Scherer,  Lunar  Orbiter's  program  manager  at  NASA  headquarters,  for 
presentation  to  Ed  Cortright  and  his  deputy  Oran  Nicks.  The  documents 
stated  explicitly  that  the  main  mission  of  Lunar  Orbiter  was  "the  acquisition 
of  photographic  data  of  high  and  medium  resolution  for  selection  of  suitable 
Apollo  and  Surveyor  landing  sites."  The  request  set  out  detailed  criteria 
for  such  things  as  identifying  "cones"  (planar  features  at  right  angles  to 
a  flat  surface),  "slopes"  (circular  areas  inclined  with  respect  to  the  plane 
perpendicular  to  local  gravity),  and  other  subtle  aspects  of  the  lunar  surface. 
Obtaining  information  about  the  size  and  shape  of  the  moon  and  about 
the  lunar  gravitational  field  was  deemed  less  important.  By  omitting  a 
detailed  description  of  the  secondary  objectives  in  the  request,  Langley  made 
clear  that  "under  no  circumstances"  could  anything  "be  allowed  to  dilute 
the  major  photo-reconnaissance  mission."32  The  urgency  of  the  national 
commitment  to  a  manned  lunar  landing  mission  was  the  force  driving  Lunar 
Orbiter.  Langley  wanted  no  confusion  on  that  point. 


The  Source  Evaluation  Board 

Cliff  Nelson  and  LOPO  moved  quickly  in  September  1963  to  create  a 
Source  Evaluation  Board  that  would  possess  the  technical  expertise  and 
good  judgment  to  help  NASA  choose  wisely  from  among  the  industrial 
firms  bidding  for  Lunar  Orbiter.  A  large  board  of  reviewers  (comprising 
more  than  80  evaluators  and  consultants  from  NASA  centers  and  other 
aerospace  organizations)  was  divided  into  groups  to  evaluate  the  technical 
feasibility,  cost,  contract  management  concepts,  business  operations,  and 
other  critical  aspects  of  the  proposals.  One  group,  the  so-called  Scientists' 
Panel,  judged  the  suitability  of  the  proposed  spacecraft  for  providing 
valuable  information  to  the  scientific  community  after  the  photographic 

326 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

mission  had  been  completed.  Langley's  two  representatives  on  the  Scientists' 
Panel  were  Clint  Brown  and  Dr.  Samuel  Katzoff,  an  extremely  insightful 
engineering  analyst,  27-year  Langley  veteran,  and  assistant  chief  of  the 
Applied  Materials  and  Physics  Division. 

Although  the  opinions  of  all  the  knowledgeable  outsiders  were  taken 
seriously,  Langley  intended  to  make  the  decision.33  Chairing  the  Source 
Evaluation  Board  was  Eugene  Draley,  one  of  Floyd  Thompson's  assistant 
directors.  When  the  board  finished  interviewing  all  the  bidders,  hearing 
their  oral  presentations,  and  tallying  the  results  of  its  scoring  of  the 
proposals  (a  possible  70  points  for  technical  merit  and  30  points  for  business 
management),  it  was  to  present  a  formal  recommendation  to  Thompson.  He 
in  turn  would  pass  on  the  findings  with  comments  to  Homer  Newell's  office 
in  Washington. 

Five  major  aerospace  firms  submitted  proposals  for  the  Lunar  Orbiter 
contract.  Three  were  California  firms:  STL  in  Redondo  Beach,  Lockheed 
Missiles  and  Space  Company  of  Sunnyvale,  and  Hughes  Aircraft  Company  of 
Los  Angeles.  The  Martin  Company  of  Baltimore  and  the  Boeing  Company 
of  Seattle  were  the  other  two  bidders.34 

Three  of  the  five  proposals  were  excellent.  Hughes  had  been  developing 
an  ingenious  spin-stabilization  system  for  geosynchronous  communication 
satellites,  which  helped  the  company  to  submit  an  impressive  proposal  for  a 
rotating  vehicle.  With  Hughes's  record  in  spacecraft  design  and  fabrication, 
the  Source  Evaluation  Board  gave  Hughes  serious  consideration.  STL  also 
submitted  a  fine  proposal  for  a  spin-stabilized  rotator.  This  came  as  no 
surprise,  of  course,  given  STL's  prior  work  for  Surveyor  as  well  as  its  prior 
contractor  studies  on  lunar  orbiter  systems  for  NASA  headquarters. 

The  third  outstanding  proposal — entitled  "ACLOPS"  (Agena-Class 
Lunar  Orbiter  Project) — was  Boeing's.  The  well-known  airplane  manu- 
facturer had  not  been  among  the  companies  originally  invited  to  bid  on 
Lunar  Orbiter  and  was  not  recognized  as  the  most  logical  of  contenders. 
However,  Boeing  recently  had  successfully  completed  the  Bomarc  missile 
program  and  was  anxious  to  become  involved  with  the  civilian  space  pro- 
gram, especially  now  that  the  DOD  was  canceling  Dyna-Soar,  an  air  force 
project  for  the  development  of  an  experimental  X-20  aerospace  plane.  This 
cancellation  released  several  highly  qualified  U.S.  Air  Force  personnel,  who 
were  still  working  at  Boeing,  to  support  a  new  Boeing  undertaking  in  space. 
Company  representatives  had  visited  Langley  to  discuss  Lunar  Orbiter,  and 
Langley  engineers  had  been  so  excited  by  what  they  had  heard  that  they 
had  pestered  Thompson  to  persuade  Seamans  to  extend  an  invitation  to 
Boeing  to  join  the  bidding.  The  proposals  from  Martin,  a  newcomer  in  the 
business  of  automated  space  probes,  and  Lockheed,  a  company  with  years 
of  experience  handling  the  Agena  space  vehicle  for  the  air  force,  were  also 
quite  satisfactory.  In  the  opinion  of  the  Source  Evaluation  Board,  however, 
the  proposals  from  Martin  and  Lockheed  were  not  as  strong  as  those  from 
Boeing  and  Hughes. 

327 


Space/light  Revolution 

The  LOPO  staff  and  the  Langley  representatives  decided  early  in  the 
evaluation  that  they  wanted  Boeing  to  be  selected  as  the  contractor;  on 
behalf  of  the  technical  review  team,  Israel  Taback  had  made  this  preference 
known  both  in  private  conversations  with,  and  formal  presentations  to, 
the  Source  Evaluation  Board.  Boeing  was  Langley's  choice  because  it  pro- 
posed a  three-axis-stabilized  spacecraft  rather  than  a  spinner.  For  attitude 
reference  in  orbit,  the  spacecraft  would  use  an  optical  sensor  similar  to  the 
one  that  was  being  planned  for  use  on  the  Mariner  C  spacecraft,  which  fixed 
on  the  star  Canopus. 

An  attitude-stabilized  orbiter  eliminated  the  need  for  a  focal-length  spin 
camera.  This  type  of  photographic  system,  first  conceived  by  Merton  E. 
Davies  of  the  Rand  Corporation  in  1958,  could  compensate  for  the  dis- 
tortions caused  by  a  rotating  spacecraft  but  would  require  extensive 
development.  In  the  Boeing  proposal,  Lunar  Orbiter  would  carry  a 
photo  subsystem  designed  by  Eastman  Kodak  and  used  on  DOD  spy 
satellites.35  This  subsystem  worked  automatically  and  with  the  precision  of 
a  Swiss  watch.  It  employed  two  lenses  that  took  pictures  simultaneously  on 
a  roll  of  70-millimeter  aerial  film.  If  one  lens  failed,  the  other  still  worked. 
One  lens  had  a  focal  length  of  610  millimeters  (24  inches)  and  could  take 
pictures  from  an  altitude  of  46  kilometers  (28.5  miles)  with  a  high  resolution 
for  limited-area  coverage  of  approximately  1  meter.  The  other,  which  had 
a  focal  length  of  about  80  millimeters  (3  inches),  could  take  pictures  with 
a  medium  resolution  of  approximately  8  meters  for  wide  coverage  of  the 
lunar  surface.  The  film  would  be  developed  on  board  the  spacecraft  using 
the  proven  Eastman  Kodak  "Bimat"  method.  The  film  would  be  in  contact 
with  a  web  containing  a  single-solution  dry  processing  chemical,  which  elim- 
inated the  need  to  use  wet  chemicals.  Developed  automatically  and  wound 
onto  a  storage  spool,  the  processed  film  could  then  be  "read  out"  and  trans- 
mitted by  the  spacecraft's  communications  subsystem  to  receiving  stations 
of  JPL's  worldwide  Deep  Space  Network,  which  was  developed  for  commu- 
nication with  spacefaring  vehicles  destined  for  the  moon  and  beyond.36 

How  Boeing  had  the  good  sense  to  propose  an  attitude-stabilized  plat- 
form based  on  the  Eastman  Kodak  camera,  rather  than  to  propose  a  rotator 
with  a  yet-to-be  developed  camera  is  not  totally  clear.  Langley  engineers  had 
conversed  with  representatives  of  all  the  interested  bidders,  so  Boeing's  peo- 
ple might  possibly  have  picked  up  on  Langley's  concerns  about  the  quality  of 
photographs  from  spinners.  The  other  bidders,  especially  STL  and  Hughes, 
with  their  expertise  in  spin-stabilized  spacecraft,  might  also  have  picked  up 
on  those  concerns  but  were  too  confident  in  the  type  of  rotationally  stabilized 
system  they  had  been  working  on  to  change  course  in  midstream. 

Furthermore,  Boeing  had  been  working  closely  with  RCA,  which  for 
a  time  was  also  thinking  about  submitting  a  proposal  for  Lunar  Orbiter. 
RCA's  idea  was  a  lightweight  (200-kilogram),  three-axis,  attitude-stabilized, 
and  camera-bearing  pay  load  that  could  be  injected  into  lunar  orbit  as  part  of 
a  Ranger-type  probe.  A  lunar  orbiter  study  group,  chaired  by  Lee  Scherer 

328 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-67-5655 

Lunar  Orbiter  was  essentially  a  flying  camera.  The  payload  structure  was  built 
around  a  pressurized  shell  holding  Eastman  Kodak's  dual-imaging  photographic 
system,  which  used  a  camera  with  wide-angle  and  telephoto  lenses  that  could 
simultaneously  take  two  kinds  of  pictures  on  the  same  film. 


at  NASA  headquarters,  had  evaluated  RCA's  approach  in  October  1962, 
however,  and  found  it  lacking.  It  was  too  expensive  ($20.4  million  for  flying 
only  three  spacecraft),  and  its  proposed  vidicon  television  unit  could  not 
cover  the  lunar  surface  either  in  the  detail  or  the  wide  panoramas  NASA 
wanted.37 

Boeing  knew  all  about  this  rejected  RCA  approach.  After  talking  to 
Langley's  engineers,  the  company  shrewdly  decided  to  stay  with  an  attitude- 
stabilized  orbiter  but  to  dump  the  use  of  the  inadequate  vidicon  television. 
Boeing  replaced  the  television  system  with  an  instrument  with  a  proven 
track  record  in  planetary  reconnaissance  photography:  the  Eastman  Kodak 
spy  camera.38 

On  20  December  1963,  two  weeks  after  the  Source  Evaluation  Board  made 
its  formal  recommendation  to  Administrator  James  Webb  in  Washington, 
NASA  announced  that  it  would  be  negotiating  with  Boeing  as  prime  con- 
tractor for  the  Lunar  Orbiter  project.  Along  with  the  excellence  of  its  pro- 
posed spacecraft  design  and  Kodak  camera,  NASA  singled  out  the  strength 
of  Boeing's  commitment  to  the  project  and  its  corporate  capabilities  to 

329 


Space/light  Revolution 


Capt.  Lee  R.  Scherer  served  as  Lunar 
Orbiter's  program  manager  at  NASA 
headquarters. 


L-66-6301 


complete  it  on  schedule  without  relying  on  many  subcontractors.  Still,  the 
choice  was  a  bit  ironic.  Only  14  months  earlier,  the  Scherer  study  group 
had  rejected  RCA's  approach  in  favor  of  a  study  of  a  spin-stabilized  space- 
craft proposed  by  STL.  Now  Boeing  had  outmaneuvered  its  competition  by 
proposing  a  spacecraft  that  incorporated  essential  features  of  the  rejected 
RCA  concept  and  almost  none  from  the  STL's  previously  accepted  one. 

Boeing  won  the  contract  even  though  it  asked  for  considerably  more 
money  than  any  of  the  other  bidders.  The  lowest  bid,  from  Hughes,  was 
$41,495,339,  less  than  half  of  Boeing's  $83,562,199,  a  figure  that  would 
quickly  rise  when  the  work  started.  Not  surprisingly,  NASA  faced  some 
congressional  criticism  and  had  to  defend  its  choice.  The  agency  justified 
its  selection  by  referring  confidently  to  what  Boeing  alone  proposed  to  do 
to  ensure  protection  of  Lunar  Orbiter's  photographic  film  from  the  hazards 
of  solar  radiation.39 

This  was  a  technical  detail  that  deeply  concerned  LOPO.  Experiments 
conducted  by  Boeing  and  by  Dr.  Trutz  Foelsche,  a  Langley  scientist  in  the 
Space  Mechanics  (formerly  Theoretical  Mechanics)  Division  who  specialized 
in  the  study  of  space  radiation  effects,  suggested  that  even  small  doses 
of  radiation  from  solar  flares  could  fog  ordinary  high-speed  photographic 
film.  This  would  be  true  especially  in  the  case  of  an  instrumented  probe 
like  Lunar  Orbiter,  which  had  thin  exterior  vehicular  shielding.  Even  if 
the  thickness  of  the  shielding  around  the  film  was  increased  tenfold  (from 
1  g/cm2  to  10  g/cm2),  Foelsche  judged  that  high-speed  film  would  not  make 
it  through  a  significant  solar-particle  event  without  serious  damage.40  Thus, 

330 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-64-4141 

Representatives  of  NASA  Langley  and  Boeing  signed  the  Lunar  Orbiter  contract 
on  16  April  1964  ana  sent  it  to  NASA  headquarters  for  final  review.  Three  weeks 
later,  on  7  May,  Administrator  James  E.  Webb  approved  the  $80-million  incentives 
contract  to  build  five  Lunar  Orbiter  spacecraft. 


something  extraordinary  had  to  be  done  to  protect  the  high-speed  film.  A 
better  solution  was  not  to  use  high-speed  film  at  all. 

As  NASA  explained  successfully  to  its  critics,  the  other  bidders  for  the 
Lunar  Orbiter  contract  relied  on  high-speed  film  and  faster  shutter  speeds 
for  their  on-board  photographic  subsystems.  Only  Boeing  did  not.  When 
delegates  from  STL,  Hughes,  Martin,  and  Lockheed  were  asked  at  a  bidders' 
briefing  in  November  1963  about  what  would  happen  to  their  film  if  a  solar 
event  occurred  during  an  orbiter  mission,  they  all  had  to  admit  that  the 
film  would  be  damaged  seriously.  Only  Boeing  could  claim  otherwise.  Even 
with  minimal  shielding,  the  more  insensitive,  low-speed  film  used  by  the 
Kodak  camera  would  not  be  fogged  by  high-energy  radiation,  not  even  if 
the  spacecraft  moved  through  the  Van  Allen  radiation  belts.41  This,  indeed, 
proved  to  be  the  case.  During  the  third  mission  of  Lunar  Orbiter  in  February 
1967,  a  solar  flare  with  a  high  amount  of  optical  activity  did  occur,  but  the 
film  passed  through  it  unspoiled.42 

Negotiations  with  Boeing  did  not  take  long.  Formal  negotiations  began 
on  17  March  1964,  and  ended  just  four  days  later.  On  7  May  Administrator 
Webb  signed  the  document  that  made  Lunar  Orbiter  an  official  NASA 

331 


Space/light  Revolution 

commitment.  Hopes  were  high.  But  in  the  cynical  months  of  1964,  with 
Ranger's  setbacks  still  making  headlines  and  critics  still  faulting  NASA  for 
failing  to  match  Soviet  achievements  in  space,  everyone  doubted  whether 
Lunar  Orbiter  would  be  ready  for  its  first  scheduled  flight  to  the  moon  in 
just  two  years. 

Nelson's  Team 

Large  projects  are  run  by  only  a  handful  of  people.  Four  or  five  key 
individuals  delegate  jobs  and  responsibilities  to  others.  This  was  certainly 
true  for  Lunar  Orbiter.  From  start  to  finish,  Langley's  LOPO  remained 
a  small  organization;  its  original  nucleus  of  9  staff  members  never  grew 
any  larger  than  50  professionals.  Langley  management  knew  that  keeping 
LOPO's  staff  small  meant  fewer  people  in  need  of  positions  when  the  project 
ended.  If  all  the  positions  were  built  into  a  large  project  office,  many  careers 
would  be  out  on  a  limb;  a  much  safer  organizational  method  was  for  a  small 
project  office  to  draw  people  from  other  research  and  technical  divisions  to 
assist  the  project  as  needed.43 

In  the  case  of  Lunar  Orbiter,  four  men  ran  the  project:  Cliff  Nelson, 
the  project  manager;  Israel  Taback,  who  was  in  charge  of  all  activities 
leading  to  the  production  and  testing  of  the  spacecraft;  Bill  Boyer,  who 
was  responsible  for  planning  and  integrating  launch  and  flight  operations; 
and  James  V.  Martin,  the  assistant  project  manager.  Nelson  had  accepted 
the  assignment  with  Thompson's  assurance  that  he  would  be  given  wide 
latitude  in  choosing  the  men  and  women  he  wanted  to  work  with  him  in  the 
project  office.  As  a  result,  virtually  all  of  his  top  people  were  handpicked. 

The  one  significant  exception  was  his  chief  assistant,  Jim  Martin.  In 
September  1964,  the  Langley  assistant  director  responsible  for  the  project 
office,  Gene  Draley,  brought  in  Martin  to  help  Nelson  cope  with  some  of 
the  stickier  details  of  Lunar  Orbiter 's  management.  A  senior  manager 
in  charge  of  Republic  Aviation's  space  systems  requirements,  Martin  had 
a  tremendous  ability  for  anticipating  business  management  problems  and 
plenty  of  experience  taking  care  of  them.  Furthermore,  he  was  a  well- 
organized  and  skillful  executive  who  could  make  schedules,  set  due  dates, 
and  closely  track  the  progress  of  the  contractors  and  subcontractors.  This 
"paper"  management  of  a  major  project  was  troublesome  for  Cliff  Nelson, 
a  quiet  people-oriented  person.  Draley  knew  about  taskmaster  Martin  from 
Republic's  involvement  in  Project  Fire  and  was  hopeful  that  Martin's  acer- 
bity and  business-mindedness  would  complement  Nelson's  good-heartedness 
and  greater  technical  depth,  especially  in  dealings  with  contractors. 

Because  Cliff  Nelson  and  Jim  Martin  were  so  entirely  opposite  in  per- 
sonality, they  did  occasionally  clash,  which  caused  a  few  internal  problems 
in  LOPO.  On  the  whole,  however,  the  alliance  worked  quite  well,  although 
it  was  forced  by  Langley  management.  Nelson  generally  oversaw  the  whole 
endeavor  and  made  sure  that  everybody  worked  together  as  a  team.  For 

332 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-66-6301 

In  contrast  to  the  quiet,  people- oriented  LOPO  director  Cliff  Nelson  (left),  James 
"Big  Jim"  Martin  (right)  breathed  fire  when  it  came  to  getting  consistent  top-grade 
performance  from  the  NASA  contractors. 


the  monitoring  of  the  day-to-day  progress  of  the  project's  many  operations, 
Nelson  relied  on  the  dynamic  Martin.  For  example,  when  problems  arose 
with  the  motion-compensation  apparatus  for  the  Kodak  camera,  Martin 
went  to  the  contractor's  plant  to  assess  the  situation  and  decided  that 
its  management  was  not  placing  enough  emphasis  on  following  a  schedule. 
Martin  acted  tough,  pounded  on  the  table,  and  made  the  contractor  put 
workable  schedules  together  quickly.  When  gentler  persuasion  was  called  for 
or  subtler  interpersonal  relationships  were  involved,  Nelson  was  the  person 
for  the  job.  Martin,  who  was  technically  competent  but  not  as  technically 
talented  as  Nelson,  also  deferred  to  the  project  manager  when  a  decision  re- 
quired particularly  complex  engineering  analysis.  Thus,  the  two  men  worked 
together  for  the  overall  betterment  of  Lunar  Orbiter.44 

Placing  an  excellent  person  with  just  the  right  specialization  in  just 
the  right  job  was  one  of  the  most  important  elements  behind  the  success 
of  Lunar  Orbiter,  and  for  this  eminently  sensible  approach  to  project 
management,  Cliff  Nelson  and  Floyd  Thompson  deserve  the  lion's  share  of 
credit.  Both  men  cultivated  a  management  style  that  emphasized  direct 
dealings  with  people  and  often  ignored  formal  organizational  channels. 
Both  stressed  the  importance  of  teamwork  and  would  not  tolerate  any 
individual,  however  talented,  willfully  undermining  the  esprit  de  corps. 
Before  filling  any  position  in  the  project  office,  Nelson  gave  the  selection 
much  thought.  He  questioned  whether  the  people  under  consideration  were 
compatible  with  others  already  in  his  project  organization.  He  wanted  to 
know  whether  candidates  were  goal-oriented — willing  to  do  whatever  was 

333 


Space/light  Revolution 

necessary  (working  overtime  or  traveling)  to  complete  the  project.45  Because 
Langley  possessed  so  many  employees  who  had  been  working  at  the  center 
for  many  years,  the  track  record  of  most  people  was  either  well  known  or  easy 
to  ascertain.  Given  the  outstanding  performance  of  Lunar  Orbiter  and  the 
testimonies  about  an  exceptionally  healthy  work  environment  in  the  project 
office,  Nelson  did  an  excellent  job  predicting  who  would  make  a  productive 
member  of  the  project  team.46 

Considering  Langley's  historic  emphasis  on  fundamental  applied  aero- 
nautical research,  it  might  seem  surprising  that  Langley  scientists  and  en- 
gineers did  not  try  to  hide  inside  the  dark  return  passage  of  a  wind  tunnel 
rather  than  be  diverted  into  a  spaceflight  project  like  Lunar  Orbiter.  As  has 
been  discussed,  some  researchers  at  Langley  (and  agencywide)  objected  to 
and  resisted  involvement  with  project  work.  The  Surveyor  project  at  JPL 
had  suffered  from  staff  members'  reluctance  to  leave  their  own  specialties 
to  work  on  a  space  project.  However,  by  the  early  1960s  the  enthusiasm 
for  spaceflight  ran  so  rampant  that  it  was  not  hard  to  staff  a  space  project 
office.  All  the  individuals  who  joined  LOPO  at  Langley  came  enthusiasti- 
cally; otherwise  Cliff  Nelson  would  not  have  had  them.  Israel  Taback,  who 
had  been  running  the  Communications  and  Control  Branch  of  IRD,  remem- 
bers having  become  distressed  with  the  thickening  of  what  he  calls  "the 
paper  forest" :  the  preparation  of  five-year  plans,  ten-year  plans,  and  other 
lengthy  documents  needed  to  justify  NASA's  budget  requests.  The  work  he 
had  been  doing  with  airplanes  and  aerospace  vehicles  was  interesting  (he 
had  just  finished  providing  much  of  the  flight  instrumentation  for  the  X-15 
program),  but  not  so  interesting  that  he  wanted  to  turn  down  Cliff  Nelson's 
offer  to  join  Lunar  Orbiter.  "The  project  was  brand  new  and  sounded  much 
more  exciting  than  what  I  had  been  doing,"  Taback  remembers.  It  appealed 
to  him  also  because  of  its  high  visibility  both  inside  and  outside  the  center. 
Everyone  had  to  recognize  the  importance  of  a  project  directly  related  to 
the  national  goal  of  landing  a  man  on  the  moon.47 

Norman  L.  Crabill,  the  head  of  LOPO's  mission  design  team,  also  decided 
to  join  the  project.  On  a  Friday  afternoon,  he  had  received  the  word  that 
one  person  from  his  branch  of  the  Applied  Materials  and  Physics  Division 
would  have  to  be  named  by  the  following  Monday  as  a  transfer  to  LOPO;  as 
branch  head,  Crabill  himself  would  have  to  make  the  choice.  That  weekend 
he  asked  himself,  "What's  your  own  future,  Crabill?  This  is  space.  If  you 
don't  step  up  to  this,  what's  your  next  chance.  You've  already  decided  not 
to  go  with  the  guys  to  Houston."  He  immediately  knew  who  to  transfer, 
"It  was  me."  That  was  how  he  "got  into  the  space  business."  And  in  his 
opinion,  it  was  "the  best  thing"  that  he  ever  did.48 

The  Boeing  Team 

Cliff  Nelson's  office  had  the  good  sense  to  realize  that  monitoring  the 
prime  contractor  did  not  entail  doing  Boeing's  work  for  Boeing.  Nelson 

334 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

approached  the  management  of  Lunar  Orbiter  more  practically:  the  con- 
tractor was  "to  perform  the  work  at  hand  while  the  field  center  retained 
responsibility  for  overseeing  his  progress  and  assuring  that  the  job  was  done 
according  to  the  terms  of  the  contract."  For  Lunar  Orbiter,  this  philosophy 
meant  specifically  that  the  project  office  would  have  to  keep  "a  continu- 
ing watch  on  the  progress  of  the  various  components,  subsystems,  and  the 
whole  spacecraft  system  during  the  different  phases  of  designing,  fabricating 
and  testing  them."4  Frequent  meetings  would  take  place  between  Nelson 
and  his  staff  and  their  counterparts  at  Boeing  to  discuss  all  critical  mat- 
ters, but  Langley  would  not  assign  all  the  jobs,  solve  all  the  problems,  or 
micromanage  every  detail  of  the  contractor's  work. 

This  philosophy  sat  well  with  Robert  J.  Helberg,  head  of  Boeing's 
Lunar  Orbiter  team.  Helberg  had  recently  finished  directing  the  company's 
work  on  the  Bomarc  missile,  making  him  a  natural  choice  for  manager  of 
Boeing's  next  space  venture.  The  Swedish-born  Helberg  was  absolutely 
straightforward,  and  all  his  people  respected  him  immensely — as  would 
everyone  in  LOPO.  He  and  fellow  Swede  Cliff  Nelson  got  along  famously. 
Their  relaxed  relationship  set  the  tone  for  interaction  between  Langley  and 
Boeing.  Ideas  and  concerns  passed  freely  back  and  forth  between  the  project 
offices.  Nelson  and  his  people  "never  had  to  fear  the  contractor  was  just 
telling  [them]  a  lie  to  make  money,"  and  Helberg  and  his  tightly  knit,  220- 
member  Lunar  Orbiter  team  never  had  to  complain  about  uncaring,  paper- 
shuffling  bureaucrats  who  were  mainly  interested  in  dotting  all  the  i's  and 
crossing  all  the  t's  and  making  sure  that  nothing  illegal  was  done  that  could 
bother  government  auditors  and  put  their  necks  in  a  wringer.50 

The  Langley/NASA  headquarters  relationship  was  also  harmonious  and 
effective.  This  was  in  sharp  contrast  to  the  relationship  between  JPL 
and  headquarters  during  the  Surveyor  project.  Initially,  JPL  had  tried 
to  monitor  the  Surveyor  contractor,  Hughes,  with  only  a  small  staff  that 
provided  little  on-site  technical  direction;  however,  because  of  unclear 
objectives,  the  open-ended  nature  of  the  project  (such  basic  things  as 
which  experiment  packages  would  be  included  on  the  Surveyor  spacecraft 
were  uncertain),  and  a  too  highly  diffused  project  organization  within 
Hughes,  JPL's  "laissez-faire"  approach  to  project  management  did  not 
work.  As  the  problems  snowballed,  Cortright  found  it  necessary  to  intervene 
and  compelled  JPL  to  assign  a  regiment  of  on-site  supervisors  to  watch 
over  every  detail  of  the  work  being  done  by  Hughes.  Thus,  as  one 
analyst  of  Surveyor's  management  has  observed,  "the  responsibility  for 
overall  spacecraft  development  was  gradually  retrieved  from  Hughes  by  JPL, 
thereby  altering  significantly  the  respective  roles  of  the  field  center  and  the 
spacecraft  systems  contractors."51 

Nothing  so  unfortunate  happened  during  Lunar  Orbiter,  partly  because 
NASA  had  learned  from  the  false  steps  and  outright  mistakes  made  in 
the  management  of  Surveyor.  For  example,  NASA  now  knew  that  before 
implementing  a  project,  everyone  involved  must  take  part  in  extensive 

335 


Spaceflight  Revolution 

preliminary  discussions.  These  conversations  ensured  that  the  project's 
goals  were  certain  and  each  party's  responsibilities  clear.  Each  office  should 
expect  maximum  cooperation  and  minimal  unnecessary  interference  from 
the  others.  Before  Lunar  Orbiter  was  under  way,  this  excellent  groundwork 
had  been  laid. 

As  has  been  suggested  by  a  1972  study  done  by  the  National  Academy 
of  Public  Administration,  the  Lunar  Orbiter  project  can  serve  as  a  model 
of  the  ideal  relationship  between  a  prime  contractor,  a  project  office,  a  field 
center,  a  program  office,  and  headquarters.  From  start  to  finish  nearly 
everything  important  about  the  interrelationship  worked  out  superbly  in 
Lunar  Orbiter.  According  to  LOPO's  Israel  Taback,  "Everyone  worked 
together  harmoniously  as  a  team  whether  they  were  government,  from 
headquarters  or  from  Langley,  or  from  Boeing."  No  one  tried  to  take 
advantage  of  rank  or  to  exert  any  undue  authority  because  of  an  official 
title  or  organizational  affiliation.52  That  is  not  to  say  that  problems  never 
occurred  in  the  management  of  Lunar  Orbiter.  In  any  large  and  complex 
technological  project  involving  several  parties,  some  conflicts  are  bound  to 
arise.  The  key  to  project  success  lies  in  how  differences  are  resolved. 


The  "Concentrated"  versus  the  "Distributed"  Mission 

The  most  fundamental  issue  in  the  premission  planning  for  Lunar  Or- 
biter was  how  the  moon  was  to  be  photographed.  Would  the  photography 
be  "concentrated"  on  a  predetermined  single  target,  or  would  it  be  "dis- 
tributed" over  several  selected  targets  across  the  moon's  surface?  On  the 
answer  to  this  basic  question  depended  the  successful  integration  of  the 
entire  mission  plan  for  Lunar  Orbiter. 

For  Lunar  Orbiter,  as  with  any  other  spaceflight  program,  mission  plan- 
ning involved  the  establishment  of  a  complicated  sequence  of  events:  When 
should  the  spacecraft  be  launched?  When  does  the  launch  window  open  and 
close?  On  what  trajectory  should  the  spacecraft  arrive  in  lunar  orbit?  How 
long  will  it  take  the  spacecraft  to  get  to  the  moon?  How  and  when  should 
orbital  "injection"  take  place?  How  and  when  should  the  spacecraft  get  to 
its  target  (s),  and  at  what  altitude  above  the  lunar  surface  should  it  take  the 
pictures?  Where  does  the  spacecraft  need  to  be  relative  to  the  sun  for  taking 
optimal  pictures  of  the  lunar  surface?  Answering  these  questions  also  meant 
that  NASA's  mission  planners  had  to  define  the  lunar  orbits,  determine  how 
accurately  those  orbits  could  be  navigated,  and  know  the  fuel  requirements. 
The  complete  mission  profile  had  to  be  ready  months  before  launch.  And 
before  the  critical  details  of  the  profile  could  be  made  ready,  NASA  had  to 
select  the  targeted  areas  on  the  lunar  surface  and  decide  how  many  of  them 
were  to  be  photographed  during  the  flight  of  a  single  orbiter.53 

Originally  NASA's  plan  was  to  conduct  a  concentrated  mission.  The 
Lunar  Orbiter  would  go  up  and  target  a  single  site  of  limited  dimensions. 

336 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


Top  NASA  officials  listen  to 
a  LOPO  briefing  at  Langley 
in  December  1966.  Sitting  to 
the  far  right  with  his  hand  on 
his  chin  is  Floyd  Thompson. 
To  the  left  sits  Dr.  George 
Mueller,  NASA  associate  ad- 
ministrator for  Manned  Space 
Flight.  On  the  wall  is  a  di- 
agram of  the  sites  selected  for 
the  "concentrated  mission. " 
The  chart  below  illustrates  the 
primary  area  of  photographic 
interest. 


L-66- 10290 


MOON'S  ROTATION 
ON  ITS  AXIS 


-  TARGET  SITES 


TYPICAL 

SPACECRAFT 

ORBIT 


MOON'S  TRAVEL 
AROUND  EARTH 
AS  OBSERVED 
FROM  EARTH 


PRIMARY  AREA  OF 
PHOTOGRAPHIC  INTEREST 
(REGION  FOR  SELECTION 
OF  APOLLO  LANDING  AREA) 


LUNAR  SURFACE  PHOTOGRAPHIC  COVERAGE 


337 


Spaceflight  Revolution 

The  country's  leading  astrogeologists  would  help  in  the  site  selection  by 
identifying  the  smoothest,  most  attractive  possibilities  for  a  manned  lunar 
landing.  The  U.S.  Geological  Survey  had  drawn  huge,  detailed  maps  of  the 
lunar  surface  from  the  best  available  telescopic  observations.  With  these 
maps,  NASA  would  select  one  site  as  the  prime  target  for  each  of  the  five 
Lunar  Orbiter  missions.  During  a  mission,  the  spacecraft  would  travel  into 
orbit  and  move  over  the  target  at  the  "perilune,"  or  lowest  point  in  the 
orbit  (approximately  50  kilometers  [31.1  miles]  above  the  surface);  then 
it  would  start  taking  pictures.  Successive  orbits  would  be  close  together 
longitudinally,  and  the  Lunar  Orbiter's  camera  would  resume  photographing 
the  surface  each  time  it  passed  over  the  site.  The  high- resolution  lens  would 
take  a  1-meter-resolution  picture  of  a  small  area  (4  x  16  kilometers)  while  at 
exactly  the  same  time,  the  medium-resolution  lens  would  take  an  8-meter- 
resolution  picture  of  a  wider  area  (32  x  37  kilometers).  The  high-resolution 
lens  would  photograph  at  such  a  rapid  interval  that  the  pictures  would  just 
barely  overlap.  The  wide-angle  pictures,  taken  by  the  medium-resolution 
lens,  would  have  a  conveniently  wide  overlap.  All  the  camera  exposures 
would  take  place  in  24  hours,  thus  minimizing  the  threat  to  the  film  from  a 
solar  flare.  The  camera's  capacity  of  roughly  200  photographic  frames  would 
be  devoted  to  one  location.  The  result  would  be  one  area  shot  in  adjacent, 
overlapping  strips.  By  putting  the  strips  together,  NASA  had  a  picture  of  a 
central  1-meter-resolution  area  that  was  surrounded  by  a  broader  8-meter- 
resolution  area — in  other  words,  it  would  be  one  large,  rich  stereoscopic 
picture  of  a  choice  lunar  landing  site.  NASA  would  learn  much  about  that 
one  ideal  place,  and  the  Apollo  program  would  be  well  served.54 

The  plan  sounded  fine  to  everyone — at  least  in  the  beginning.  Langley's 
Request  for  Proposals  had  specified  the  concentrated  mission,  and  Boeing 
had  submitted  the  winning  proposal  based  on  that  mission  plan.  Moreover, 
intensive,  short-term  photography  like  that  called  for  in  a  concentrated 
mission  was  exactly  what  Eastman  Kodak's  high-resolution  camera  system 
had  been  designed  for.  The  camera  was  a  derivative  of  a  spy  satellite  photo 
system  created  specifically  for  earth  reconnaissance  missions  specified  by  the 
DOD.* 


U- 

In  the  top-secret  DOD  system,  the  camera  with  the  film  inside  apparently  would  reenter  the 
atmosphere  inside  a  heat-shielded  package  that  parachuted  down,  was  hooked,  and  was  physically 
retrieved  in  midair  (if  all  went  as  planned)  by  a  specially  equipped  U.S.  Air  Force  C-119  cargo  airplane. 
It  was  obviously  a  very  unsatisfactory  system,  but  in  the  days  before  advanced  electronic  systems,  it 
was  the  best  high-resolution  satellite  reconnaissance  system  that  modern  technology  could  provide.  Few 
NASA  people  were  ever  privy  to  many  of  the  details  of  how  the  "black  box"  actually  worked,  because 
they  did  not  have  "the  need  to  know."  However,  they  figured  that  it  had  been  designed,  as  one  LOPO 
engineer  has  described  in  much  oversimplified  layman's  terms,  "so  when  a  commander  said,  'we've  got 
the  target,'  bop,  take  your  snapshots,  zap,  zap,  zap,  get  it  down  from  orbit,  retrieve  it  and  bring  it 
home,  rush  it  off  to  Kodak,  and  get  your  pictures."  (Norman  Crabill  interview  with  author,  Hampton, 
Va.,  28  August  1991.) 

338 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

As  LOPO's  mission  planners  gave  the  plan  more  thought,  however, 
they  realized  that  the  concentrated  mission  approach  was  flawed.  Norman 
Crabill,  Langley's  head  of  mission  integration  for  Lunar  Orbiter,  remembers 
the  question  he  began  to  ask  himself,  "What  happens  if  only  one  of  these 
missions  is  going  to  work?  This  was  in  the  era  of  Ranger  failures  and 
Surveyor  slippage.  When  you  shoot  something,  you  had  only  a  twenty 
percent  probability  that  it  was  going  to  work.  It  was  that  bad."  On  that 
premise,  NASA  planned  to  fly  five  Lunar  Orbiters,  hoping  that  one  would 
operate  as  it  should.  "Suppose  we  go  up  there  and  shoot  all  we  [have]  on  one 
site,  and  it  turns  out  to  be  no  good?"  fretted  Crabill,  and  others  began  to 
worry  as  well.  What  if  that  site  was  not  as  smooth  as  it  appeared  on  the  U.S. 
Geological  Survey  maps,  or  a  gravitational  anomaly  or  orbital  perturbation 
was  present,  making  that  particular  area  of  the  moon  unsafe  for  a  lunar 
landing?  And  what  if  that  Lunar  Orbiter  turned  out  to  be  the  only  one  to 
work?  What  then?55 

In  late  1964,  over  the  course  of  several  weeks,  LOPO  became  more 
convinced  that  it  should  not  be  putting  all  its  eggs  in  one  basket.  "We 
developed  the  philosophy  that  we  really  didn't  want  to  do  the  concentrated 
mission;  what  we  really  wanted  to  do  was  what  we  called  the  'distributed 
mission,'"  recalls  Crabill.  The  advantage  of  the  distributed  mission  was 
that  it  would  enable  NASA  to  inspect  several  choice  targets  in  the  Apollo 
landing  zone  with  only  one  spacecraft.56 

In  early  1965,  Norm  Crabill  and  Tom  Young  of  the  LOPO  mission 
integration  team  traveled  to  the  office  of  the  U.S.  Geological  Survey  in 
Flagstaff,  Arizona.  There,  the  Langley  engineers  consulted  with  U.S. 
government  astrogeologists  John  F.  McCauley,  Lawrence  Rowan,  and  Harold 
Masursky.  Jack  McCauley  was  Flagstaff's  top  man  at  the  time,  but  he 
assigned  Larry  Rowan,  "a  young  and  upcoming  guy,  very  reasonable  and 
very  knowledgeable,"  the  job  of  heading  the  Flagstaff  review  of  the  Lunar 
Orbiter  site  selection  problem.  "We  sat  down  with  Rowan  at  a  table  with 
these  big  lunar  charts,"  and  Rowan  politely  reminded  the  Langley  duo  that 
"the  dark  areas  on  the  moon  were  the  smoothest."  Rowan  then  pointed  to 
the  darkest  places  across  the  entire  face  of  the  moon.57 

Rowan  identified  10  good  targets.  When  Crabill  and  Young  made  orbital 
calculations,  they  became  excited.  In  a  few  moments,  they  had  realized  that 
they  wanted  to  do  the  distributed  mission.  Rowan  and  his  colleagues  in 
Flagstaff  also  became  excited  about  the  prospects.  This  was  undoubtedly 
the  way  to  catch  as  many  landing  sites  as  possible.  The  entire  Apollo  zone  of 
interest  was  ±45°  longitude  and  ±5°  latitude,  along  the  equatorial  region  of 
the  facing,  or  near  side  of  the  moon.  Within  that  zone,  the  area  that  could 
be  photographed  via  a  concentrated  mission  was  small.  A  single  Lunar 
Orbiter  that  could  photograph  10  sites  of  that  size  all  within  that  region 
would  be  much  more  effective.  If  the  data  showed  that  a  site  chosen  by  the 
astrogeologists  was  not  suitable,  NASA  would  have  excellent  photographic 
coverage  of  nine  other  prime  sites.  In  summary,  the  distributed  mode  would 

339 


Space/light  Revolution 


JETTISON 

ATLAS 

•OOSTH 


NUKING  OMIT) 


,  AGENA  SCNUATION 


TYPICAL  FLIGHT  SEQUENCE  OF  EVENTS 

L-66-5819 

Lunar  Orbiter's  "Typical  Flight  Sequence  of  Events"  turned  out  to  be  quite  typical 
indeed,  as  all  five  spacecraft  performed  exactly  as  planned. 


give  NASA  the  flexibility  to  ensure  that  Lunar  Orbiter  would  provide  the 
landing  site  information  needed  by  Apollo  even  if  only  one  Lunar  Orbiter 
mission  proved  successful. 

But  there  was  one  big  hitch:  Eastman  Kodak's  photo  system  was  not 
designed  for  the  distributed  mission.  It  was  designed  for  the  concentrated 
mission  in  which  all  the  photography  would  involve  just  one  site  and  be 
loaded,  shot,  and  developed  in  24  hours.  If  Lunar  Orbiter  must  photograph 
10  sites,  a  mission  would  last  at  least  two  weeks.  The  film  system  was 
designed  to  sustain  operations  for  only  a  day  or  two;  if  the  mission  lasted 
longer  than  that,  the  Bimat  film  would  stick  together,  the  exposed  parts  of 
it  would  dry  out,  the  film  would  get  stuck  in  the  loops,  and  the  photographic 
mission  would  be  completely  ruined. 

When  Boeing  first  heard  that  NASA  had  changed  its  mind  and  now 
wanted  to  do  the  distributed  mission,  Helberg  and  his  men  balked.  Accord- 
ing to  LOPO's  Norman  Crabill,  Boeing's  representatives  said,  "Look,  we  un- 
derstand you  want  to  do  this.  But,  wait.  The  system  was  designed,  tested, 
used,  and  proven  in  the  concentrated  mission  mode.  You  can't  change  it 

340 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

now  because  it  wasn't  designed  to  have  the  Bimat  film  in  contact  for  long 
periods  of  time.  In  two  weeks'  time,  some  of  the  Bimat  is  just  going  to  go, 
pfft!  It's  just  going  to  fail!"  Boeing  understood  the  good  sense  of  the  dis- 
tributed mission,  but  as  the  prime  contractor,  the  company  faced  a  classic 
technological  dilemma.  The  customer,  NASA,  wanted  to  use  the  system  to 
do  something  it  was  not  designed  to  do.  This  could  possibly  cause  a  disas- 
trous failure.  Boeing  had  no  recourse  but  to  advise  the  customer  that  what 
it  wanted  to  do  could  endanger  the  entire  mission.58 

The  Langley  engineers  wanted  to  know  whether  Boeing  could  solve  the 
film  problem.  "We  don't  know  for  sure,"  the  Boeing  staff  replied,  "and  we 
don't  have  the  time  to  find  out."  NASA  suggested  that  Boeing  conduct  tests 
to  obtain  quantitative  data  that  would  define  the  limits  of  the  film  system. 
Boeing's  response  was  "That's  not  in  the  contract."59  The  legal  documents 
specified  that  the  Lunar  Orbiter  should  have  the  capacity  to  conduct  the 
concentrated  mission.  If  NASA  now  wanted  to  change  the  requirements  for 
developing  the  Orbiter,  then  a  new  contract  would  have  to  be  negotiated.  A 
stalemate  resulted  on  this  issue  and  lasted  until  early  1965.  The  first  launch 
was  only  a  year  away. 

If  LOPO  hoped  to  persuade  Boeing  to  accept  the  idea  of  changing 
a  basic  mission  requirement,  it  had  to  know  the  difference  in  reliability 
between  the  distributed  and  concentrated  missions.  If  analysis  showed  that 
the  distributed  mission  would  be  far  less  reliable,  then  even  LOPO  might 
want  to  reconsider  and  proceed  with  the  concentrated  mission.  Crabill  gave 
the  job  of  obtaining  this  information  to  Tom  Young,  a  young  researcher 
from  the  Applied  Materials  and  Physics  Division.  Crabill  had  specifically 
requested  that  Young  be  reassigned  to  LOPO  mission  integration  because, 
in  his  opinion,  Young  was  "the  brightest  guy  [he]  knew."  On  the  day  Young 
had  reported  to  work  with  LOPO,  Crabill  had  given  him  "a  big  pile  of  stuff  to 
read,"  thinking  he  would  be  busy  and,  as  Crabill  puts  it,  "out  of  my  hair  for 
quite  a  while."  But  two  days  later,  Young  returned,  having  already  made  his 
way  through  all  the  material.  When  given  the  job  of  the  comparative  mission 
reliability  analysis,  Young  went  to  Boeing  in  Seattle.  In  less  than  two  weeks, 
he  found  what  he  needed  to  know  and  figured  out  the  percentages:  the 
reliability  for  the  concentrated  mission  was  an  unspectacular  60  percent, 
but  for  the  distributed  mission  it  was  only  slightly  worse,  58  percent.  "It 
was  an  insignificant  difference,"  Crabill  thought  when  he  heard  Young's 
numbers,  especially  because  nobody  then  really  knew  how  to  do  that  type 
of  analysis.  "We  didn't  gag  on  the  fact  that  it  was  pretty  low  anyway,  but 
we  really  wanted  to  do  this  distributed  mission."  The  Langley  researchers 
decided  that  the  distributed  mission  was  a  sensible  choice,  if  the  Kodak 
system  could  be  made  to  last  for  the  extra  time  and  if  Boeing  could  be 
persuaded  to  go  along  with  the  mission  change.60 

LOPO  hoped  that  Young's  analysis  would  prove  to  Boeing  that  no 
essential  difference  in  reliability  existed  between  the  two  types  of  missions, 
but  Boeing  continued  to  insist  that  the  concentrated  mission  was  the  legal 

341 


Space/light  Revolution 

requirement,  not  the  distributed  mission.  The  dispute  was  a  classic  case 
of  implementing  a  project  before  even  the  customer  was  completely  sure  of 
what  that  project  should  accomplish.  In  such  a  situation,  the  only  sensible 
thing  to  do  was  to  be  flexible. 

The  problem  for  Boeing,  of  course,  was  that  such  flexibility  might  cost 
the  company  its  financial  incentives.  If  a  Lunar  Orbiter  mission  failed,  the 
company  worried  that  it  would  not  be  paid  the  bonus  money  promised  in  the 
contract.  Helberg  and  Nelson  discussed  this  issue  in  private  conversations. 
Floyd  Thompson  participated  in  many  of  these  talks  and  even  visited  Seattle 
to  try  to  facilitate  an  agreement.  In  the  end,  Langley  convinced  Helberg 
that  the  change  from  a  concentrated  to  a  distributed  mission  would  not 
impact  Boeing's  incentives.  If  a  mission  failed  because  of  the  change,  LOPO 
promised  that  it  would  assume  the  responsibility.  Boeing  would  have  done 
its  best  according  to  the  government  request  and  instructions — and  for  that 
they  would  not  be  penalized.61 

The  missions,  however,  would  not  fail.  NASA  and  Boeing  would  handle 
the  technical  problems  involving  the  camera  by  testing  the  system  to 
ascertain  the  definite  limits  of  its  reliable  operation.  Prom  Kodak,  the 
government  and  the  prime  contractor  obtained  hard  data  regarding  the 
length  of  time  the  film  could  remain  set  in  one  place  before  the  curls 
or  bends  in  the  film  around  the  loops  became  permanent  and  the  torque 
required  to  advance  the  film  exceeded  the  capability  of  the  motor.  From 
these  tests,  Boeing  and  LOPO  established  a  set  of  mission  "rules"  that 
had  to  be  followed  precisely.  For  example,  to  keep  the  system  working, 
Lunar  Orbiter  mission  controllers  at  JPL  had  to  advance  the  film  one  frame 
every  eight  hours.  The  rules  even  required  that  film  sometimes  be  advanced 
without  opening  the  door  of  the  camera  lens.  Mission  controllers  called  these 
nonexposure  shots  their  "film-set  frames"  and  the  schedule  of  photographs 
their  "film  budget."62 

As  a  result  of  the  film  rules,  the  distributed  mission  turned  out  to  be  a 
much  busier  operation  than  a  concentrated  mission  would  have  been.  Each 
time  a  photograph  was  taken,  including  film-set  frames,  the  spacecraft  had  to 
be  maneuvered.  Each  maneuver  required  a  command  from  mission  control. 
LOPO  staff  worried  about  the  ability  of  the  spacecraft  to  execute  so  many 
maneuvers  over  such  a  prolonged  period.  They  feared  something  would 
go  wrong  during  a  maneuver  that  would  cause  them  to  lose  control  of  the 
spacecraft.  Lunar  Orbiter  /,  however,  flawlessly  executed  an  astounding 
number  of  commands,  and  LOPO  staff  were  able  to  control  spacecraft 
attitude  during  all  374  maneuvers.63 

Ultimately,  the  trust  between  Langley  and  Boeing  allowed  each  to  take 
the  risk  of  changing  to  a  distributed  mission.  Boeing  trusted  Langley  to 
assume  responsibility  if  the  mission  failed,  and  Langley  trusted  Boeing  to 
put  its  best  effort  into  making  the  revised  plan  a  success.  Had  either  not 
fulfilled  its  promise  to  the  other,  Lunar  Orbiter  would  not  have  achieved  its 
outstanding  record. 

342 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


Simple  as  this  diagram  of  Lunar  Orbiter 
(left)  may  look,  no  spacecraft  in  NASA  his- 
tory operated  more  successfully  than  Lunar 
Orbiter.  Below,  Lunar  Orbiter  I  goes 
through  a  final  inspection  in  the  NASA 
Hanger  S  clean  room  at  Kennedy  Space 
Center  prior  to  launch  on  10  August  1966. 
The  spacecraft  was  mounted  on  a  three- axis 
test  stand  with  its  solar  panels  deployed 
and  high-gain  dish  antenna  extended  from 
the  side. 


P36982 


L-66-6387 


343 


Spaceflight  Revolution 


Lunar  Orbiter  I  lifts  off  from  Cape 
Kennedy  on  10  August  1966.  With  a 
payload  weighing  only  860  pounds,  the 
spacecraft  was  light  enough  to  fly  on 
an  Atlas- Agena  instead  of  the  more 
expensive  Atlas-Centaur. 


L-66-6381 


"The  Picture  of  the  Century" 

The  switch  to  the  distributed  mission  was  not  the  only  instance  during 
the  Lunar  Orbiter  mission  when  contract  specifications  were  jettisoned  to 
pursue  a  promising  idea.  Boeing  engineers  realized  that  the  Lunar  Orbiter 
project  presented  a  unique  opportunity  for  photographing  the  earth.  When 
the  LOPO  staff  heard  this  idea,  they  were  all  for  it,  but  Helberg  and 
Boeing  management  rejected  the  plan.  Turning  the  spacecraft  around  so 
that  its  camera  could  catch  a  quick  view  of  the  earth  tangential  to  the 
moon's  surface  entailed  technical  difficulties,  including  the  danger  that, 
once  the  spacecraft's  orientation  was  changed,  mission  controllers  could 
lose  command  of  the  spacecraft.  Despite  the  risk,  NASA  urged  Boeing  to 
incorporate  the  maneuver  in  the  mission  plan  for  Lunar  Orbiter  I.  Helberg 
refused.64 

In  some  projects,  that  might  have  been  the  end  of  the  matter.  Peo- 
ple would  have  been  forced  to  forget  the  idea  and  to  live  within  the  cir- 
cumscribed world  of  what  had  been  legally  agreed  upon.  Langley,  how- 
ever, was  not  about  to  give  up  on  this  exciting  opportunity.  Cliff  Nelson, 

344 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-66-7999 

With  "the  picture  of  the  century"  proudly  displayed  before  them,  key  members  of 
the  LOPO  team  report  the  success  of  Lunar  Orbiter  I  at  a  press  conference  in 
August  1966.  Left  to  right  are  Oran  W.  Nicks,  director  of  Lunar  and  Planetary 
Programs  at  NASA  headquarters;  Floyd  Thompson;  Cliff  Nelson;  and  Isadore 
G.  Recant,  the  Langley  scientist  in  charge  of  data  handling  for  the  spacecraft.  At 
the  podium  is  the  U.S.  Geological  Survey's  Dr.  Larry  Rowan,  the  young  geologist 
who  helped  LOPO  identify  the  most  promising  landing  sites. 

Floyd  Thompson,  and  Lee  Scherer  went  to  mission  control  at  JPL  to  talk 
to  Helberg  and  at  last  convinced  him  that  he  was  being  too  cautious — that 
"the  picture  was  worth  the  risk."  If  any  mishap  occurred  with  the  space- 
craft during  the  maneuver,  NASA  again  promised  that  Boeing  would  still 
receive  compensation  and  part  of  its  incentive  for  taking  the  risk.  The  en- 
thusiasm of  his  own  staff  for  the  undertaking  also  influenced  Helberg  in  his 
final  decision  to  take  the  picture.65 

On  23  August  1966  just  as  Lunar  Orbiter  /was  about  to  pass  behind  the 
moon,  mission  controllers  executed  the  necessary  maneuvers  to  point  the 
camera  away  from  the  lunar  surface  and  toward  the  earth.  The  result  was 
the  world's  first  view  of  the  earth  from  space.  It  was  called  "the  picture  of  the 
century"  and  "the  greatest  shot  taken  since  the  invention  of  photography."  * 


jk 

The  unprecedented  photo  also  provided  the  first  oblique  perspectives  of  the  lunar  surface.  All 
other  photographs  taken  during  the  first  mission  were  shot  from  a  position  perpendicular  to  the  surface 
and  thus,  did  not  depict  the  moon  in  three  dimensions.  In  subsequent  missions,  NASA  made  sure  to 
include  this  sort  of  oblique  photography.  Following  the  first  mission,  Boeing  prepared  a  booklet  entitled 
Lunar  Orbiter  I — Photography  (NASA  Langley,  1968),  which  gave  a  detailed  technical  description  of  the 
earth- moon  photographs;  see  especially  pp.  64-71. 

345 


Space/light  Revolution 

Not  even  the  color  photos  of  the  earth  taken  during  the  Apollo  missions 
superseded  the  impact  of  this  first  image  of  our  planet  as  a  little  island  of 
life  floating  in  the  black  and  infinite  sea  of  space.66 


Mission  More  Than  Accomplished 

Lunar  Orbiter  defied  all  the  probability  studies.  All  five  missions  worked 
extraordinarily  well,  and  with  the  minor  exception  of  a  short  delay  in  the 
launch  of  Lunar  Orbiter  I — the  Eastman  Kodak  camera  was  not  ready- 
all  the  missions  were  on  schedule.  The  launches  were  three  months  apart 
with  the  first  taking  place  in  August  1966  and  the  last  in  August  1967. 
This  virtually  perfect  flight  record  was  a  remarkable  achievement,  especially 
considering  that  Langley  had  never  before  managed  any  sort  of  flight 
program  into  deep  space. 

Lunar  Orbiter  accomplished  what  it  was  designed  to  do,  and  more.  Its 
camera  took  1654  photographs.  More  than  half  of  these  (840)  were  of  the 
proposed  Apollo  landing  sites.  Lunar  Orbiter s  /,  //,  and  ///  took  these  site 
pictures  from  low-flight  altitudes,  thereby  providing  detailed  coverage  of 
22  select  areas  along  the  equatorial  region  of  the  near  side  of  the  moon.  One 
of  the  eight  sites  scrutinized  by  Lunar  Orbiters  II  and  ///  was  a  very  smooth 
area  in  the  Sea  of  Tranquility.  A  few  years  later,  in  July  1969,  Apollo  11 
commander  Neil  Armstrong  would  navigate  the  lunar  module  Eagle  to  a 
landing  on  this  site.67 

By  the  end  of  the  third  Lunar  Orbiter  mission,  all  the  photographs  needed 
to  cover  the  Apollo  landing  sites  had  been  taken.  NASA  was  then  free  to 
redesign  the  last  two  missions,  move  away  from  the  pressing  engineering 
objective  imposed  by  Apollo,  and  go  on  to  explore  other  regions  of  the 
moon  for  the  benefit  of  science.  Eight  hundred  and  eight  of  the  remaining 
814  pictures  returned  by  Lunar  Orbiters  IV  and  V  focused  on  the  rest  of  the 
near  side,  the  polar  regions,  and  the  mysterious  far  side  of  the  moon.  These 
were  not  the  first  photographs  of  the  "dark  side";  a  Soviet  space  probe, 
Zond  III,  had  taken  pictures  of  it  during  a  fly-by  into  a  solar  orbit  a  year 
earlier,  in  July  1965.  But  the  Lunar  Orbiter  photos  were  higher  quality 
than  the  Russian  pictures  and  illuminated  some  lunarscapes  that  had  never 
before  been  seen  by  the  human  eye.  The  six  remaining  photos  were  of  the 
spectacular  look  back  at  the  distant  earth.  By  the  time  all  the  photos  were 
taken,  about  99  percent  of  the  moon's  surface  had  been  covered. 

When  each  Lunar  Orbiter  completed  its  photographic  mission,  the 
spacecraft  continued  its  flight  to  gather  clues  to  the  nature  of  the  lunar 
gravitational  environment.  NASA  found  these  clues  valuable  in  the  planning 
of  the  Apollo  flights.  Telemetry  data  clearly  indicated  that  the  moon's 
gravitational  pull  was  not  uniform.  The  slight  dips  in  the  path  of  the  Lunar 
Orbiters  as  they  passed  over  certain  areas  of  the  moon's  surface  were  caused 
by  gravitational  perturbations,  which  in  turn  were  caused  by  the  mascons. 

346 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-67-6702 

Lunar  Orbiter  V  provided  the  first  wide-angle  view  of  the  moon's  mysterious  "dark 
side. " 


347 


Spaceflight  Revolution 


L-66-9788 

One  of  the  most  spectacular  mosaics  produced  by  Lunar  Orbiter  II  was  this  close-up 
of  the  enormous  crater  Copernicus  with  its  300-meter-high  (984-3  feet)  mountains 
rising  from  the  crater  floor.  On  the  horizon  are  the  Carpathian  mountains  with  the 
920-meter-high  (3018.4  feet)  Guy-Lussac  Promontory. 

The  extended  missions  of  the  Lunar  Orbiters  also  helped  to  confirm  that 
radiation  levels  near  the  moon  were  quite  low  and  posed  no  danger  to 
astronauts  unless  a  major  solar  flare  occurred  while  they  were  exposed  on 
the  lunar  surface.  A  few  months  after  each  Lunar  Orbiter  mission,  NASA 
deliberately  crashed  the  spacecraft  into  the  lunar  surface  to  study  lunar 
impacts  and  their  seismic  consequences.  Destroying  the  spacecraft  before  it 
deteriorated  and  mission  controllers  had  lost  command  of  it  ensured  that  it 
would  not  wander  into  the  path  of  some  future  mission.68 

Whether  the  Apollo  landings  could  have  been  made  successfully  without 
the  photographs  from  Lunar  Orbiter  is  a  difficult  question  to  answer. 
Without  the  photos,  the  manned  landings  could  certainly  still  have  been 
attempted.  In  addition  to  the  photographic  maps  drawn  from  telescopic 
observation,  engineers  could  use  some  good  pictures  taken  from  Ranger  and 
Surveyor  to  guide  them.  However,  the  detailed  photographic  coverage  of 
22  possible  landing  sites  definitely  made  NASA's  final  selection  of  ideal  sites 
much  easier  and  the  pinpointing  of  landing  spots  possible. 


348 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 


L-67-4023 

In  August  1967,  Lunar  Orbiter  V  photographed  the  90-kilometer-wide  (55.9  miles) 
Tycho  crater,  one  of  the  brightest  craters  seen  from  earth.  A  young  impact  crater, 
Tycho  reveals  its  central  peak,  rough  floor,  and  precipitous  walls. 

Furthermore,  Lunar  Orbiter  also  contributed  important  photometric 
information  that  proved  vital  to  the  Apollo  program.  Photometry  involves 
the  science  of  measuring  the  intensity  of  light.  Lunar  Orbiter  planners  had 
to  decide  where  to  position  the  camera  to  have  the  best  light  for  taking 
the  high-resolution  photographs.  When  we  take  pictures  on  earth,  we 
normally  want  to  have  the  sun  behind  us  so  it  is  shining  directly  on  the 
target.  But  a  photo  taken  of  the  lunar  surface  in  these  same  circumstances 
produces  a  peculiar  photometric  function:  the  moon  looks  flat.  Even 
minor  topographical  features  are  indistinguishable  because  of  the  intensity 
of  the  reflecting  sunlight  from  the  micrometeorite-filled  lunar  surface.  The 
engineers  in  LOPO  had  to  determine  the  best  position  for  photographing 
the  moon.  After  studying  the  problem  (Taback,  Crabill,  and  Young  led  the 
attack  on  this  problem),  LOPO's  answer  was  that  the  sun  should  indeed  be 
behind  the  spacecraft,  but  photographs  should  be  taken  when  the  sun  was 
only  15  degrees  above  the  horizon.69 

Long  before  it  was  time  for  the  first  Apollo  launch,  LOPO's  handling  of 
the  lunar  photometric  function  was  common  knowledge  throughout  NASA 

349 


Space/light  Revolution 

and  the  aerospace  industry.  The  BellComm  scientists  and  engineers  who 
reviewed  Apollo  planning  quickly  realized  that  astronauts  approaching  the 
moon  to  make  a  landing  needed,  like  Lunar  Orbiter,  to  be  in  the  best 
position  for  viewing  the  moon's  topography.  Although  a  computer  program 
would  pinpoint  the  Apollo  landing  site,  the  computer's  choice  might  not  be 
suitable.  If  that  was  the  case,  astronauts  would  have  to  rely  on  their  own 
eyes  to  choose  a  spot.  If  the  sun  was  in  the  wrong  position,  they  would  not 
make  out  craters  and  boulders,  the  surface  would  appear  deceptively  flat, 
and  the  choice  might  be  disastrous.  Apollo  11  commander  Neil  Armstrong 
did  not  like  the  spot  picked  by  the  computer  for  the  Eagle  landing.  Because 
NASA  had  planned  for  him  to  be  in  the  best  viewing  position  relative  to  the 
sun,  Armstrong  could  see  that  the  place  was  "littered  with  boulders  the  size 
of  Volkswagens."  So  he  flew  on.  He  had  to  go  another  1500  meters  before 
he  saw  a  spot  where  he  could  set  the  lunar  module  down  safely.70 

NASA  might  have  considered  the  special  photometric  functions  involved 
in  viewing  the  moon  during  Apollo  missions  without  Lunar  Orbiter,  but 
the  experience  of  the  Lunar  Orbiter  missions  took  the  guesswork  out  of  the 
calculations.  NASA  knew  that  its  astronauts  would  be  able  to  see  what  they 
needed  to  see  to  avoid  surface  hazards.  This  is  a  little-known  but  important 
contribution  from  Lunar  Orbiter. 


Secrets  of  Success 

In  the  early  1970s  Erasmus  H.  Kloman,  a  senior  research  associate  with 
the  National  Academy  of  Public  Administration,  completed  an  extensive 
comparative  investigation  of  NASA's  handling  of  its  Surveyor  and  Lunar 
Orbiter  projects.  After  a  lengthy  review,  NASA  published  a  shortened 
and  distilled  version  of  Kloman 's  larger  study  as  Unmanned  Space  Project 
Management:  Surveyor  and  Lunar  Orbiter.  The  result — even  in  the 
expurgated  version,  with  all  names  of  responsible  individuals  left  out — was 
a  penetrating  study  in  "sharp  contrasts"  that  should  be  required  reading  for 
every  project  manager  in  business,  industry,  or  government. 

Based  on  his  analysis  of  Surveyor  and  Lunar  Orbiter,  Kloman  concluded 
that  project  management  has  no  secrets  of  success.  The  key  elements  are 
enthusiasm  for  the  project,  a  clear  understanding  of  the  project's  objective, 
and  supportive  and  flexible  interpersonal  and  interoffice  relationships.  The 
history  of  Surveyor  and  Lunar  Orbiter,  Kloman  wrote,  "serves  primarily 
as  a  confirmation  of  old  truths  about  the  so-called  basic  principles  of 
management  rather  than  a  revelation  of  new  ones."  Kloman  writes  that 
Langley  achieved  Lunar  Orbiter's  objectives  by  "playing  it  by  the  book." 
By  this,  Kloman  meant  that  Langley  applied  those  simple  precepts  of 
good  management;  he  did  not  mean  that  success  was  achieved  through 
a  thoughtless  and  strict  formula  for  success.  Kloman  understood  that 
Langley's  project  engineers  broke  many  rules  and  often  improvised  as 
they  went  along.  Enthusiasm,  understanding,  support,  and  flexibility 

350 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

allowed  project  staff  to  adapt  the  mission  to  new  information,  ideas,  or 
circumstances.  "Whereas  the  Surveyor  lessons  include  many  illustrations  of 
how  'not  to'  set  out  on  a  project  or  how  to  correct  for  early  misdirections," 
Kloman  argued,  "Lunar  Orbiter  shows  how  good  sound  precepts  and 
directions  from  the  beginning  can  keep  a  project  on  track."71 

Lunar  Orbiter,  however,  owes  much  of  its  success  to  Surveyor.  LOPO 
staff  were  able  to  learn  from  the  mistakes  made  in  the  Surveyor  project. 
NASA  headquarters  was  responsible  for  some  of  these  mistakes.  The 
complexity  of  Surveyor  was  underestimated,  unrealistic  manpower  and 
financial  ceilings  were  imposed,  an  "unreasonably  open-ended  combination 
of  scientific  experiments  for  the  payload"  was  insisted  upon  for  too  long, 
too  many  changes  in  the  scope  and  objectives  of  the  project  were  made, 
and  the  project  was  tied  to  the  unreliable  Centaur  launch  vehicle.72  NASA 
headquarters  corrected  these  mistakes.  In  addition,  Langley  representatives 
learned  from  JPL's  mistakes  and  problems.  They  talked  at  great  length 
to  JPL  staff  in  Pasadena  about  Surveyor  both  before  and  after  accepting 
the  responsibility  for  Lunar  Orbiter.  Prom  these  conversations,  Langley 
acquired  a  great  deal  of  knowledge  about  the  design  and  management  of  an 
unmanned  space  mission.  JPL  scientists  and  engineers  even  conducted  an 
informal  "space  school"  that  helped  to  educate  several  members  of  LOPO 
and  Boeing's  team  about  key  details  of  space  mission  design  and  operations. 

The  interpersonal  skills  of  the  individuals  responsible  for  Lunar  Orbiter, 
however,  appear  to  have  been  the  essential  key  to  success.  These  skills  cen- 
tered more  on  the  ability  to  work  with  other  people  than  they  did  on  what 
one  might  presume  to  be  the  more  critical  and  esoteric  managerial,  concep- 
tual, and  technical  abilities.  In  Kloman's  words,  "individual  personal  qual- 
ities and  management  capabilities  can  at  times  be  a  determining  influence 
in  overall  project  performance."73  Compatibility  among  individual  man- 
agers, Nelson  and  Helberg,  and  the  ability  of  those  managers  to  stimulate 
good  working  relationships  between  people  proved  a  winning  combination 
for  Lunar  Orbiter. 

Norman  Crabill  made  these  comments  about  Lunar  Orbiter's  manage- 
ment: "We  had  some  people  who  weren't  afraid  to  use  their  own  judgment 
instead  of  relying  on  rules.  These  people  could  think  and  find  the  essence 
of  a  problem,  either  by  discovering  the  solution  themselves  or  energizing 
the  troops  to  come  up  with  an  alternative  which  would  work.  They  were 
absolute  naturals  at  that  job." 74 

Lunar  Orbiter  was  a  pathfinder  for  Apollo,  and  it  was  an  outstanding 
contribution  by  Langley  Research  Center  to  the  early  space  program.  The 
old  NACA  aeronautics  laboratory  proved  not  only  that  it  could  handle  a 
major  deep  space  mission,  but  also  that  it  could  achieve  an  extraordinary 
record  of  success  that  matched  or  surpassed  anything  yet  tried  by  NASA. 
When  the  project  ended  and  LOPO  members  went  back  into  functional 
research  divisions,  Langley  possessed  a  pool  of  experienced  individuals  who 
were  ready,  if  the  time  came,  to  plan  and  manage  yet  another  major 

351 


Space/light  Revolution 


L-67-6777 

Although  most  people  have  come  to  associate  the  first  picture  of  the  "Whole  Earth" 
with  the  Apollo  program,  Lunar  Orbiter  V  actually  captured  this  awesome  view  of 
the  home  planet.  When  this  picture  was  taken  on  8  August  1967,  the  spacecraft  was 
about  5860  kilometers  (3641-2  miles)  above  the  moon  in  near  polar  orbit,  so  that  the 
lunar  surface  is  not  seen.  Clearly  visible  on  the  left  side  of  the  globe  is  the  eastern 
half  of  Africa  and  the  entire  Arabian  peninsula. 


352 


To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

project.  That  opportunity  came  quickly  in  the  late  1960s  with  the  inception 
of  Viking,  a  much  more  complicated  and  challenging  project  designed  to 
send  unmanned  reconnaissance  orbiters  and  landing  probes  to  Mars.  When 
Viking  was  approved,  NASA  headquarters  assigned  the  project  to  "those 
plumbers"  at  Langley.  The  old  LOPO  team  formed  the  nucleus  of  Langley's 
much  larger  Viking  Project  Office.  With  this  team,  Langley  would  once 
again  manage  a  project  that  would  be  virtually  an  unqualified  success. 


353 


11 


In  the  Service  of  Apollo 


We  were  working  beyond  the  state  of  the  art.  Nobody 
had  done  things  like  this  before. 

— E.  Barton  Geer,  associate  chief 
of  Langley's  Flight  Vehicle  and 
Systems  Division  during  the 
Apollo  era  and  member  of  the 
Apollo  204  Review  Board 

And  just  as  we  got  to  the  transonic  field,  then  all  of 
a  sudden  we  opened  up  with  the  supersonic  field  and 
find  out  we  're  flying — militarily  anyway — we  're  flying 
at  speeds  of  [MachJ  2  and  3.  And  you  just  get  that 
pretty  well  understood  and,  Holy  Smoke,  here  we  are 
going  to  the  Moon  and  things  like  that. 

—Floyd  L.  Thompson,  Langley  director 
and  chairman  of  the  Apollo  204 
Review  Board 


The  crowning  moment,  as  well  as  the  denouement,  of  the  spaceflight 
revolution  came  at  4:18  p.m.  on  Sunday,  20  July  1969,  when  American  as- 
tronauts Neil  A.  Armstrong  and  Edwin  E.  "Buzz"  Aldrin,  Jr.,  made  the 
first  manned  lunar  landing.  The  realization  of  this  spectacular  moment 
required  the  most  sudden  burst  of  technological  creativity  and  the  largest 
commitment  of  resources  ever  made  by  any  nation  in  peacetime:  an  esti- 
mated $24  billion.  At  its  peak  the  Apollo  program  employed  approximately 
400,000  Americans  and  enlisted  the  support  of  over  20,000  industrial  firms 
and  universities.  As  President  Kennedy  had  said  in  his  May  1961  speech, 
"It  will  not  be  one  man  going  to  the  moon — it  will  be  an  entire  nation.  For 
all  of  us  must  work  to  put  him  there." l 

355 


Space/light  Revolution 

"All  of  us"  also  meant  all  the  NASA  facilities,  including  the  research 
centers.  To  be  sure,  Langley  did  not  serve  as  the  heart  of  the  Apollo 
program,  as  it  temporarily  had  for  the  man-in-space  effort  before  the  STG 
left  for  Texas.  Apollo  would  not  be  managed  by  any  of  the  field  centers 
but  by  a  well-staffed  central  program  office  within  the  Office  of  Manned 
Space  Flight  at  NASA  headquarters.  Of  course,  the  spaceflight  centers  (the 
Manned  Spacecraft  Center  in  Houston  and  the  Marshall  Space  Flight  Center 
in  Huntsville)  were  deeply  involved.  A  large,  well-funded  Apollo  program 
office  at  Houston  was  responsible  for  the  budget,  schedule,  technical  design, 
and  production  of  the  three- module  Apollo  spacecraft;  moreover,  Houston 
was  the  home  of  Mission  Control,  the  nerve  center  of  NASA's  manned  flight 
operations  and  the  place  where  all  the  news  reporters  went  after  the  launch 
at  Cape  Kennedy.  At  Marshall  the  von  Braun  team  handled  the  awesome 
task  of  developing  the  giant  Saturn  rocket.2  These  were  the  two  NASA 
centers  where  the  staff  "lived  and  breathed"  Apollo.  But  neither  Langley 
nor  any  of  the  other  NASA  facilities  were  left  out.  Nor  could  they  be.  There 
was  too  much  work  to  do,  too  much  to  learn,  and  too  little  time.  All  NASA 
centers  eventually  became  heavily  involved  in  the  program,  and  with  the 
exception  of  Houston  and  Huntsville,  none  was  more  involved  than  Langley. 

Yet,  when  it  came  to  the  Apollo  flights  themselves  and  their  worldwide 
publicity,  Langley  and  the  other  centers  were  not  part  of  the  big  show.  By 
the  time  the  Apollo  spacecraft  sat  atop  the  huge  Saturn  V  at  Launch  Com- 
plex 39A  at  Kennedy  Space  Center  in  July  1969,  Langley's  contributions  to 
the  Apollo  program  had  already  been  made  and  mostly  forgotten.  That, 
after  all,  was  the  purpose  and  the  predicament  of  a  research  center:  to  lay 
the  groundwork  for  the  technological  achievements  that  others  would  pursue 
and  for  which  others  would  receive  the  credit.  Of  course,  John  Houbolt's 
concept  of  LOR  would  not  be  forgotten  in  the  dramatic  days  of  Apollo  11, 
nor  would  the  flights  of  the  Lunar  Orbiter  spacecraft.  But  many  other  el- 
emental tasks  that  Langley  had  done  in  the  service  of  Apollo  would  not 
be  remembered:  the  basic  rendezvous  and  docking  studies,  the  wind-tunnel 
investigations  of  the  aerodynamic  integrity  of  the  Saturn- Apollo  launch  com- 
bination, the  work  on  reentry  heating  and  its  potentially  fatal  effects  on  the 
returning  Apollo  spacecraft,  and  the  simulation  training  that  helped  prepare 
the  astronauts  not  only  for  the  rendezvous  and  docking  in  space  but  also 
for  the  actual  landing  of  a  manned  spacecraft  and  for  astronaut  locomotion 
activities  on  the  moon.  From  launch  to  splashdown,  there  was  no  aspect 
of  the  Apollo  mission  that  scientists,  engineers,  and  technicians  at  Langley 
had  not  helped  to  develop  in  one  way  or  another.  All  this  work,  however, 
had  been  done  long  before  the  two  Apollo  astronauts  skillfully  maneuvered 
their  lunar  module  Eagle  down  to  the  Sea  of  'Tranquility  on  that  historic 
day  in  July  1969.  All  that  Langley  employees  could  do  on  that  hot  Sunday 
afternoon  in  July  was  sit  in  front  of  televisions  in  their  own  living  rooms, 
sip  cool  glasses  of  lemonade,  and  applaud  with  the  rest  of  the  country  what 
everyone  together  had  accomplished. 

356 


In  the  Service  of  Apollo 
Langley's  "Undercover  Operation"  in  Houston 

Langley  was  in  a  novel  situation  during  the  heyday  of  Apollo.  It  was 
the  original  (and  for  over  20  years  only)  NACA  center,  not  to  mention  the 
"mother"  of  many  of  the  other  centers  (including  the  Manned  Spacecraft 
Center  in  Houston),  but  during  the  Apollo  period,  it  was  relegated  to 
the  periphery  of  NASA's  most  urgent  task.  Langley  management  was  not 
accustomed  to  being  in  a  marginal  position  and  did  not  especially  like  being 
there. 

Members  of  the  local  Hampton  elite  did  not  like  it  either.  For  them  and 
other  groups  of  Langley  supporters,  the  unfortunate  situation  was  the  work 
of  conniving  politicians  who  had  stolen  their  precious  STG  for  Texas.  The 
bitterness  over  this  did  not  fade  quickly.  On  occasion,  concern  for  Langley's 
displacement  led  those  who  knew  better  to  make  imprudent  remarks.  For 
example,  in  December  1961  during  James  Webb's  first  visit  to  Langley  after 
becoming  NASA  administrator,  Floyd  Thompson,  as  savvy  a  person  as  ever 
sat  in  the  Langley  director's  chair,  mentioned  to  Webb  that  the  area's  city 
fathers  were  wondering  exactly  what  they  might  expect  in  terms  of  new 
jobs  and  government  contracts  as  a  result  of  the  lunar  landing  program. 
Webb  responded  sharply  with  a  statement  that  echoed  the  already  famous 
line  from  President  Kennedy's  inaugural  address:  "The  city  fathers  should 
be  asking  not  what  Apollo  can  do  for  them,  but  what  they  can  do  for 
Apollo."3  Thompson  never  forgot  the  sting  of  this  retort  and  resolved 
never  to  give  Webb  or  anyone  else  reason  to  question  the  effort  that  his 
own  research  center  was  putting  into  the  nation's  lunar  landing  program. 

At  a  meeting  of  NASA  center  directors  and  program  office  directors  held 
in  the  so-called  control  room  on  the  top  floor  of  the  NASA  headquarters 
building  in  early  1962,  the  question  of  Langley's  contribution  to  Apollo  did 
come  up.  For  Jim  Webb,  Apollo  was  everything,  and  he  wanted  assurances 
from  the  center  directors  that  they  were  doing  all  they  could  to  support  the 
program.  Characteristically,  Thompson  bided  his  time,  waiting  while  his 
counterparts  at  the  other  NASA  centers  answered.  Then  in  his  deliberate 
and  rather  high-pitched  midwestern  voice,  he  reported  with  confidence, 
"Well,  we  have  a  senior  man  in  Houston  who  keeps  track  of  all  that  we're 
doing  for  it,  and  is  on  the  spot  ready  to  contact  us  whenever  anyone  down 
there  wants  us  to  do  anything  more."4  Such  a  remark  was  typical  "Tommy" 
Thompson,  and  Webb  let  it  pass.  He  had  no  idea  that  such  a  thing  was  going 
on  down  in  Houston  or  that  the  Houston  organization  would  even  allow  it, 
but  he  took  Thompson  at  his  word. 

The  Langley  center  director  did  have  a  "senior  man"  on  duty  in  Houston 
and  only  in  part  so  that  he  could  give  such  smart  answers  to  "Big  Jim"  Webb. 
In  April  1962,  which  was  a  few  months  before  the  STG  had  completed  its 
move  to  the  Southwest,  Thompson  had  dispatched  Langley  veteran  Axel 
T.  Mattson  to  Houston  to  serve  as  research  assistant  for  Manned  Spacecraft 
Projects.  The  experienced  assistant  chief  of  the  Full-Scale  Research  Division, 

357 


Spaceflight  Revolution 


Under  the  wily  direction  of  Floyd 
Thompson  (left),  Langley  did  every- 
thing it  could  to  support  Project 
Apollo  and  satisfy  Administrator 
Jim  Webb  (right). 

L-61-8539 


Mattson  was  to  report  his  findings  to  Langley  Associate  Director  Charles 
Donlan,  Thompson's  right-hand  man.5 

This  was  an  interesting  and  unusual  development;  no  other  NASA  center 
had  such  an  arrangement,  and  certainly  no  other  center  but  Langley  could 
have  gotten  away  with  it.  Langley  had  an  official  explanation  of  the 
purpose  of  this  liaison:  "to  create  a  mechanism  for  the  timely  exchange 
of  information  on  manned  space  programs  and  projects  of  mutual  interest 
to  the  Langley  Center  and  the  Manned  Spacecraft  Center  and  to  provide  a 
means  for  quickly  initiating  action  at  Langley  as  may  be  required  in  support 
of  manned  spacecraft  projects."6  In  truth,  Mattson  was  acting  as  a  sort  of 
spy — a  Langley  agent. 

"I  was  running  an  undercover  operation,  really,  in  the  technical  sense," 
Mattson  remembers,  his  eyes  glimmering.  "Here  we  were  a  research  outfit 
trying  to  get  involved  more  directly  not  only  with  the  Apollo  work  but  with 
the  big  Apollo  money.  It  had  to  do  with  the  way  the  funding  was  set  up;  this 
was  the  big  controlling  factor.  For  everything  Langley  was  doing  in  support 
of  the  Office  of  Manned  Space  Flight  and  Apollo,  it  was  my  assignment 
to  try  to  get  a  transfer  of  funds  from  Houston,  which  was  getting  a  lot  of 
money,  to  Langley,  which  wasn't."  The  scheme  seems  a  bit  dishonest,  but 
Mattson  liked  even  that  aspect  of  his  work.  "I  was  bootlegging  material. 

358 


In  the  Service  of  Apollo 

I  was  transferring  certain  material  [from  Houston  to  Langley]  without  any 
official  authorization."  At  Langley,  Donlan  made  it  clear  to  Mattson  that 
he  should  not  initiate  any  work  on  Apollo  for  which  Langley  would  have  to 
go  running  to  NASA  headquarters  for  approval  and  funding.  Donlan  had 
told  him,  "Get  it  from  Houston."  That  was  the  real  reason  Langley  had 
quietly  sent  Axel  Mattson  to  Houston  in  April  1962.7 

At  the  same  time  that  Americans  were  becoming  enamored  with  the 
much  exaggerated  romance  of  spies  through  the  first  James  Bond  movies, 
the  free-spirited  Mattson  fell  in  love  with  the  intrigues  of  his  espionage 
work.  What  made  it  "so  beautiful"  for  him  was  the  freewheeling  nature  of 
the  assignment — the  freedom  to  innovate  and  take  some  chances.  "I  didn't 
fool  with  protocol  or  with  the  big  shots.  I  dealt  almost  exclusively  with  the 
troops  in  the  field  who  would  be  honest  and  open  with  me."  During  the 
working  day,  he  walked  from  office  to  office,  starting  conversations,  finding 
out  what  was  going  on,  what  the  problems  were,  and  what  Langley  might  be 
able  to  do  about  them.  In  the  evening,  he  socialized  at  the  many  Houston 
area  cocktail  and  dinner  parties,  where  he  made  the  important  personal 
contacts  that  led  to  office  meetings  the  next  day.  Besides  enjoying  convivial 
relations  with  his  Houston  buddies,  many  of  whom  he  had  known  at  Langley, 
Mattson  also  developed  close  professional  and  social  relationships  with  many 
of  the  NASA  contractor  representatives.  These  relationships  allowed  him  to 
tap  a  gushing  pipeline  of  Apollo-related  technical  data.8 

Through  this  technological  espionage,  Mattson  was  able  to  obtain  valu- 
able information  for  Langley.  Whenever  an  important  new  technical  mat- 
ter was  being  discussed  by  Houston  engineers  and  industry  representatives, 
Mattson  would  say,  "Gee,  I'm  awfully  interested  in  that,"  and  those  involved 
would  give  him  all  the  information  they  had.  "You  wouldn't  believe  it.  I 
went  over  to  their  offices  the  next  day  and  they  just  gave  it  to  me,"  all 
the  data,  all  the  technical  reports.  Conversely,  "when  they  wanted  some 
information,  they  called  me  and  I'd  get  it.  It  would  come  in  my  suitcase 
[from  Langley]  and  I'd  go  peddle  it  out."  Much  of  it  was  raw  data,  "no  for- 
mality, nothing  to  it,"  a  transfer  of  basic  and  sometimes  even  unprocessed 
information  that  went  unrecorded  and  was  never  revealed  in  any  formal  doc- 
umentation. Mattson  tried  to  feel  out  "when  a  guy  was  just  having  troubles 
even  with  calculations.  He  would  be  grunting  them  out  and  not  too  sure 
of  the  inputs,  and  I'd  bring  them  back  to  Langley  and  have  the  guys  look 
at  it  and  say,  'How  about  critiquing  it  and  see  if  you  can  change  numbers 
by  using  the  latest  test  information'  and  things  like  that."  It  was  basically 
"a  back-scratching  operation."  Mattson  sometimes  went  even  farther.  He 
admits  now  that  he  occasionally  would  "take  a  man's  calculations  right  off 
his  desk,  make  a  carbon  copy,"  without  the  man's  knowledge,  and  put  it  in 
his  suitcase  or  in  the  mail  for  Langley.9 

"Let  me  tell  you,  my  suitcase  was  full  whenever  I  came  back  to  Langley," 
not  only  with  papers  but  also  with  the  slides  and  films  that  were  being 
produced  at  Houston  by  the  carton.  Such  goodies  made  Mattson  something 

359 


Space/light  Revolution 

of  a  star  attraction  at  Langley  department  meetings  and  senior  staff  get- 
togethers  where  he  provided  some  special  entertainment.  At  some  point 
in  such  proceedings,  Mattson  remembers,  Floyd  Thompson  would  typically 
pause  and  joke,  "  'Well,  I  guess  we  might  just  as  well  see  if  Mattson's  got 
another  film  to  show  us.'  And  sure  enough,  I  had  another  film  to  show  them, 
or  slides,  or  something."10  This  was  the  way  Langley  kept  up  with  what 
was  going  on  in  Houston  and  minimized  the  loss  of  its  STG  to  Texas. 

The  other  NASA  centers,  with  no  direct  ties  to  the  STG,  could  not  have 
placed  a  man  inside  the  Manned  Spacecraft  Center;  Bob  Gilruth  and  his 
men  would  not  have  tolerated  such  interlopers.  Other  centers  were  left  out 
in  the  cold.  Sometimes  those  who  knew  about  Langley's  operation  tried  to 
tap  Mattson  for  information.  "Once  in  a  while,"  Mattson  remembers,  "I 
would  get  a  call  from  a  guy  at  Lewis  or  Ames  about  something  he  wanted 
that  was  available  down  in  Houston."  For  such  casual  requests,  especially  if 
he  knew  the  guy,  Mattson  usually  did  his  best,  but  he  always  kept  Langley's 
interests  foremost  in  his  mind.11 

It  is  rather  amazing  that  the  fledgling  Houston  organization  put  up  with 
Mattson  for  as  long  as  it  did,  which  was  to  the  time  of  the  first  Apollo  flights 
in  1968.  Certainly  this  would  not  have  been  possible  without  the  good  graces 
of  Gilruth,  who  was  Floyd  Thompson's  good  friend.  Not  everyone  at  the 
Manned  Spacecraft  Center  wanted  Mattson  nosing  around.  In  one  meeting 
not  long  after  the  opening  of  the  Houston  center,  Gilruth  was  surprised  to 
encounter  Mattson  and  wanted  to  know  what  Mattson  was  doing  there.  Max 
Faget  answered,  "He's  doing  nothing  as  far  as  I  know."  In  Mattson's  opinion, 
Faget  did  not  know.  Mattson  had  stayed  as  far  away  from  Faget  as  possible, 
knowing  that  the  former  member  of  PARD  and  the  STG  did  not  like  him 
or  the  idea  of  his  walking  the  halls  of  the  new  Houston  space  center.  To 
the  extent  that  he  reported  to  anyone  at  Houston,  Mattson  dealt  exclusively 
with  Paul  Purser,  Gilruth's  deputy  director,  a  longtime  friend  and  Langley 
veteran.  Purser  knew  what  Mattson  was  up  to,  generally  speaking,  but  for 
a  short  period  of  time  at  the  beginning  had  failed  to  tell  his  boss  anything 
about  it.12 

Immediately  after  that  first  encounter  with  Mattson,  Gilruth  called  both 
Purser  and  Mattson  "on  the  carpet"  and  demanded  the  details  of  the  latter's 
assignment.  Mattson  explained  his  operation  and  what  Thompson  and 
Donlan  wanted  out  of  it,  and  Gilruth  grudgingly  gave  his  okay.  Whether 
Gilruth  ever  brought  up  the  matter  of  Mattson's  presence  in  Houston  with 
his  former  Langley  compatriots,  Thompson  and  Donlan,  is  unknown,  but  he 
probably  did.  Whatever  the  details  of  that  conversation  were,  the  outcome 
was  that  Mattson  stayed  at  Houston  off  and  on  for  the  next  several  years — 
despite  Max  Faget's  objections.  In  key  respects,  the  assignment  was  a 
thankless  job,  not  only  professionally  but  also  personally;  by  Mattson's 
own  admission,  it  forced  him  to  neglect  his  family  in  Virginia.13  But  such 
sacrifices  were  consistent  with  the  demands  of  the  spaceflight  revolution, 


360 


In  the  Service  of  Apollo 


L-67-1562 

Axel  Mattson  (far  left)  flashes  the  winning  smile  that  made  it  possible  for  him,  an 
outsider,  to  walk  the  halls  of  the  Manned  Spacecraft  Center  in  Houston  for  months  at 
a  time.  Robert  R.  Gilruth  (center),  the  Houston  center  director  grudgingly  indulged 
Mattson's  presence.  (This  photo  was  actually  taken  at  Langley  in  1967  at  the  foot 
of  the  Lunar  Landing  Research  Facility.)  To  the  right  of  Gilruth  is  Charles  Donlan, 
Langley 's  deputy  director;  to  his  right  is  Donald  Hewes,  the  Langley  engineer  in 
charge  of  the  Lunar  Landing  Research  Facility.  The  identity  of  the  man  in  the 
NASA  overalls  is  unknown. 


and  he  could  always  take  comfort  in  the  romantic  notion  that  he  was  an 
agent  in  Langley's  secret  service  for  Apollo. 


The  Dynamics  of  Having  an  Impact 

Only  in  one  or  two  isolated  instances  did  Mattson's  presence  in  Houston 
contribute  in  any  dramatic  way  to  Apollo's  ultimate  success.  As  discussed 
in  chapter  eight,  in  early  1962,  not  long  after  starting  his  assignment  in 
Houston,  Mattson  helped  win  support  for  John  Houbolt's  LOR  concept  in 
the  face  of  some  stubborn  opposition  at  the  Texas  center.  Mattson  took 
Houbolt  to  every  person  Mattson  thought  might  be  willing  to  listen,  and 
at  the  end  of  the  day,  at  least  in  Mattson's  opinion,  the  majority  of  the 

361 


Space/light  Revolution 

Houston  center  engineers  supported  the  LOR  concept  as  the  best  mission 
mode  for  Apollo.  A  few  months  later,  thanks  to  some  rather  keen  technical 
insight  regarding  the  kind  of  test  data  NASA  might  need  to  assure  the 
design  integrity  of  the  returning  Apollo  command  module,  Mattson  made  a 
less  significant  but  still  notable  contribution — one  that  does  not  appear  in 
the  formal  NASA  record. 

The  story  of  that  contribution  begins  three  months  after  Mattson 's 
arrival  in  Houston — that  was  when  Mattson  first  got  wind  of  an  important 
full-scale  test  being  planned  to  measure  the  impact  dynamics  of  the  Apollo 
command  module.  The  North  American  Aviation  Corporation,  which  in 
November  1961  had  won  the  contract  to  design  and  build  the  command  and 
service  modules  of  the  Apollo  spacecraft,  had  built  a  swimming-pool-like 
facility  with  a  big  gantry  next  to  it  at  the  site  of  its  Space  Information 
Systems  Division  in  Downey,  California;  the  purpose  of  this  facility  was  to 
acquire  data  on  the  pressure  and  acceleration  loads  placed  on  the  command 
module  upon  impact  with  the  ocean.  In  late  1962  and  early  1963,  North 
American  and  NASA  engineers  had  already  put  the  command  module 
through  preliminary  impact  tests  at  the  facility  and  were  planning  for  a 
final  verification  test  involving  the  highest  and  most  severe  impact  drop 
angle  and  an  Apollo  capsule  configuration  systematically  equipped  with 
pressure  transducers  to  measure  the  impact  loads.  Inside  the  capsule  the 
engineers  were  even  installing  instrumented  mannequins,  as  in  automobile 
crash  tests,  to  see  how  astronauts  would  come  through  the  jolting  splash 
into  the  water.14 

Mattson  became  interested  in  the  details  of  this  important  test.  He 
made  contacts  with  some  of  the  informed  NASA  researchers  and  contractor 
representatives  and  started  reading  up  on  it.  From  his  knowledge  of  similar 
capsule-drop  tests  carried  out  in  the  Back  River  at  Langley  on  the  Mercury 
capsule,  he  understood  that  such  a  test  would  prove  troublesome.  First,  a 
flexible  bottom  surface  on  a  space  capsule,  such  as  the  heat-shield  system 
designed  for  the  underneath  side  of  the  Apollo  command  module,  would 
cave  in  or  "oilcan"  upon  any  hard  impact.  Second,  when  transducers  were 
put  on  a  boilerplate  capsule  for  the  purpose  of  measuring  the  loads  of 
the  impact  pressure,  the  Langley  engineers  had  found  that  the  pressure 
would  just  "wash  out"  or  dissipate,  thus  making  it  impossible  for  them  to 
obtain  the  needed  data.  "There  would  be  so  much  resonance  and  what-not 
in  the  structure,"  Mattson  explains,  "that  the  damn  pressure  transducers 
wouldn't  give  you  that  instantaneous  spike  [i.e.,  the  unusually  high  and 
sharply  defined  maximum  or  pointed  element  in  the  graph].  And  that  was 
what  busted  things  up  [when  one  dropped  a  space  capsule  from  a  great 
height  into  the  water].  It  wasn't  all  that  other  trash.  It  was  that  instant 
spike."  In  the  Mercury  program  the  only  way  that  the  Langley  engineers 
could  get  that  spike  in  the  recorded  data  was  to  work  not  with  boilerplates 
but  with  solid  models.  This  solution  had  worked  perfectly.15 


362 


In  the  Service  of  Apollo 

Immediately  upon  hearing  about  the  proposed  Apollo  tests,  Mattson 
called  Sandy  M.  Stubbs,  a  bright  young  engineer  in  the  Impacting  Struc- 
tures Section  of  Langley's  Structures  Research  Division.  Stubbs  was  then 
conducting  a  test  program  on  the  water  landing  characteristics  of  various 
spacecraft  models.  Mattson  asked  Stubbs  if  he  was  getting  any  data  on  a 
solid  model  that  duplicated  the  Apollo  command  module.  Stubbs  answered 
that  he  had  no  such  data;  he  was  not  using  the  Apollo  configuration  because 
North  American's  capsule  design  was  already  fixed  and  was  not  going  to  be 
changed.  That  was  not  the  answer  Mattson  wanted  to  hear,  so  he  tried 
coaxing  Stubbs  into  adding  an  Apollo  model  to  his  test.  Stubbs  replied  that 
he  was  "running  a  little  short"  on  funds  and  was  in  charge  of  a  "low-key 
kind  of  operation."  He  would  like  to  do  it,  of  course,  as  he  was  trying  to  do 
a  systematic  study  and  was  planning  to  write  a  formal  technical  report.  If 
Mattson  could  obtain  the  funding,  Stubbs  told  Langley's  agent  in  Houston, 
then  he  would  be  glad  to  go  ahead  with  the  test.16 

Mattson  made  an  appointment  with  the  Manned  Spacecraft  Center's 
Joseph  N.  Kotanchik,  NASA's  main  technical  monitor  for  spacecraft  struc- 
tures work  on  the  North  American  contract.  Kotanchik,  like  many  of  the 
other  Houston  officials,  was  a  Langley  old-timer;  he  had  been  a  key  member 
of  the  design  team  that  had  built  Langley's  Structures  Research  Laboratory 
in  1939-1940.  Unfortunately,  he  and  Mattson  did  not  get  along  well.  Per- 
haps it  was  because  Kotanchik  was  quite  formal,  a  northerner,  and  an  MIT 
graduate,  whereas  Mattson  was  usually  informal,  a  loquacious  southerner 
by  disposition  if  not  by  birth  (he  was  born  in  New  Jersey),  who  was  edu- 
cated at  North  Carolina  State  University  in  Raleigh.  Perhaps  the  friction 
also  stemmed  from  the  professional  difficulties  Kotanchik  had  experienced 
while  at  Langley.  If  Mattson  had  not  had  previous  dealings  with  Kotanchik, 
he  would  not  have  even  bothered  to  make  an  appointment.  Mattson  usu- 
ally just  walked  into  someone's  office,  gave  a  warm  greeting,  sat  down,  and 
started  talking.  No  one  just  walked  in  on  Kotanchik  except  his  bosses. 

Kotanchik  did  not  want  to  be  bothered  with  Mattson  and  told  him  to  be 
quick.  Mattson  promptly  asked  for  a  transfer  of  $40,000  to  cover  the  costs  of 
Sandy  Stubbs's  impact  loads  test  on  a  solid  model  of  the  Apollo  command 
module  at  Langley.  Kotanchik  flatly  refused,  and  essentially  told  Mattson  to 
get  out  of  his  office.  According  to  Mattson,  Kotanchik  said,  "Don't  you  know 
that  we're  going  to  have  a  full-scale  test  that  we've  spent  a  million  dollars  on? 
Don't  you  know  that  it  will  get  us  all  the  information  we  need?"  Mattson 
tried  to  explain  about  Langley's  experience  with  the  problem  of  flexible 
bottom  structures  oilcanning  and  the  resonance  of  impacting  boilerplate 
capsules,  about  the  character  of  the  data  spike,  and  about  the  need  for 
solid-model  testing,  but  Kotanchik  would  not  listen.  "Well,  I  got  kind  of 
hot  about  that,"  Mattson  remembers.  "But  the  business  I  was  in,  you 
couldn't  stay  hot  too  long.  You  just  momentarily  got  shook  up,  but  then 
forgot  about  it  real  quick.  But  I  put  it  on  the  back  burner  and  I  made 
my  contacts  with  North  American."  In  addition,  he  told  Charles  Donlan 

363 


Spaceflight  Revolution 


L-65-3649 


L-61-8040 

One  of  Langley's  notable  behind-the-scenes  contributions  to  the  success  of  Project 
Apollo  was  its  testing  to  assure  the  landing  integrity  of  the  returning  command 
module  whether  the  capsule  splashed  down  into  water  (top)  or  maneuvered  to  a  soft 
landing  on  earth  (below). 


364 


In  the  Service  of  Apollo 

about  Kotanchik's  refusal.  Donlan,  understanding  the  good  sense  of  what 
Mattson  was  trying  to  do,  told  him  to  tell  Sandy  Stubbs  to  go  ahead  and 
include  a  solid  model  of  the  Apollo  command  module  in  his  tests,  because 
Stubbs  wanted  to  do  the  systematic  testing  anyway.  Langley  management 
would  just  have  to  find  another  way  to  pay  for  it.17 

By  the  day  of  the  big  drop  test  in  California  in  late  September  1964, 
Mattson  had  developed  such  good  rapport  with  one  of  the  North  American 
representatives  that  Mattson's  office  had  been  hooked  up  by  telephone  to 
the  North  American  office  in  Houston,  which  was  linked  to  the  test  site  in 
California,  so  he  could  hear  a  blow-by-blow,  "real-time"  narration  of  the 
entire  command  module  drop  test.  The  test  did  not  go  well.  As  Mattson 
remembers,  the  gist  of  the  telephone  narration  originating  in  Downey  was 
as  follows:  "All  right,  the  capsule  drops.  It  lands  in  the  water.  My  God, 
it's  sinking.  It  has  gone  in  and  split  wide  open.  All  the  mannequins  are 
drowning.  The  whole  spacecraft  is  in  ruins."  At  that  moment,  it  was  clear 
to  everyone  that  a  minor  catastrophe  was  at  hand;  at  that  moment,  also, 
Mattson's  telephone  hookup  was  cut  off.18 

Mattson  knew  he  had  "a  hot  one."  Raw  technical  data  existed  at  Langley, 
or  soon  would,  that  might  be  the  key  to  solving  the  problem;  but  what 
exactly  was  he  to  do  with  this  knowledge,  and  when?  He  decided  it  was 
best  to  bide  his  time  and  see  what  developed,  because  before  long  something 
would  surely  "hit  the  fan."  In  a  few  days,  NASA  and  North  American  put 
together  an  Impact  Test  Program  Review  to  look  into  the  splashdown  of 
the  wrecked  command  module.  "You  wouldn't  believe  the  collection  of 
structural  experts  and  spacecraft  experts  that  all  convened  at  Houston." 
Into  these  meetings  sneaked  Axel  Mattson.  He  sat  in  the  back  trying  to  be 
inconspicuous  but  listening  carefully  to  all  that  was  said.  By  the  third  day 
of  these  meetings,  chaired  by  Kotanchik,  the  engineers  involved  had  come 
up  with  a  plan  of  attack  for  solving  the  Apollo  spacecraft's  splashdown 
structures  problem;  the  plan  was  basically  to  retest  after  strengthening  the 
overall  structure  of  the  command  module  with  special  attention  to  increasing 
the  thickness  and  stiffness  of  the  heat-shield  structure.  The  only  way  to 
strengthen  the  structure  was  to  add  material  and  move  toward  a  solid-model 
capsule  test  program  similar  to  the  one  that  had  been  conducted  at  Langley 
for  Mercury.  At  this  point,  very  late  in  the  course  of  the  meetings,  Mattson 
began  to  "lick  his  chops,"  because  the  questions  then  became:  "Who  has  got 
the  impact  loads?  Who  knows  what  they  are?"  These  data  were  apparently 
not  available,  but  Mattson — and  Mattson  alone — knew  how  to  get  them.19 

Joe  Kotanchik  knew  Mattson  knew.  Prom  his  seat  in  the  front  of  the 
auditorium,  the  chairman  of  the  meeting  nervously  looked  around  the  room 
until  he  spotted  the  one  man  in  the  back  with  a  wry  smile  on  his  face;  then, 
Kotanchik  called  for  a  coffee  break.  He  immediately  got  up  and  pointed  his 
finger  at  Mattson  and  motioned  the  Langley  interloper  to  follow  him  into 
his  office.  Inside  he  asked  Mattson  if  he  had  heard  what  it  was  that  was 
needed  to  move  on  with  the  spacecraft  verification  program.  Mattson,  with 

365 


Space/light  Revolution 

more  than  a  little  personal  satisfaction,  answered  that  he  had  indeed  heard 
and  that  it  was  "exactly  what  I  was  talking  about  in  here  when  I  came  in 
two  months  ago,  Joe.  I  was  trying  to  get  that  damn  spike.  They  don't  know 
what  that  spike  is.  They  don't  know  the  magnitude  of  the  impact  loads 
that  busts  things  up."  Kotanchik,  however,  close  to  panic,  was  in  no  mood 
for  such  a  speech  and  told  Mattson  to  shut  up.  Mattson  did,  his  victory 
sweet  enough  without  rubbing  it  in.  He  told  Kotanchik  that  he  would  have 
Sandy  Stubbs  on  an  airplane  and  in  Houston  by  tomorrow  morning — which 
he  did.  "And  that's  how  NASA  got  the  information  it  needed."  Stubbs 
did  a  terrific  job  presenting  his  research  findings  to  Kotanchik's  assembled 
experts:  he  discussed  the  pertinent  parameters  of  his  1/4-scale  model  of  the 
Apollo  command  module,  presented  the  acceleration  and  pressure  data  for 
the  capsule's  landings  on  water,  and  explained  a  slide  with  a  cross  section 
of  the  solid  model  showing  its  construction  details.  In  the  ensuing  years, 
Stubbs  went  on  to  make  a  major  contribution  to  understanding  the  impact 
dynamics  that  would  affect  the  Apollo  command  module  upon  splashdown.20 
Mattson  never  wrote  a  report  on  the  affair,  but  he,  too,  had  made  a 
contribution.  It  was  the  sort  of  undercover  technical  work  that  did  not 
make  its  way  into  the  history  books,  but  nothing  in  his  service  for  Apollo, 
or  for  Langley,  ever  gave  Mattson  greater  pride  or  satisfaction. 

Inside  the  Numbers 

Mattson  did  write  other  reports.  At  least  once  a  year,  he  put  together  a 
detailed  inventory,  "Langley  Research  Center  Tests  of  Interest  to  Project 
Apollo,"  for  Bob  Gilruth  at  the  Manned  Spacecraft  Center  and  Floyd 
Thompson  at  Langley.  The  document  described  briefly  the  tests  being  done 
by  Langley  researchers  in  support  of  Apollo,  identified  the  research  divisions 
and  facilities  in  which  the  work  was  being  done  and  the  project  engineers 
conducting  it,  gave  the  scheduled  test  dates,  and  offered  some  associated 
remarks.  The  purpose  of  the  report  was  dual:  to  demonstrate  how  much 
Langley  was  doing  to  support  the  lunar  landing  program  and  to  show  the 
Manned  Spacecraft  Center  the  wide  range  of  Langley  research  that  could 
perhaps  be  useful  to  Apollo.  Mattson  distributed  copies  of  this  unpublished 
typescript  to  four  or  five  of  the  most  important  Apollo  program  managers 
in  Houston,  including  the  manager  of  the  Apollo  Spacecraft  Program  Office 
(in  1963,  Joseph  F.  Shea).21 

Although  it  was  self-serving,  Mattson 's  regular  inventory  nevertheless 
provided  a  rather  accurate  survey  of  everything  being  done  at  Langley 
that  was  in  any  way  related  to  Apollo.  Every  item  listed  was  legitimate. 
Nothing  was  invented,  although  the  breadth  or  depth  of  certain  studies  was 
sometimes  stretched  a  bit  to  make  them  seem  more  applicable  to  Apollo 
than  they  really  were.  Still,  these  informal  reports  provide  an  unusually 
complete  record  of  how  much  Langley  did  in  the  service  of  Apollo,  at  least 
as  reported  by  Mattson  between  December  1962  and  February  1968. 

366 


In  the  Service  of  Apollo 


L-62-9222 

Project  Fire  explored  the  effects  of  reentry  heating  on  Apollo  spacecraft  materials. 
Although  the  ultimate  tests  involved  Atlas  rockets  carrying  recoverable  reentry 
packages,  the  flight  tests  from  Cape  Canaveral  were  preceded  by  a  series  of  important 
wind-tunnel  tests  at  Langley,  as  shown  in  this  photo  from  1962. 

Langley's  work  on  Apollo  grew  steadily  from  1962  to  1968,  with  an 
apparent  peak  in  the  period  covered  by  Mattson's  March  1966  report.  Of  the 
three  major  research  groups  in  Langley's  organization  after  the  laboratory's 
formal  reorganization  in  late  1964,  Group  1,  under  the  direction  of  Francis 
B.  Smith,  did  the  most  work  for  Apollo — 133  projects  (85  of  them  discrete) 
compared  with  92  (46  discrete)  for  Group  2,  headed  by  John  E.  Duberg, 
and  98  (51  discrete)  for  Group  3,  headed  by  Laurence  K.  Loftin,  Jr. 

Besides  the  three  research  directorates,  two  other  major  organizations 
at  the  center  contributed  to  Apollo.  One  of  them,  the  Office  for  Flight 
Projects,  led  by  Eugene  C.  Draley,  actively  supported  Apollo;  within  the 
office,  the  Applied  Materials  and  Physics  Division  reported  43  Apollo-related 
projects.  This  organization,  formerly  PARD,  was  particularly  busy  with 
work  for  Project  Fire,  an  important  NASA  effort  to  study  the  effects  of 
reentry  heating  on  a  spacecraft  returning  to  earth  at  a  high  speed.  Fire, 
for  which  Langley  created  a  special  program  office  in  May  1961  (under  the 
direction  of  Herbert  A.  "Hack"  Wilson),  not  only  consisted  of  flight  tests 
from  Cape  Canaveral  involving  Atlas  rockets  carrying  recoverable  reentry 
packages  but  also  involved  a  considerable  amount  of  Langley  wind-tunnel 
testing.  Two  Fire  missions  eventually  took  place.  The  first,  launched  on 
14  April  1964  from  Wallops  Island,  fired  a  payload  into  the  atmosphere 

367 


Spaceflight  Revolution 


Number  of  Langley  Research  Projects 
Directly  Related  to  Apollo  Program,  1962-1968 


Dec. 

Dec. 

Nov. 

Nov. 

Mar. 

Feb. 

Directorate 

1962 

1963 

1964 

1965 

1966 

1968 

Group  1 

ACD 

0 

0 

0 

1 

1 

0 

FID 

X 

X 

X 

X 

X 

8 

IRD 

1 

6 

4 

9 

9 

1 

SMD 

4 

9 

19 

19 

19 

23 

Total 

5 

15 

23 

29 

29 

32 

Group  2 

OLD 

7 

7 

9 

12 

13 

8 

SRD 

2 

5 

9 

7 

8 

5 

Total 

9 

12 

18 

19 

21 

13 

Group  3 

APD 

6 

5 

9 

5 

5 

4 

FMTD 

10 

5 

2 

2 

3 

1 

FSRD 

9 

10 

6 

6 

6 

4 

Total 

25 

20 

17 

13 

14 

9 

Flight  Projects 

LOPO 

X 

0 

0 

1 

1 

4 

MORL 

X 

0 

0 

2 

2 

0 

AMPD 

0 

2 

5 

10 

11 

15 

Total 

0 

2 

5 

13 

14 

19 

Engineering  and 

Technical  Services 

FVSD 

X 

X 

X 

X 

X 

1 

ESD 

0 

0 

0 

0 

0 

0 

MSD 

0 

0 

0 

0 

0 

0 

PMD 

0 

0 

0 

0 

0 

0 

RMFD 

0 

0 

0 

0 

0 

0 

Total 

0 

0 

0 

0 

0 

1 

x                division  did  not  exist                                       FSRD        Full-Scale  Research  Div. 

ACD          Analysis  and  Computation  Div.                     LOPO       Lunar  Orbiter  Project  Office 

FID            Flight  Instrumentation  Div.                             MORL       Manned  Orbiting  Research  Laboratory 

IRD           Instrument  Research  Div.                               AMPD      Applied  Materials  and  Physics  Div. 

SMD          Space  Mechanics  Div.                                      FVSD        Flight  Vehicles  and  Systems  Div. 

OLD          Dynamic  Loads  Div.                                        ESD           Electrical  Systems  Div. 

SRD           Structures  Research  Div.                                MSD          Mechanical  Services  Div. 

APD          Aero-Physics  Div.                                              PMD          Plant  Maintenance  Div. 

FMTD       Flight  Mechanics  and  Technology  Div.          RMFD       Research  Models  and  Facilities  Div. 

368 


In  the  Service  of  Apollo 

at  a  speed  in  excess  of  40,000  kilometers  an  hour,  the  velocity  that  the 
Apollo  spacecraft  returning  from  the  moon  was  expected  to  reach.  The 
second,  on  22  May  1965,  basically  confirmed  the  findings  of  the  first 
mission,  which  indicated  that  "the  radiation  and  the  temperatures  that 
would  be  experienced  by  an  Apollo  spacecraft  reentry  were  less  severe  than 
expected."22  Spacecraft  engineers  at  Houston  and  North  American  made 
use  of  this  important  data  from  Project  Fire  in  designing  and  qualifying 
Apollo's  heat  shield. 

Even  Langley's  Office  of  Engineering  and  Technical  Services,  Langley's 
fifth  directorate,  which  normally  was  not  involved  in  much  research,  took  on 
a  special  Apollo  project  by  conducting  a  study  of  an  electrolytic  chlorinator 
for  water  sterilization. 


The  Simulators 

The  most  active  of  all  Langley  divisions  in  Apollo  work  was  the  Space 
Mechanics  Division  (SMD)  of  Group  1.  This  division  conducted  more  studies 
related  to  the  lunar  landing  mission  than  did  any  other  division  at  Langley, 
perhaps  because  it  was  in  this  division,  previously  known  as  the  Theoretical 
Mechanics  Division,  that  LOR  had  germinated. 

The  essence  of  SMD's  contributions  to  Apollo  rested  in  its  simulation 
research.  This  was  a  field  of  work  that  dated,  at  the  center,  to  the  early  1940s 
when  manned  simulation  for  aeronautical  R&D  began  because  of  the  need 
for  World  War  II  pilot  trainers.  In  the  following  15  years,  with  significant 
advances  in  servomechanism  and  control  theory  and,  more  importantly,  in 
analog  and  digital  computers,  simulation  technology  made  a  quantum  leap 
forward.  A  new  generation  of  intricate  and  capable  machines  was  developed 
just  when  such  simulators  were  needed  for  spaceflight.  As  Langley's  foremost 
simulation  expert,  Arthur  W.  Vogeley,  head  of  SMD's  Guidance  and  Control 
Branch,  said  in  a  speech  to  the  American  Society  of  Mechanical  Engineers 
in  1966,  "Simulation  is  now  big  business.  In  total  investment  of  professional 
manpower  and  facilities  it  is  larger  now  than  the  whole  aviation  industry 
was  not  many  years  ago.  Simulation  is  growing  rapidly — exponentially,  it 
seems.  Where  it  will  go  in  the  next  15  years  is  anybody's  guess."23 

The  Sputnik  crisis  had  made  simulation  research  and  astronaut  training 
absolutely  vital — in  large  part  because  human  ambitions  began  to  outstrip 
human  understanding;  simulators  were  needed  to  fill  in  the  gaps  in  the 
basic  knowledge  about  spaceflight.  Building  simulators  to  investigate  the 
interface  between  the  airplane  and  the  pilot  had  been  a  difficult  challenge 
for  aeronautical  researchers  prior  to  the  spaceflight  revolution;  building 
machines  that  simulated  the  interface  between  astronaut  and  spacecraft  in 
the  weightless  environment  of  outer  space  was  an  even  tougher  job — but 
one  that  had  to  be  done.  In  Projects  Gemini  and  Apollo,  astronauts  and 
spacecraft  were  to  be  committed  to  major,  complex,  and  untried  maneuvers 

369 


Spaceflight  Revolution 

which,  of  necessity,  had  to  be  carried  through  to  a  successful  completion, 
and  usually  on  the  first  attempt.  Simulation  was  vital  to  success. 

Very  little  simulation  was  necessary  for  Project  Mercury,  but  Project 
Gemini  entailed  orbital  rendezvous  and  docking,  which  was  a  more  danger- 
ous and  complex  maneuver  than  sending  a  capsule  into  orbit.  Learning  how 
to  link  two  spacecraft  or  spacecraft  modules  in  orbit — an  operation  required 
by  Apollo's  LOR  mode — ultimately  became  the  primary  purpose  of  Project 
Gemini,  NASA's  second  man-in-space  program.  Many  people  fail  to  appre- 
ciate the  basic  purpose  of  Project  Gemini,  which  was  to  serve  as  a  bridge 
between  Mercury  and  Apollo  and  to  develop  the  techniques  of  rendezvous 
and  docking,  spacewalking,  and  long-duration  flights  required  by  the  Apollo 
lunar  landing  mission.24 

Docking  itself  was  a  straightforward  operation,  very  much  like  the  mid- 
air refueling  of  a  jet  airplane,  which  was  a  maneuver  that  experienced  pilots 
routinely  managed  just  by  flying  "all  eyeballs"  and  by  the  seat-of-the-pants. 
Rendezvous,  however,  seemed  to  be  an  altogether  different  matter.  Michael 
Collins  had  more  than  his  share  of  experience  with  the  "dark  mysteries"  of 
rendezvous  during  his  Gemini  X  and  Apollo  11  flights:  "Sir  Isaac  Newton, 
when  formulating  the  laws  of  gravity  and  motion,  had  no  idea  how  difficult 
he  was  making  it  for  those  of  us  who  would  fly  his  circles  and  ellipses.  It 
was  simple  enough  to  explain,  with  a  chalkboard  or,  better  yet,  a  powerful 
digital  computer,  but  in  flight  one  had  to  be  extraordinarily  careful  not 
to  make  a  false  move,  not  to  trust  the  eyes  alone,  not  to  fire  the  engines 
unless  each  maneuver  had  been  checked  and  double-checked."  Collins,  one 
of  the  most  thoughtful  astronauts  (and  best  writers)  to  comment  on  his 
experiences  in  space,  captures  what  it  was  like  for  a  Gemini  pilot  to  try  to 
catch  and  rendezvous  with  his  Agena  target/docking  vehicle  some  2  miles 
ahead  of  him: 

The  pilot  sees  the  Agena's  twinkling  light  out  the  window,  points  the  nose  of  the 
Gemini  at  it,  and  fires  a  thruster  to  move  toward  the  Agena.  For  a  short  time 
all  seems  well,  and  the  Agena  grows  in  size.  Then  a  strange  thing  happens:  the 
Agena  begins  to  sink  and  disappears  under  the  Gemini's  nose.  Then  minutes  later 
it  reappears  from  below,  but  now  it  is  going  faster  than  the  Gemini  and  vanishes 
out  front  somewhere.  What  has  happened?  When  the  Gemini  fired  its  thruster  it 
increased  its  velocity  but  also  its  centrifugal  force,  causing  its  orbit  to  become  larger. 
As  it  climbed  toward  its  new  apogee,  it  slowed  down,  so  that  it  began  to  lose  ground 
compared  to  the  Agena.  The  Gemini  pilot  should  have  fired  a  thruster  to  move  away 
from  the  Agena,  causing  him  to  drop  down  below  it  into  a  faster  orbit,  and  begin  to 
overtake  it.  Then,  when  the  Agena  reached  a  precisely  calculated  angle  above  him, 
he  could  thrust  toward  it  and  his  resulting  orbit  would  intercept  that  of  the  Agena. 
Sir  Isaac  demands  that  you  play  his  game  his  way. 

Rendezvous  in  space  could  turn  sour  with  paralyzing  swiftness.  An  on-board 
computer  might  fail,  a  gyroscope  might  tilt  the  wrong  way,  or  some  other 
glitch  might  occur  to  complicate  the  performance  of  a  necessary  maneuver. 

370 


In  the  Service  of  Apollo 

Pilots  of  both  the  LEM  and  the  CM  had  to  be  ready  to  make  crucial  decisions 
instantaneously.  They  could  not  simply  say  to  one  another  "Meet  Me  Over 
St.  Louis"  and  expect  their  two  spacecraft  to  rendezvous  successfully  in 
space.  The  usual  way  of  piloting  in  atmospheric  flight  just  would  not  work. 
Because  of  these  stark  new  realities  about  flight  in  space,  Collins  notes,  "a 
great  amount  of  care  and  pre-planning  had  to  go  into  the  planning  of,  and 
hardware  for,  the  rendezvous  missions."25 

Engineers  in  Langley's  Theoretical  Mechanics  Division  (which  was  ac- 
tually renamed  the  Aero-Space  Mechanics  Division  before  becoming  SMD) 
had  become  heavily  involved  in  certain  aspects  of  spaceflight  simulation 
even  before  Sputnik.  Their  work  for  the  X-15  program  (when  they  were 
still  NACA  employees)  had  led  them  to  construct  and  "fly"  an  X-15  atti- 
tude and  control  simulator.  But  Sputnik  set  them  loose;  over  the  course 
of  the  next  10  years  they  conceived,  built,  and  operated  nine  new  simu- 
lators. These  included  a  Rendezvous  and  Docking  Simulator,  a  Rotating 
Vehicle  Simulator  to  study  the  effects  on  astronauts  of  long  stays  in  a  ro- 
tating space  station,  and  a  Reduced  Gravity  Walking  Simulator  that  was 
used  to  evaluate  the  effects  of  lunar  gravity  on  man's  walking  and  running 
capabilities  (with  and  without  pressure  suits).  A  Lunar  Orbit  and  Let- 
down Approach  Simulator  (LOLA),  a  $1.9-million  facility,  was  designed  so 
that  pilots  could  experience  the  same  sort  of  visual  cues  that  they  would 
encounter  while  navigating  and  controlling  a  spacecraft  in  the  vicinity  of 
the  moon.  The  Lunar  Landing  Research  Facility,  a  mammoth  $3.5-million 
facility,  simulated  manned  lunar  landings,  and  a  Projection  Planetarium 
was  built  that  could  either  project  stars  on  a  plastic  dome  while  test  pilots 
sat  on  a  rotating  gun  turret  "spacecraft"  to  get  the  feel  of  heavenly  move- 
ment or  generate  a  horizon-to-horizon  view  of  Florida  as  seen  from  about 
100,000  feet  for  "out-the-window"  studies  of  Apollo  launch-abort  problems. 
The  Virtual  Image  Rendezvous  Simulator  used  a  closed-circuit  television 
system  and  analog  computers  to  represent  a  moving  target  vehicle  for  ren- 
dezvous and  docking  studies;  a  Water  Immersion  Simulator  used  a  water 
tank  for  investigating  problems  associated  with  manned  extra-  and  intra- 
vehicular  activities  in  reduced-G  environments;  and  a  One-Man  Propulsion 
Research  Apparatus  suspended  a  person  equipped  with  vertical  thrusters 
and  translation  and  attitude  controllers  from  a  lightweight  gimbal  unit 
for  maneuverability  studies  in  reduced-gravity  fields.26  Two  of  the  space- 
flight  simulators  in  particular — the  Rendezvous  and  Docking  Simulator  and 
the  Lunar  Landing  Research  Facility — made  significant  contributions  to  the 
successes  of  projects  Gemini  and  Apollo. 

In  the  months  following  NASA's  adoption  of  the  LOR  concept,  a 
team  of  Langley  engineers  led  by  Arthur  W.  Vogeley  and  Max  C.  Kurbjun 
of  the  Space  Mechanics  Division,  and  including  Roy  R.  Brissendon, 
Alfred  J.  Meintel,  Jr.,  Jack  E.  Pennington,  and  Marvin  C.  Waller,  designed 
an  unusual  research  facility  explicitly  for  the  purpose  of  studying  the  special 
problems  of  rendezvous  and  docking.  In  its  essentials,  the  design  belonged 

371 


Spaceflight  Revolution 


L-63-5016 


L-64-1610  L-62-9652 

The  Langley  Rendezvous  and  Docking  Simulator  suspended  from  the  roof  of  the  West 
Area  airplane  hangar  (top).  Bottom  left,  a  time-lapse  look  at  a  successful  docking. 
Bottom  right,  an  unidentified  pilot  "eyeballs  "  his  way  to  a  docking  by  peering  through 
the  portal  in  his  capsule. 


372 


In  the  Service  of  Apollo 

to  Vogeley,  the  head  of  SMD's  Guidance  and  Control  Branch,  and  entailed 
full-scale  mock-ups  of  the  Gemini  and  Apollo  cockpits  that  could  be  hung 
from  an  overhead  carriage  and  cable-suspended  gimbal  system.  This  entire 
assembly  could  then  be  attached  to  an  overhead  moving  crane  that  moved 
along  a  210-foot  track  running  along  the  rafters  of  Langley's  cavernous  West 
Area  hangar.  Pilot  astronauts  could  then  "fly"  the  cockpits  both  in  night- 
time and  daylight  conditions  to  rehearse  and  perfect  rendezvous  and  docking 
skills. 

Upon  its  completion  in  early  1963,  SMD  began  using  this  ingenious 
simulator  to  study  the  finer  points  of  various  rendezvous  missions.  (The 
simulator  included  a  general-purpose  analog  computer,  which  made  it 
possible  for  the  pilot  inside  the  gimbal  to  experience  all  six  degrees  of 
motion  freedom.)  Experience  with  the  facility  confirmed  something  that 
Vogeley  and  other  experienced  guidance  and  control  experts  at  Langley  had 
believed  for  some  time:  rendezvous,  if  practiced,  was  not  as  mysterious  or 
as  difficult  as  many  people  imagined.  In  fact,  it  could  be  accomplished  quite 
easily.27 

Nonetheless,  without  their  experience  in  the  Rendezvous  and  Docking 
Simulator,  the  Gemini  and  Apollo  astronauts  would  not  have  been  as  well 
prepared  for  handling  the  pressures  of  rendezvous.  "We  trained  an  awful 
lot  of  astronauts,"  Vogeley  remembers  with  pride,  "who  all  appreciated  the 
realism  of  the  simulator's  visual  scene.  It  gave  us  a  lot  of  satisfaction  just 
to  show  that  NASA  could  do  that  sort  of  thing  in  a  unique  piece  of  ground 
equipment  that  only  cost  about  $300,000.  I  think  we  got  our  money's 
worth."28  In  his  branch  office  Vogeley  kept  a  picture  of  the  simulator 
that  was  signed  by  all  the  pilots  who  ever  used  the  facility — with  one 
exception.  Apollo  astronaut  Ed  White,  who  visited  Langley  twice  to  fly 
on  the  simulator,  died  in  the  tragic  launchpad  fire  at  Cape  Kennedy  on 
27  January  1967  before  he  could  add  his  autograph  to  the  list. 

The  other  simulator  to  contribute  in  a  significant  way  to  the  success 
of  Apollo  was  the  Lunar  Landing  Research  Facility,  an  imposing  250-foot- 
high,  400-foot-long  gantry  structure  that  became  operational  in  1965  at  a 
cost  of  nearly  $4  million.  Conceived  in  1962  by  engineer  Donald  Hewes  and 
built  under  the  careful  direction  of  his  quiet  but  ingenious  division  chief, 
W.  Hewitt  Phillips,  this  gigantic  facility  was  designed  to  develop  techniques 
for  landing  the  rocket-powered  LEM  on  the  moon's  surface.  Because  the 
moon  had  no  atmosphere  and  its  gravitational  pull  was  only  one-sixth  that 
of  the  earth's,  piloting  the  LEM  would  be  completely  unlike  atmospheric 
flying.  The  thrust  of  the  LEM's  rockets  in  a  vacuum  would  produce  unusual 
and  abrupt  up-and-down,  side-to-side,  or  rolling  motions.  In  addition,  the 
lack  of  an  atmosphere  would  create  a  harsh  light.  As  the  astronauts  landed, 
they  would  face  this  bright  glare,  as  well  as  deep,  dark  shadows,  which  would 
skew  depth  perception.  Some  unique  problems  had  to  be  overcome  to  make 
a  pinpoint  lunar  landing.  Some  means  of  simulation  seemed  called  for;  the 
question  was  how  to  do  it. 

373 


Space/light  Revolution 


L-69-7140  L-63-3268 

The  two  men  responsible  for  the  design  and  early  operation  of  the  Lunar  Landing 
Research  Facility  were  Donald  Hewes  (left)  and  his  division  chief,  William  Hewitt 
Phillips  (right). 


Hewitt  Phillips,  a  soft-spoken,  MIT-educated  engineer  born  in  Port 
Sunlight,  England,  remembers  how  the  idea  for  the  Lunar  Landing  Research 
Facility  originated  between  1962  and  1963:  "Since  we  knew  that  the 
moon's  gravity  is  one-sixth  that  of  the  Earth's,  we  needed  to  support 
five-sixths  of  the  vehicle's  weight  to  simulate  the  actual  conditions  on  the 
moon."29  Perhaps,  some  practical  method  could  be  devised  to  lower  the 
apparent  weight  of  a  mock-up  LEM  to  its  lunar  equivalent  by  a  method  of 
suspension  using  vertical  cables  attached  to  a  traveling  bridge  crane. 

From  this  basic  notion,  the  design  evolved.  A  huge  gantry  structure  was 
built  that  would  dominate  Langley's  landscape  for  years  to  come.  Phillips 
and  Hewes  wanted  the  supporting  gantry  to  be  even  taller,  but  because 
of  the  heavy  military  air  traffic  from  adjacent  Langley  AFB,  the  structure 
was  limited  to  200  feet.  The  completed  facility,  however,  stood  240  feet 
6  inches,  excluding  the  top  warning  lights  and  antennae.  Two  long  cables 
provided  the  desired  vertical  lifting  force  equal  to  five-sixths  of  the  vehicle's 
weight,  thereby  opposing  the  pull  of  the  earth's  gravity  and  simulating  the 
low  gravitational  force  of  the  moon's  surface.  The  cables  were  attached  to 
a  servocontrolled  hoist  system  in  a  dolly  unit  mounted  under  a  traveling 
bridge;  the  hoist  system  was  controlled  automatically  by  load  cells  in  each 
support  strut.  As  the  test  vehicle  moved  up  and  down  and  back  and  forth  in 
response  to  the  controlling  pilot,  the  bridge  and  dolly  responded  to  signals 
from  the  vehicle  and  from  cable  angle  sensors  at  the  top  of  the  cables  to 


374 


In  the  Service  of  Apollo 


L-69-6324 

Langley's  Lunar  Landing  Research  Facility,  completed  in  1965,  helped  to  prepare  the 
Apollo  astronauts  for  the  final  150  feet  of  their  lunar  landing  mission  by  simulating 
both  the  lunar  gravity  environment  and  the  full-scale  LEM  vehicle  dynamics. 

stay  directly  over  the  vehicle  at  all  times  and  to  keep  the  cables  vertical. 
Because  the  bridge  and  dolly  system  were  driven  hydraulically,  they  provided 
a  responsive  servocontrol  system.  Moreover,  safety  features  were  built  into 
the  system  to  prevent  the  lunar  landing  vehicle  from  crashing  or  the  bridge 
and  the  dolly  from  overrunning  their  tracks  in  the  event  of  an  equipment 
malfunction  or  the  pilot  exceeding  the  safety  limits  of  the  system. 

The  lunar  landing  test  vehicle  itself  could  be  flown  up  to  about  17  miles 
per  hour  within  the  confines  of  the  overhead  structure,  which  provided  a 
travel  range  400  feet  long,  50  feet  wide,  and  180  feet  high.  The  vehicle  could 
also  be  hoisted  to  the  overhead  platform,  where  two  cables  connected  to 
the  trolley  units  on  the  lower  horizontal  truss  structure  could  catapult  the 
vehicle  downward  at  35  miles  per  hour.  To  make  the  simulated  landings 
more  authentic,  Hewes  and  his  men  filled  the  base  of  the  huge  eight-legged, 
red-and-white  structure  with  dirt  and  modeled  it  to  resemble  the  moon's 
surface.  They  erected  floodlights  at  the  proper  angles  to  simulate  lunar 
light  and  installed  a  black  screen  at  the  far  end  of  the  gantry  to  mimic  the 
airless  lunar  "sky."  Hewes  personally  climbed  into  the  fake  craters  with 
cans  of  everyday  black  enamel  to  spray  them  so  that  the  astronauts  could 
experience  the  shadows  that  they  would  see  during  the  actual  moon  landing. 

375 


Space/light  Revolution 


L-68-547 


L-67-9108 

Langley  engineers  designed  the  control  cab  of  the  Lunar  Landing  Research  Facility 's 
original  landing  module  (top  left)  from  the  cockpit  of  a  Bell  helicopter.  To  make  it 
similar  to  the  actual  LEM,  they  eventually  redesigned  it  with  a  stand-up  cab  (top 
right  and  bottom). 


376 


In  the  Service  of  Apollo 


L-69-4361  L-67-3857 

With  floodlights  shining  down  to  simulate  lunar  light  and  the  base  modeled  to 
resemble  the  lunar  surface,  24  astronauts  practiced  landings  at  the  Lunar  Landing 
Research  Facility  between  1965  and  1969. 


As  a  final  touch  to  the  facility,  Hewes  attached  to  an  overhead,  lightweight 
trolley  track  a  simple  contrivance,  which  came  to  be  known  as  the  Reduced 
Gravity  Walking  Simulator.  Made  of  canvas  slings,  steel  cables,  a  small 
trolley,  and  a  wooden  walking  surface,  this  rig  tilted  a  walker  some  80  de- 
grees from  vertical  by  holding  him  up  with  two  cables.  Astronauts,  thus 
suspended,  could  practice  moonwalking  down  the  plywood  surface.  They 
made  12-foot  jumps  with  ease,  rapidly  climbed  a  "vertical"  pole  with  one 
hand,  and  generally  got  a  feel  for  what  it  would  be  like  to  traverse  the  lunar 
surface.30  The  Reduced  Gravity  Walking  Simulator  became  quite  a  hit  with 
all  press  members  who  visited  Langley  during  the  Apollo  era.  In  1968,  for 
example,  CBS  anchorman  Walter  Cronkite  suited  up  in  an  orange  astronaut 
outfit  for  what  turned  out  to  be  a  rather  comical  televised  walk  on  the  moon. 

Of  course,  the  landing  facility  was  a  complicated  system  and  had  kinks 
that  had  to  be  ironed  out.  The  electrohydraulic  system  that  kept  the  crane 
platform  directly  over  the  flight  vehicle  and  the  cables  vertical  was  extremely 
involved.  The  facility  had  a  wonderful  assortment  of  structural,  cable- 
stretch,  and  pendulous  frequencies  that  were  unpredictable  and  required 
innovative  compensation  systems.  The  LEM  model  was  attached  to  the 
cables  through  gimbal  rings  allowing  pitch,  roll,  and  yaw  motions  produced 
by  a  hydrogen  peroxide  rocket  attitude  control  system.  The  one-sixth  weight 
not  lifted  by  the  cable  system  had  to  be  lifted  by  throttleable  hydrogen 

377 


Spaceflight  Revolution 


L-68-8308 

The  spaceflight  revolution  captivated  many  in  the  news  media,  including  then-CBS 
news  anchorman  Walter  Cronkite.  During  a  1968  visit  to  Langley,  the  adventurous 
Cronkite  tried  out  the  Reduced  Gravity  Walking  Simulator — a  series  of  cable- 
supported  slings  hanging  down  from  the  Lunar  Landing  Research  Facility  designed 
to  approximate  lunar  locomotion. 


peroxide  thrusters  fixed  to  the  vehicle  structure.  With  such  marvelous 
complexities,  it  is  no  wonder  that  it  took  some  time  for  the  Langley  engineers 
to  perfect  their  gargantuan  but  sensitive  and  responsive  mechanism. 

As  Neil  Armstrong  testified,  once  the  kinks  were  ironed  out,  the  Lunar 
Landing  Research  Facility  was  "an  engineer's  delight."  The  flying  volume 
was  "limiting,  but  adequate  to  give  pilots  a  substantive  introduction  to 
Lunar  flight  characteristics."  Moreover,  thanks  to  the  built-in  safety  fea- 
tures, whereby  the  cable  system  could  be  locked  if  the  vehicle  was  out  of 
control,  the  astronauts  were  able  to  "investigate  unorthodox  attitude,  tra- 
jectory and  control  combinations  which  would  be  impractical  in  a  free-flying 
simulator."31 

Armstrong  knew  the  limitations  of  other  simulators  from  personal  expe- 
rience. On  6  May  1968,  during  a  test  flight  of  a  free-flying  Lunar  Landing 
Training  Vehicle  at  NASA's  Flight  Research  Center  at  Edwards  AFB  in 
California,  he  was  almost  killed.  Historian  Richard  Hallion,  author  of  the 

378 


In  the  Service  of  Apollo 

definitive  book  on  the  history  of  NASA's  Flight  Research  Center,  relates  the 
accident: 

While  hovering  10  meters  above  the  ground,  the  vehicle  suffered  a  loss  of  helium 
pressure  in  its  propellant  tanks,  causing  shutdown  of  its  attitude  control  rockets.  It 
started  nosing  up  and  rolling  over,  and  Armstrong  immediately  ejected.  His  zero- 
zero  seat  kicked  him  away  from  the  stricken  craft,  which  tumbled  into  the  ground 
and  exploded  as  the  astronaut  safely  descended  by  parachute. 

"It  was  a  sad  fate  for  a  pioneering  flight  craft,"  writes  Hallion,  a  great 
lover  of  flying  machines.  Indeed,  but  it  almost  sealed  a  far  worse  fate  for 
America's  first  man  on  the  moon.32  Prom  that  day  on,  Neil  Armstrong 
made  no  flights  in  the  Lunar  Landing  Training  Vehicle  or  any  other  free- 
flying  test  vehicle  simulating  lunar  landings.  He  did,  however,  continue  to 
use  Langley's  facility  to  practice  landings. 

Ironically,  in  the  early  days  of  the  facility's  development,  many  people 
associated  with  the  Apollo  program  did  not  see  the  need  for  such  a  facility 
and  questioned  whether  it  would  ever  work.  How  could  a  vehicle  hanging 
from  cables,  like  a  child's  top  jumping  at  the  end  of  a  string,  adequately 
simulate  moon  landings?  Axel  Mattson  recalls  that  Gilruth's  engineers 
in  Houston  never  expressed  much  enthusiasm  for  the  device.  They  felt 
that  the  best  simulations  would  be  provided  by  helicopters  approximating 
final  descent  trajectories  (all  Apollo  crew  members  did  in  fact  become 
qualified  helicopter  pilots)  by  a  test  program  involving  a  modified  Bell 
X-14A  VTOL  aircraft,  or  by  the  special  free-flight  lunar  landing  research 
vehicles  being  developed  by  NASA  for  the  test  flights  at  Edwards.  All  these 
research  vehicles  made  significant  contributions  to  developing  techniques 
for  the  lunar  landing.  But  Langley's  controversial  Lunar  Landing  Facility 
provided  astronauts  with  a  unique  experience.  So  realistic  was  its  imitation 
moonscape,  for  example,  that  Neil  Armstrong  remarked  that  when  he  saw 
his  shadow  fall  upon  the  lunar  dust,  it  was  exactly  as  he  had  seen  it  at  the 
landing  facility  at  Langley.33 

Some  of  Langley's  other  simulators  did  not  make  significant  contributions 
to  Apollo — or  to  any  other  program.  The  clearest  case  in  point  was  the 
intricate  LOLA,  which  started  operating  in  1965  at  an  imposing  cost  of 
nearly  $2  million.  This  simulator  was  designed  to  provide  a  pilot  with 
a  detailed  visual  encounter  with  the  lunar  surface;  the  machine  consisted 
primarily  of  a  cockpit,  a  closed-circuit  TV  system,  and  four  large  murals  or 
scale  models  representing  portions  of  the  lunar  surface  as  seen  from  various 
altitudes.  The  pilot  in  the  cockpit  moved  along  a  track  past  these  murals, 
which  would  accustom  him  to  the  visual  cues  for  controlling  a  spacecraft  in 
the  vicinity  of  the  moon.  Unfortunately,  such  a  simulation — although  great 
fun  and  quite  aesthetic — was  not  helpful  because  flight  in  lunar  orbit  posed 
no  special  problems  other  than  the  rendezvous  with  the  LEM,  which  the 
device  did  not  simulate.  Not  long  after  the  end  of  Apollo,  the  expensive 
machine  was  dismantled.34 

379 


Space/light  Revolution 


L-65-5579 

Although  as  much  fun  as  riding  through  the  fun  house  at  the  county  fair,  the 
$2  million  LOLA  proved  unnecessary.  In  this  photo  from  1965,  a  Langley  technician 
takes  great  care  to  make  sure  that  the  surface  features  of  the  moon  are  being 
represented  exactly. 


Rogallo's  Flexible  Wing 

More  than  any  other  division  at  Langley,  the  Full-Scale  Research  Division 
acted  as  a  "service  effort"  for  Apollo.  Testing  in  the  high-speed  wind 
tunnels  of  this  division  provided  essential  data  in  the  transonic  regime  for 
the  moon  shot.  Not  all  Apollo  work  carried  out  in  the  Full-Scale  Research 
Division,  however,  involved  high-speed  aerodynamics.  Perhaps  the  most 
interesting  and  potentially  significant  technologies  developed  in  this  division 
involved  low-speed  aerodynamics — specifically,  a  proposed  Apollo  capsule 
recovery  system  that  used  a  controllable  paraglider.  This  concept,  which 
was  eventually  turned  down  both  for  Gemini  and  Apollo,  was  the  brainchild 
of  Francis  M.  Rogallo,  an  ingenious  thinker  and  kite-flying  enthusiast  who 
worked  in  the  7  x  10-Foot  Tunnel  Branch. 

Although  Bob  Gilruth  and  many  other  engineers  responsible  for  Project 
Mercury  considered  the  ballistic  capsule  approach  "an  elegant  solution"  to 
the  problem  of  quickly  putting  a  person  in  orbit,  no  aeronautical  engineer 
was  especially  happy  with  the  plan.35  Their  dream  was  for  the  spacecraft 

380 


In  the  Service  of  Apollo 

to  return  to  earth  using  "wings  and  wheels" — that  is,  to  really  fly  down 
through  the  atmosphere  to  a  landing  on  a  conventional  runway. 

NASA  placed  its  hopes  for  such  an  airplane-like  landing  on  an  unusual 
inflated- fabric  flexible  wing,  or  parawing.  Such  a  wing  was  being  developed 
at  Langley  in  the  early  1960s  under  the  intellectual  direction  of  Francis 
M.  Rogallo.  Rogallo's  idea  for  Gemini,  as  well  as  for  Apollo,  was  to  pack 
away  a  carefully  designed  parawing  like  a  parachute  until  the  spacecraft  fell 
to  about  60,000  feet,  at  which  time  an  elaborate  unstowing  and  unfurling 
process  began.  By  20,000  feet,  if  all  went  well,  the  descending  spacecraft 
would  turn  into  the  world's  heaviest  hang  glider,  suspended  under  a  dart- 
shaped  parawing.  The  astronauts  themselves  would  then  bring  the  soaring 
craft  down  to  a  landing  either  on  water  or  on  soil.36 

Rogallo  had  started  at  NACA  Langley  in  1936  after  graduation  from 
Stanford  University,  and  since  1945  the  flexible  wing  had  been  a  pet  project. 
A  survey  of  the  history  of  the  parawing  provides  not  only  an  understand- 
ing of  the  genesis  of  one  of  Langley's  most  intriguing — if  never  used — 
developments  for  Apollo  but  also  insight  into  the  sudden  and  dramatic  im- 
pact of  Sputnik  and  the  spaceflight  revolution  on  the  course  of  independent 
research  at  Langley. 

Rogallo  and  his  wife,  Gertrude,  spent  their  spare  time  flying  home-built 
kites  at  their  beach  house  at  Nags  Head,  North  Carolina,  which  is  near 
Kitty  Hawk.  By  the  end  of  World  War  II,  this  hobby  had  begun  to  give  the 
couple  ideas  for  unconventional  vehicles,  such  as  hydrofoil  boats,  ground- 
effect  machines,  V/STOL  aircraft,  and  flexible  wings.  Because  they  could 
not  find  any  organization,  including  their  own  NACA,  to  support  R&D  for 
their  ideas,  they  "decided  to  do  what  we  could  privately  as  time  permitted." 
Initially  their  efforts  focused  on  configurations  resembling  boat  sails;  later, 
their  designs  were  similar  to  parachutes.  Finally,  they  concentrated  on 
shapes  somewhere  between  boat  sails  and  parachutes — flexible  wings.  By 
the  end  of  1948,  the  couple  had  developed  a  flexible  kite,  which  the  Rogallos 
called  "Flexi-Kite,"  and  a  type  of  gliding  parachute,  which  they  later  named 
a  "paraglider."37 

In  1948,  Rogallo  and  his  wife  filed  a  patent  for  a  V-shaped  flexible  wing, 
which  was  awarded  (U.S.  Patent  No.  2,546,078)  in  March  1951.  From 
the  outset,  the  inventors  had  thought  of  their  parawing  as  a  wing  not 
only  for  sport  gliders  but  also  for  military  and  civil  powered  aircraft.  No 
one,  however,  took  their  proposals  seriously.  As  Rogallo  remembers,  when 
meeting  friends  and  acquaintances,  they  were  generally  greeted  with,  "How's 
the  kite  business?"  The  Rogallos  had  resorted  to  selling  their  Flexi-Kite  as 
a  toy  in  order  to  illustrate  the  parawing  principle  and  help  finance  their 
work.  Francis  Rogallo  would  often  say  in  later  years  that  toys  should  copy 
the  real  thing  and  not  the  other  way  around.38 

For  the  first  seven  years  of  its  development,  the  motivation  behind  the 
flexible  wing  had  been  "purely  aeronautical,"  but  that  changed  in  1952 
when  the  Rogallos  saw  the  Colliers  magazine  that  ran  the  exciting  series 

381 


Space/light  Revolution 


Francis  and  Gertrude  Rogallo  (right),  in- 
ventors of  the  V-shaped  flexible  "para- 
wing."  In  December  1961,  Langley  flight- 
tested  a  50-foot  parawing  's  ability  to  bring 
down  safely  a  model  of  a  manned  space 
capsule  from  a  few  thousand  feet  above 
Plum  Tree  Island  (below),  an  old  army 
bombing  range  near  Langley  Field. 


L-63-5441 


L-61-8041 


382 


In  the  Service  of  Apollo 

of  stories  about  spaceflight.  Francis  Rogallo  was  struck  by  the  issue's 
beautiful  illustrations  of  rigid- winged  gliders  mounted  on  top  of  huge  rockets. 
As  he  recalled  later  in  a  1963  speech  to  the  American  Astronautical 
Society,  "I  thought  that  the  rigid-winged  gliders  might  better  be  replaced 
by  vehicles  with  flexible  wings  that  could  be  folded  into  small  packages 
during  the  launching."  In  August  1952  he  met  Dr.  Willy  Ley,  one  of 
Colliers  consultants,  and  told  Ley  his  thoughts  about  flexible  wings  for 
astronautics.  In  the  conversation  Rogallo  mentioned  that  the  technology  of 
flexible  wings  might  someday  prove  very  useful  when  spacecraft  commute 
regularly  between  planets:  a  rocket  ship  returning  from  Mars  could  pop  out 
flexible  wings  as  it  entered  the  earth's  atmosphere  and  glide  the  last  100 
or  200  miles  home,  saving  "the  stockholders"  that  much  fuel.  "But  the 
time  was  not  yet  ripe."39  (Note  that  Rogallo  imagined,  perhaps  in  jest, 
that  private  corporations  would  be  sponsoring  the  interplanetary  travel,  not 
governments.) 

In  April  1954,  hoping  to  gain  acceptance  of  his  concept  for  aerospace 
applications,  Rogallo  gave  a  presentation,  complete  with  glider  model 
demonstrations,  to  the  local  Tidewater  reserve  unit  of  the  Air  Force  Research 
and  Development  Command.  Two  months  later,  he  submitted  a  proposal 
to  include  parawing  research  in  the  NACA  budget,  but  the  proposal  was 
rejected.  Indefatigable,  he  submitted  a  proposal  to  discuss  his  flexible  wing 
concepts  at  the  annual  meeting  of  the  Institute  of  Aeronautical  Sciences 
(IAS).  This  was  "the  first  [proposal]  that  actually  reached  the  program 
committee  after  several  tries,"  but  it  too  was  turned  down.  The  IAS 
rejection  letter  read:  "Although  the  paper  is  out  of  the  ordinary  and  looks 
like  it  might  be  fine  to  hear,  it  just  does  not  fit  into  our  program."40 

As  it  did  for  so  many  research  projects,  the  launch  of  Russia's  Sputnik  1  in 
October  1957  changed  the  course  of  history  for  the  parawing.  Even  before 
the  formation  of  NASA  in  1958,  Rogallo  had  received  NACA  approval  to 
make  a  few  crude  wind-tunnel  and  model  flight  investigations  of  parawings 
in  the  7  x  10-Foot  Tunnel  Branch.  In  December  1958,  he  made  a  presen- 
tation to  the  Langley  Committee  on  Aerodynamics,  and  as  he  remembers, 
"gradually  people  in  other  divisions  became  interested  and  volunteered  to 
investigate  parawings  in  their  facilities."  During  1959  cloth  parawings  were 
tested  in  the  4- Foot  Supersonic  Pressure  Tunnel  at  Mach  2,  and  still  other 
parawing  models  were  deployed  at  high  altitudes  (150,000  to  200,000  feet) 
at  nearly  Mach  3  from  rocket  launchings  at  Wallops  Island.  In  August  1959, 
von  Braun  invited  Rogallo  to  Huntsville  for  a  presentation,  so  "business  was 
picking  up."41 

For  the  next  year  and  a  half,  into  early  1961,  Rogallo  gave  talk  after 
talk  on  his  parawing  concept  to  various  technical  groups.  He  spoke  at 
the  national  aeronautics  meeting  of  the  Society  of  Automotive  Engineers 
(April  1960);  at  Ryan  Aeronautical  Company  and  North  American  Aviation 
(May  1960);  at  the  annual  IAS  meeting  in  New  York  City  (Jan.  1961); 
and  at  local  IAS  chapter  meetings  in  Lancaster,  California,  and  San  Diego 

383 


Space/light  Revolution 

(March  1961).  By  the  end  of  1960,  the  Ryan  company,  the  same  company 
that  built  Charles  Lindbergh's  Spirit  of  St.  Louis,  began  building  a  powered 
man-carrying  "Ryan  Flex- Wing"  at  its  own  expense;  Rogallo  was  on  hand  in 
San  Diego  to  witness  its  first  flight.  Also,  in  early  1961,  NASA  Marshall  gave 
Ryan  and  North  American  contracts  to  study  the  feasibility  of  recovering 
Saturn  boosters  by  means  of  parawings.  NASA  in-house  studies  of  the 
technological  capabilities  of  the  wing  were  made  at  Marshall  and  Langley 
and  demonstrated  that  recovery  of  the  (later  canceled)  C-2  rocket  stage 
was  feasible.  By  the  end  of  1961,  the  DOD  let  its  first  parawing  contract, 
to  Ryan,  for  flight  tests  of  the  Flex- Wing.  The  aircraft  was  later  sent  to 
NASA  Langley  for  investigation  in  the  Full-Scale  Tunnel.  Thereafter,  the 
number  of  projects  and  contracts  related  to  parawings  increased  too  rapidly 
to  mention  in  this  brief  history.42 

"It  looked  like  parawings  were  here  to  stay,"  Rogallo  rejoiced  at  the  time, 
and  Sputnik  was  the  reason.43  By  the  summer  of  1963,  it  appeared  that  the 
concept  had  achieved  worldwide  acceptance  and  that  the  time  had  come  for 
his  parawing  study  group  to  give  the  U.S.  government  royalty- free  license  to 
use  its  patents,  which  it  did  in  a  ceremony  in  Washington  on  18  July  1963. 
In  a  short  speech,  Rogallo  expressed  his  hopes  for  the  invention:  "We  feel 
confident  that  the  civil  and  military  agencies  of  the  government  will  carry  on 
this  work,  and  we  hope  private  industry  will  promote  use  of  the  concept  for 
business  and  pleasure  as  effectively  as  they  have  for  astronautics  and  military 
aeronautics."  In  a  separate  ceremony  a  day  earlier,  Dr.  Hugh  Dryden,  depu- 
ty administrator  of  NASA,  presented  Francis  Rogallo  and  his  wife  with  a 
check  for  $35,000  for  their  development  of  the  flexible- wing  concept;  at  that 
time,  it  was  the  largest  cash  award  ever  made  by  the  space  agency  to  an 
inventor.44 

Unfortunately,  the  spaceflight  revolution,  which  had  so  quickly  turned 
circumstances  in  the  wing's  favor,  just  as  quickly  turned  circumstances 
against  it.  That  is  often  the  nature  of  revolutions — to  take  things  full 
circle.  From  the  beginning,  NASA's  interest  in  Rogallo's  paraglider  grew 
primarily  from  the  possibility  of  using  it  as  a  controllable  space  capsule 
recovery  system.  When  that  interest  waned,  so  too  did  NASA's  support  for 
the  innovative  flight  technology. 

Given  NASA's  formal  go-ahead  for  research,  Rogallo  and  his  colleagues  in 
the  Full-Scale  Research  Division  invested  much  time,  energy,  and  emotion  in 
the  paraglider  concept.  Several  Langley  employees  shared  Rogallo's  enthusi- 
asm for  the  innovative  flight  technology  and  even  conducted  manned  flexible- 
wing  flight  research  during  weekends  on  the  Outer  Banks  with  privately 
owned  equipment.  Although  qualitative  in  nature,  these  investigations 
proved  "valuable  in  providing  quick  answers  and  indicating  promising  di- 
rections for  the  much  more  costly  and  time-consuming  instrumented  but 
unmanned  NASA  flight  research."45  In  wind-tunnel  studies  at  Langley, 
this  research  covered  a  broad  spectrum  of  parawing  design  parameters — 
everything  from  the  original  concept  of  a  flexible  lifting  surface  (indicated 

384 


In  the  Service  of  Apollo 

in  the  engineering  data  as  a  "limp  paraglider")  to  rigid  frame  gliders  with 
conical  and  cylindrical  canopies. 

As  this  research  on  the  basic  technology  of  the  parawing  gained  momen- 
tum at  Langley,  NASA's  STG,  still  at  Langley  at  this  time,  grew  inter- 
ested in  the  possible  application  of  the  foldable,  deployable,  inflatable-frame 
paraglider  to  its  Gemini  EOR  program.  Specifically,  the  STG  believed  it 
might  be  used  as  part  of  Mercury  Mark  II,  the  follow-on  to  Project  Mercury, 
which  ultimately  (in  January  1962)  became  Project  Gemini.  The  STG  felt 
that  such  a  wing  could  be  deployed  either  before  or  after  reentry  to  provide 
controlled  glide  and  horizontal  landing.  Even  on  a  lifting  reentry  body — 
NASA  was  giving  "lifting  body"  technology  considerable  attention  in  re- 
lation to  space  station  studies  during  this  period  (see  the  epilogue) — tests 
at  Langley  and  other  NASA  facilities  were  showing  that  a  parawing  could 
improve  the  postentry  flight  or  landing  characteristics.46 

In  early  July  1961,  a  few  weeks  before  the  second  manned  Mercury  flight 
by  Gus  Grissom,  Gilruth's  organization  initiated  three  well-funded  design 
study  contracts  on  the  paraglider  concept  with  Ryan,  North  American,  and 
Goodyear.  Of  these  three  companies,  North  American  would  eventually 
produce  the  most  acceptable  plan — a  study  to  explore  the  parawing  as  an 
earth-landing  system  for  Project  Apollo.47  A  few  weeks  later,  the  STG 
began  requesting  that  studies  of  the  Rogallo-type  paraglider  be  conducted 
at  NASA  centers.  At  Langley  this  led,  on  31  August,  to  a  research 
authorization  for  "Free-Flight  and  Wind- Tunnel  Tests  of  Guided  Parachutes 
as  Recovery  Devices  for  the  Apollo  Type  Reentry  Vehicle."  By  late  fall,  all 
of  this  work  came  together  as  a  formalized  NASA  paraglider  development 
program,  with  Langley  and  Ames  responsible  for  the  wind-tunnel  tests  and 
the  Flight  Research  Center  for  scheduling  manned  flight  tests.  Starting  in 
mid- 1963,  12  manned  flight  tests  were  actually  made  at  Edwards  with  a 
so-called  Parasev.48 

If  the  United  States  had  not  been  in  a  hurry  to  go  to  the  moon,  the 
Rogallo  paraglider  might  have  been  used  as  the  capsule  recovery  system 
for  Gemini  and  Apollo;  of  course,  if  the  country  had  not  been  in  such  a 
hurry,  it  would  not  have  gone  to  the  moon  at  all  in  the  1960s — and  perhaps 
would  not  have  gone  there  ever.  As  it  turned  out,  the  paraglider  became 
"hopelessly  snarled  in  a  financial,  technical,  and  managerial  morass."^ 
Richard  Hallion  recollects  the  specific  problems  encountered  during  the  flight 
tests  at  Edwards: 

Paraglider  development  involved  solving  major  design  difficulties  in  deploying  the 
wing,  ensuring  that  crew  would  have  adequate  control  over  the  parawing-equipped 
craft,  and  providing  stability,  control,  and  handling  qualities.  The  Flight  Research 
Center's  technical  staff  was  never  convinced  that  the  scheme  was  workable.  Eventu- 
ally, because  of  poor  test  results  and  rising  costs  and  time  delays,  the  idea  was  dropped 
from  Gemini  in  mid-1964.  FRC  engineers  and  pilots  had  believed  that  any  vehicle 
so  equipped  might  present  a  pilot  with  a  greater  flying  challenge  than  contemporary 
advanced  airplanes. 

385 


Space/light  Revolution 


L-62-631 

An  early  version  of  the  single-seat  Paraglider  Research  Vehicle  ("Parasev")  is  test 
"flown"  in  Langley's  Full- Scale  Tunnel  in  January  1962. 


These  conclusions  were  based  on  experience.  Flights  with  the  small,  single- 
seat  experimental  Parasev  had  proved  extremely  tricky  even  in  the  hands  of 
experienced  test  pilots.  The  first  machine,  Parasev  /,  flew  as  if  "controlled  by 
a  wet  noodle."  As  Hallion  records,  during  one  ground  tow,  a  veteran  NASA 
test  pilot  "got  out  of  phase  with  the  lagging  control  system  and  developed 
a  rocking  motion  that  got  worse  and  worse;  just  as  the  tow  truck  started 
to  slow,  the  Parasev  did  a  wing-over  into  the  lakebed,  virtually  demolishing 
the  Parasev  and  injuring  [the  pilot],  though  not  seriously."  This  was  not 
the  only  time  that  a  paraglider  test  vehicle  would  slam  into  the  ground.50 

The  Parasev,  built  and  rebuilt  several  times,  eventually  made  over 
100  flights  at  Edwards  and  showed  enough  progress  that  it  might  have  proved 
feasible  for  capsule  reentry  if  further  developed.  However,  NASA  could 
not  wait  for  its  maturation.  Besides,  the  paraglider  was  "not  absolutely 
necessary,  being  more  technological  frosting  than  cake."51  NASA  did  not 
need  an  elegant  reentry  plan,  just  a  workable  one.  By  early  1964,  NASA  was 
committed  to  a  water  landing  for  Apollo.  In  mid- 1964,  Gemini's  program 

386 


In  the  Service  of  Apollo 

manager,  Charles  W.  Mathews,  a  former  Langley  STG  engineer,  canceled 
the  paraglider.  Rogallo's  idea  would  not  fly  anyone  or  anything  back  from 
space. 

Rogallo  never  gave  up  on  his  pet  concept  and  continued  to  develop  it 
even  after  he  retired  and  moved  with  Gertrude  to  Nags  Head.  There  they 
spent  all  their  time  working  on  their  paragliders  for  sport  aviation  and  other 
applications.  Before  leaving  NASA  Langley,  Rogallo  and  his  colleagues 
in  the  Low- Speed  Vehicle  Branch  had  continued  to  explore  a  very  broad 
spectrum  of  wing  shapes  and  structures  for  his  flexible  wings.  Never  again, 
however,  would  his  concept  receive  the  same  high  level  of  NASA  support 
and  funding  that  it  had  received  when  linked  to  the  manned  space  programs 
of  the  early  1960s.  Nevertheless,  a  Parawing  Project  Office  (under  engineer 
Dewey  L.  Clemmons,  Jr.)  continued  at  Langley  until  1967  and  kept  the 
research  alive. 

The  Apollo  Fire  Investigation  Board 

Langley  played  one  other  significant,  if  very  sad,  role  in  the  Apollo 
program.  The  program  seemed  to  be  moving  along  extremely  well,  so  well 
in  fact  that  by  New  Year's  Day  1967  many  observers  believed  that  President 
Kennedy's  "by  the  end  of  the  decade"  deadline  for  landing  a  man  on  the 
moon  might  be  achieved  a  couple  of  years  ahead  of  schedule.  Then  tragedy 
struck — making  it  abundantly  clear  to  NASA  and  the  nation  just  how  high  a 
price  would  have  to  be  paid  to  pursue  such  bold  ventures  into  the  unknown. 
Early  in  the  evening  of  27  January  1967,  a  fire  broke  out  inside  the  Block  I 
Command  Module  sitting  on  top  of  the  uprated  Saturn  I  204  rocket  on  the 
launchpad  of  Complex  34  at  Cape  Kennedy.  The  fire  killed  three  Apollo 
astronauts — Gus  Grissom,  Edward  H.  White,  and  Roger  B.  Chaffee — who 
were  in  the  capsule  for  a  prelaunch  test.52 

The  next  day,  Deputy  Administrator  Robert  C.  Seamans,  Jr.,  speaking 
for  Administrator  Webb,  named  Floyd  L.  Thompson  chairman  of  the 
Apollo  204  Review  Board.53  Serving  with  Thompson  were  five  other  NASA 
officials,  one  air  force  official,  and  one  official  from  the  U.S.  Bureau  of 
Mines.  No  one  institution  was  better  represented  on  the  board  than 
Langley.  Besides  Thompson  (and  probably  because  of  him),  the  board  also 
included  E.  Barton  Geer,  associate  chief  of  Langley's  Flight  Vehicles  and 
Systems  Division,  and  George  Malley,  Langley's  chief  counsel.  Furthermore, 
Max  Faget,  then  of  the  Manned  Spacecraft  Center,  but  formerly  a  Langley 
engineer,  also  served  on  the  board.* 


*  The  other  members  of  the  Apollo  204  Review  Board  were  Col.  Charles  F.  Strang,  chief  of  the 
Missiles  and  Space  Safety  Division,  air  force  inspector  general,  Norton  AFB,  Calif.;  Lt.  Col.  Frank 
Borman,  astronaut,  Manned  Spacecraft  Center;  George  C.  White,  Jr.,  director  of  reliability  and  quality, 
Apollo  Program  Office,  NASA  headquarters;  Dr.  Robert  W.  Van  Dolah,  research  director,  Explosive 
Research  Center,  Bureau  of  Mines,  U.S.  Department  of  Interior,  Pittsburgh,  Pa.;  and  John  J.  Williams, 
director  of  Spacecraft  Operations,  NASA  Kennedy  Space  Center. 

387 


Spaceflight  Revolution 


L-68-9680 

Floyd  Thompson  (left),  chairman  of  the  Apollo  Fire  investigation  board,  talks  with 
Thomas  O.  Paine  (right),  who  took  over  from  James  Webb  as  NASA  administrator 
in  September  1968. 


NASA  was  handing  a  terribly  difficult  job  to  Langley's  69-year-old  di- 
rector, who  was  only  a  year  and  some  months  away  from  retirement.  Iron- 
ically, someone  like  Floyd  Thompson,  with  no  national  public  reputation 
and  a  surface  personality  that  even  some  close  friends  described  as  a  bit 
"hayseed,"  would  have  been  out  of  the  question  for  the  investigation  com- 
mittee that  was  appointed  after  the  Challenger  accident — even  as  a  commis- 
sion member  let  alone  as  its  chairman.54  But,  as  the  results  of  the  Apollo 
fire  investigation  and  the  subsequently  successful  Apollo  program  demon- 
strated, Thompson  would  be  the  perfect  man  for  the  Apollo  job.  As  Robert 
Seamans  once  said  about  Thompson,  "I  think  he  acted  to  fool  people  a 
little  bit  so  he  could  get  their  measure — and  then  watch  out,  you  know. 
Very  adaptable.  At  Langley,  through  all  this  chaotic  period,  he  kept  it  out 
front  doing  the  right  kind  of  things.  And  then  of  course  the  real  crunch 
came  when  we  had  the  Apollo  fire.  The  question  was,  'Who  do  we  have  who 


388 


In  the  Service  of  Apollo 

has  the  ability  and  the  credibility  to  be  responsible  for  that  review?'  That's 
when  we  put  Tommy  into  that  very,  very  difficult  job."55  Interestingly,  after 
the  mishap  of  Apollo  13,  when  exploding  oxygen  tanks  in  the  service  module 
forced  the  highly  dramatic  return  of  the  three  astronauts  even  before  they 
had  reached  the  moon,  NASA  appointed  another  Langley  center  director, 
Edgar  M.  Cortright,  Thompson's  successor,  to  chair  the  accident  review 
board.  Perhaps  this  was  in  part  a  testimony  to  how  well  Floyd  Thompson 
had  conducted  the  Apollo  fire  investigation. 

Within  24  hours  of  the  command  module  inferno,  Thompson  and  the  rest 
of  his  committee  were  on  hand  at  Pad  34,  beginning  the  long  and  arduous 
process  of  finding  out  why  the  tragedy  had  happened.  Under  Thompson's 
disciplined  direction,  the  investigation  marched  along  quickly  and  quietly. 
This  low-profile  progress  was  possible  because  the  inquiry  was  an  internal 
NASA  investigation  without  the  national  media  exposure  that  those  on  the 
Challenger  investigation  committee  would  face.  By  5  April,  after  spending 
about  10  weeks  on  the  job,  most  of  it  on  site  at  the  Kennedy  Space  Center, 
the  Apollo  204  Review  Board  was  ready  to  submit  its  formal  report,  which 
was  several  thousand  pages  long  including  appendixes.  According  to  its 
terse  prose,  arcs  from  faulty  electrical  wiring  in  an  equipment  bay  inside 
the  command  module  had  started  the  fire.  In  the  100-percent  oxygen 
atmosphere,  the  crew  had  died  of  asphyxia  caused  by  inhalation  of  toxic 
gases.  The  board  report  concluded  with  a  list  of  11  major  recommendations 
for  hardware  and  operational  changes.56 

NASA  would  need  two  more  years  to  fix  all  the  problems  with  Apollo.  A 
special  Apollo  Configuration  Control  Board,  chaired  by  George  Low,  even- 
tually oversaw  the  completion  of  over  1300  design  changes  for  the  space- 
craft. The  mending  process  had  really  begun,  however,  with  the  Apollo  204 
Review  Board's  fast  action  to  a  first  draft  of  an  investigation  report.  As 
chairman  and  as  a  person  well  aware  of  the  inherent  dangers  of  flight  re- 
search, Floyd  Thompson  wanted  everyone  to  know  that  his  board's  written 
description  of  "the  defects  in  the  Apollo  Program  that  led  to  the  condi- 
tion existing  at  the  time  of  the  Apollo  204  accident"  should  not  be  in- 
terpreted as  "an  indictment  of  the  entire  manned  space  flight  program" 
or  as  a  "castigation  of  the  many  people  associated  with  that  program." 
"Nothing  is  further  from  the  Board's  intent,"  Thompson  emphatically  de- 
clared. The  function  of  his  board  had  been  "to  search  for  error  in  the 
largest  and  most  complex  research  and  development  program  ever  under- 
taken," and  that  was  why  the  report  on  the  fire  commented  only  on  the 
deficiencies  uncovered  and  did  not  present  a  total  picture  of  the  Apollo 
program,  including  the  good  points  with  the  bad,  or  look  for  scapegoats. 
However,  the  report  tried  to  make  clear  to  the  nation  that  such  tragedies 


389 


Space/light  Revolution 


All  three  astronauts  who  died  in  the 
Apollo  fire  had  spent  a  consider- 
able amount  of  time  in  simulators 
at  Langley.  Gus  Grissom  (right) 
at  the  controls  of  the  Rendezvous 
and  Docking  Simulator  in  1963;  be- 
low, Roger  Chaffee  strapped  into  the 
Lunar  Landing  Research  Facility's 
Reduced  Gravity  Walking  Simulator 
in  1965.  Unfortunately,  a  picture  of 
Edward  White  while  in  training  at 
Langley  was  not  found. 


L-65-8545 


390 


In  the  Service  of  Apollo 
would  occasionally  occur  if  the  bold  venture  into  space  was  to  continue  and 

j-riy 

progress. 

It  is  unfortunate  that  tragic  accidents  such  as  the  Apollo  fire  and 
the  Challenger  explosion  have  to  happen  for  errors  to  be  discovered  and 
corrected.  Both  events  made  NASA  and  its  contractors  more  cautious,  and 
in  the  case  of  the  Apollo  fire,  they  actually  slowed  the  pace  of  work  so  that 
tasks  could  be  performed  more  carefully.  "It  gave  everyone  not  working  on 
fire-related  matters  a  breather,  a  period  to  catch  up  on  their  work."  "In 
the  race  for  the  moon,"  as  Apollo  astronaut/historian  Michael  Collins  has 
written,  "no  one  wanted  his  piece  of  the  machine  to  be  the  laggard,  the  one 
to  hold  up  the  whole  procession.  Consequently,  no  one  wanted  to  admit 
being  'the  long  pole  in  the  tent'  as  it  was  called,  and  managers  were  apt 
to  fudge  their  schedules  a  bit,  hoping  someone  else  would  admit  to  being 
even  farther  behind.  Many  long  poles  got  whittled  down  to  manageable 
size  during  the  time  North  American  was  struggling  to  get  the  Command 
Module  back  on  track."58 

This  scenario  really  did  not  apply  to  NASA  Langley;  its  work  to  achieve 
the  lunar  landing  objective  was  for  the  most  part  over  by  the  time  the  Apollo 
command  module  was  fixed  and  ready  for  its  first  manned  flight  (Apollo  7, 
11-22  Oct.  1968).  Langley's  contributions  to  Apollo  had  little  to  do  with 
final  preparations  but  rather  rested  largely  in  the  groundwork  for  such  an 
ambitious  program.  By  the  time  of  Apollo  11  's  historic  first  lunar  landing 
on  20  July  1969,  Langley's  multifarious  R&D  efforts  for  Apollo  had  been 
largely  forgotten;  except  for  the  Hampton  area  press,  the  media  gave  the 
center  little  attention.  But  without  Langley,  an  American  lunar  landing 
that  summer  day  may  not  have  been  possible. 

"We  had  a  target  and  a  goal,"  says  John  Houbolt,  one  of  the  few  from 
Langley  privileged  with  an  invitation  to  watch  the  historic  lunar  landing 
event  from  the  viewing  room  at  Mission  Control  in  Houston.  "Congress 


By  the  authority  granted  to  the  center  in  a  letter  from  Deputy  Administrator  Robert  Seamans  on 
27  February  1967,  NASA  Langley  became  "the  custodian  of  all  pertinent  physical  evidence,  reports,  files, 
and  working  materials  dealing  with  the  investigations  and  review  of  the  Apollo  204  accident."  (Copy  in 
Apollo  204  Review  Board  files,  Langley  Central  Files.)  In  1978,  Langley  shipped  all  the  documentary 
records  of  the  review  board  to  the  National  Archives;  however,  it  kept  all  the  related  hardware,  including 
the  Apollo  capsule  itself.  In  1990  preparations  were  made  to  send  the  "Apollo  storage  container,"  which 
included  the  capsule,  to  the  Kennedy  Space  Center  for  an  appropriate  burial  with  remnants  of  the  Space 
Shuttle  Challenger.  However,  those  preparations  halted  after  a  controversy  ensued  over  the  historical 
significance  of  the  Apollo  capsule  and  its  possible  use  in  a  museum  exhibition.  Thus,  in  the  summer 
of  1990,  NASA  made  the  decision  to  keep  the  Apollo  hardware  right  where  it  was,  in  a  warehouse 
at  Langley.  (The  press  was  allowed  to  view  the  remains  briefly,  in  part  to  confirm  that  Langley  still 
possessed  them.)  On  7  November  1990,  Langley  Director  Richard  H.  Petersen  ordered  his  director  of 
operations  "to  seal  the  entrance  to  the  Apollo  204  storage  container"  and  not  to  break  it  "without  my 
written  approval."  For  the  relevant  documentation,  see  the  Apollo  204  Review  Board  files,  Langley 
Central  Files. 

391 


Space/light  Revolution 


L-67-1697 

The  Lunar  Landing  Research  Facility  staff  crowds  around  Apollo  astronaut  Neil 
Armstrong  (center)  in  March  1967,  28  months  before  he  was  to  become  the  first 
human  to  set  foot  on  the  moon. 


was  behind  it.  Funding  was  available.  The  entire  nation  mobilized  for  a 
common  goal."  In  his  opinion,  to  this  day,  "the  landing  on  the  Moon  was 
undoubtedly  mankind's  greatest  technological  achievement  and  engineering 
accomplishment.  We  started  essentially  from  scratch  in  1962  and  seven  years 
later  we  were  on  the  Moon.  It  was  a  remarkable  achievement  and  remains 
unsurpassed."  59 

Indeed,  Apollo  was  the  crowning  achievement  of  the  spaceflight  revolu- 
tion, but  it  had  also  served  as  NASA's  only  guiding  star  through  nearly  all 
of  the  1960s.  Apollo  shone  so  spectacularly,  few  aboard  NASA  suspected 
that  it  would  ever  dim.  Unfortunately,  near  the  end  of  the  program,  in- 
terest in  spaceflight  waned,  and  Apollo's  brightness  proved  to  be  that  of  a 
supernova — dazzling  yet  brief.  Once  Apollo  had  faded,  NASA  found  itself 
traveling  without  direction. 


392 


12 


The  Cortright  Synthesis 


A  new  era  had  already  started.  All  I  did  was  acceler- 
ate it.  Thompson  was  wedded  to  the  Langley  way  of 
doing  things  and  under  Tommy  everything  would  be 
done  well,  but  done  their  way  and  in  their  good  time. 
All  I  did  was  speed  it  up,  I  think,  what  was  bound  to 
happen  eventually. 

—Edgar  M.  Cortright,  Langley  director 

Cortright  came  down  here  with  the  idea  that  he  knew 
everything  and  that  he  was  going  to  control  every- 
thing. And  this  made  it  very  difficult  for  some  of  us. 
Some  of  us  could  adjust  to  it,  some  of  us  didn't.  I'm 
one  of  those  who  didn't. 

— Laurence  K.  Loftin,  Jr.,  former 
Langley  director  of  aeronautics 


Every  revolution  needs  its  culminating  figure,  its  rationalizer,  its 
Napoleon,  who  synthesizes  chosen  elements  of  the  old  regime  and  the  rev- 
olution to  create  a  new  order.  Under  the  firm  and  confident  direction  of 
this  dynamic  leader,  a  revolutionary  episode  calms  down,  grows  structured, 
becomes  what  is  expected,  and  establishes  norms.  The  revolution  eventually 
becomes  the  social  and  intellectual  world  in  which  the  new  generation  lives, 
awaiting  the  next  major  upheaval. 

For  NASA  Ames  Research  Center  in  Sunnyvale,  California,  this 
culminating — and  dominating — figure  was  Dr.  Hans  Mark,  who  succeeded 
NACA  Langley  veteran  H.  Julian  "Harvey"  Allen  as  center  director  in 
February  1969. 1  For  Langley,  it  was  Edgar  M.  Cortright,  who  would  serve 
as  Langley  center  director  from  May  1968  to  August  1975.  In  the  first 
36  months  of  his  tenure,  Cortright  put  Langley  through  the  most  sweeping 

393 


Space/light  Revolution 

reorganization  in  the  center's  history,  be  it  NACA  or  NASA.  At  the  end  of 
it,  Langley  was  not  the  same  place  it  had  been.  Many  of  Langley's  most 
vital  links  to  the  old  culture  of  NACA  research  were  eliminated  or  retired. 
In  their  place  would  be  established  a  still  very  reputable  and  effective  orga- 
nization but  one  completely  adapted  to — even  tamed  by — the  criteria  and 
standards  set  by  the  spaceflight  revolution. 

Putting  a  comprehensive  treatment  of  the  Cortright  reorganization  and 
its  aftermath  at  the  end  of  this  long  study  of  the  spaceflight  revolution 
at  Langley,  however,  would  be  like  piggybacking  a  complete  study  of  the 
Napoleonic  period  on  top  of  a  history  of  the  French  Revolution  from  the 
fall  of  the  Bastille  to  the  coup  d'etat  of  18  Brumaire.  The  two  subjects, 
although  intricately  related,  need  to  be  treated  separately  because  they  are 
both  so  vast.  In  this  conclusion,  the  reader  will  find  discussed  only  a  few 
of  Cortright 's  changes  and  their  ramifications.  I  include  them  to  illustrate 
the  dialectical  process  by  which  Cortright  institutionalized  the  spaceflight 
revolution  at  Langley. 


The  Stranger 

Up  to  the  time  of  Cortright,  whenever  a  Langley  center  director  left  his 
job,  he  had  been  replaced  by  someone  already  working  at  the  laboratory. 
In  1925  upon  the  resignation  of  Leigh  M.  Griffith,  young  Henry  Reid,  an 
electrical  engineer  who  had  been  working  in  Langley's  instrument  research 
laboratory  since  1921,  became  the  engineer  in  charge.  In  1960  upon  Reid's 
long-awaited  retirement,  Floyd  Thompson  succeeded  him;  Thompson  had 
been  working  at  Langley  since  1927,  and  for  the  past  several  years  he  had 
been  Reid's  associate  director.  In  office,  Thompson  immediately  faced  the 
problem  of  naming  his  own  second-in-command.  Prior  to  the  spaceflight 
revolution,  the  director  had  always  given  this  position  to  a  close  and  trusted 
associate,  someone  who  had  been  working  at  Langley  for  some  time.  But 
in  the  new  political  and  bureaucratic  environment  of  NASA,  Thompson 
hesitated.  For  over  a  year,  he  acted  as  his  own  associate  director,  naming  no 
one  to  replace  him  in  his  old  position  until  he  could  thoroughly  think  through 
the  appointment.  Cagey  Thompson  was  considering  an  unprecedented 
move:  the  appointment  of  a  non-Langley  person,  Dr.  Ernst  Stuhlinger 
of  NASA  Marshall  Space  Flight  Center.  By  naming  Stuhlinger,  one  of 
von  Braun's  rocketeers,  Thompson  would  prove  to  NASA  headquarters 
that  he  was  not  so  parochial  as  to  only  consider  Langley  researchers  for 
the  job.  Approached  confidentially  so  that  no  one  at  Langley  would  hear 
about  the  offer  until  it  was  finalized,  Stuhlinger  eventually  turned  down  the 
job.  Thus,  almost  no  one  heard  about — or  even  now  know  of — the  offer 
to  Stuhlinger.  Only  after  Stuhlinger's  refusal  did  Thompson  turn  to  his 
talented  young  friend  at  Langley,  Charles  Donlan.  By  selecting  Donlan  as 


394 


The  Cortright  Synthesis 

the  associate  director,  according  to  Langley  tradition,  Donlan  was  anointed 
as  Thompson's  heir  apparent. 

Donlan,  however,  was  never  given  the  directorship  because  the  spaceflight 
revolution  would  interfere  with  the  tradition  of  succession  at  Langley.  In 
March  1968,  NASA  named  not  Donlan  but  Edgar  M.  Cortright,  a  virtual 
stranger  to  Langley,  as  the  center's  new  director.  Donlan  found  himself  out 
in  the  cold;  he  soon  left  Langley  to  serve  as  deputy  associate  administrator 
for  manned  spaceflight  at  NASA  headquarters.  Donlan  did  this  even  before 
Cortright  named  another  outsider,  Oran  W.  Nicks,  his  former  assistant  in 
the  office  of  unmanned  spaceflight  in  Washington,  as  his  associate  director. 
Floyd  Thompson  made  a  fuss  over  none  of  this;  after  all,  eight  years  earlier, 
he  had  himself  tried  to  bring  in  Stuhlinger  as  his  number  two  man.  Moreover, 
Thompson  had  not  retired  voluntarily  as  Langley's  director.  Instead,  NASA 
headquarters  announced  unilaterally  that  "Dr.  Floyd  L.  Thompson,  Director 
of  Langley  Research  Center,  will  retire  when  he  reaches  the  age  of  70 
on  November  25,  1968,  and  that  Edgar  M.  Cortright,  Deputy  Associate 
Administrator  for  Manned  Space  Flight  at  NASA  Headquarters,  will  replace 
him  as  Director  of  Langley  Center  on  May  1,  1968."  This  would  enable 
Thompson,  NASA  headquarters  said,  to  "utilize  a  large  part  of  his  time 
on  agency- wide  planning  and  evaluation  activities."  His  first  extra-Langley 
task  was  to  be  as  special  consultant  to  the  NASA  administrator,  Dr.  Thomas 
O.  Paine,  who  in  March  had  succeeded  James  Webb.2 

In  certain  respects  the  coming  of  Cortright  was  in  keeping  with  the 
Langley  tradition.  Like  Thompson  and  Henry  Reid,  he  was  an  engineer 
whose  first  professional  employment  had  been  with  the  NACA.  After  grad- 
uating with  a  bachelor's  degree  in  aeronautical  engineering  from  Rensselaer 
Polytechnic  Institute  in  1947,  Cortright  had  gone  to  work  at  the  NACA's 
Lewis  laboratory,  where  he  had  specialized  in  the  propulsion  aerodynamics  of 
supersonic  aircraft  and  guided  missiles.  While  in  Cleveland,  he  had  become 
protege  to  Abe  Silverstein,  Lewis's  dynamic  associate  director.  Because 
Silverstein  had  worked  at  Langley  from  1929  to  1943,  Cortright  became 
familiar  with  many  of  the  traditions  of  the  NACA's  first  laboratory.3 

In  keeping  with  the  spaceflight  revolution,  however,  the  coming  of 
Cortright  also  meant  dramatic  change.  Unlike  his  two  predecessors  as  direc- 
tor, he  had  never  worked  at  Langley.  Instead  of  making  his  way  to  a  high 
position  through  leadership  in  the  laboratory's  general  research  program, 
Cortright  had  earned  the  directorship  through  his  project  management  work 
at  NASA  headquarters.  When  Abe  Silverstein  came  to  Washington  in  the 
summer  of  1958  to  prepare  for  the  transition  to  NASA,  he  had  brought  young 
Cortright  with  him.  For  most  of  his  years  in  Washington,  Cortright  was  asso- 
ciated with  the  unmanned  space  program — including  the  Mariner,  Ranger, 
and  Surveyor  projects.  In  that  program,  his  boss  was  Homer  E.  Newell, 
the  former  chief  scientist  at  NRL.  (The  core  staff  of  Goddard  Space  Flight 
Center  had  come  from  NRL,  an  organization,  as  we  have  seen,  that  would 
often  be  at  odds  with  Langley.)  In  1963,  Cortright  became  NASA's  deputy 

395 


Spaceflight  Revolution 


J&ngky 


VOLUME  7,  NUMBER  5 


MARCH  22,  1968 


DR.  THompson,  ran.  ooninn  in  TOP  nnsn  POSTS,- 

EDGRR  HI.  CORTRICHT  (MIIIED  HEW  IBHCIEV  DIRECTOR 


Edgar  M.  Cortright  Charles  J. 

The  National  Aeronautics  and  Space  Administration  an- 
nounced this  week  that  Dr.  Floyd  L.  Thompson,  Director 
of  its  Langley  Research  Center,  will  retire  when  he  reaches 
the  age  of  70  on  November  25,  1968,  and  that  Edgar  M. 
Cortright,  Deputy  Associate  Administrator  for  Manned 
Space  Flight  at  NASA  Headquarters,  will  replace  him  as 
Director  of  the  Langley  Center  on  May  1,  1968,  enabling 
Dr.  Thompson  to  utilize  a  large  part  of  his  time  on  agency- 
wide  planning  and  evaluation  activities. 

Dr.  Thompson  will  continue  to  serve  as  Special  Assistant 
to  the  Administrator  until  his  retirement  and  will  serve 
as  Consultant  to  the  Administrator  on  a  part-time  basis 
after  retirement.  On  May  1,  Charles  J.  Donlan,  Deputy  Di- 
rector of  the  Langley  Research  Center,  will  transfer  from 
the  Langley  Center  to  NASA  Headquarters  to  serve  as 
Deputy  Associate  Administrator  for  Manned  Space  Flight 
(Technical). 

Replacing  Cortright  May  1  as  Deputy  Associate  Admin- 
istrator for  Manned  Space  Flight  will  be  Charles  Mathews, 
Director,  Apollo  Applications  Program,  and  formerly  Pro- 
gram Manager  for  Gemini. 

The  successor  to  Mathews  will  be  Harold  Luskin,  former 
Chief  of  Advanced  Design  Engineering  of  Lockheed-Cali- 
fornia Company  and  the  former  President  of  the  American 
Institute  of  Aeronautics  and  Astronautics.  Until  May  1, 
Luskin  will  serve  as  Deputy  Associate  Administrator  for 
Manned  Space  Flight  (Technical). 


Donlan  Dr.  Floyd  L.  Thompson 

As  Program  Manager  for  Apolio  Applications,  Luskin 
will  have  responsibility  for  planning  and  carrying  out  the 
utilization  of  the  large  Saturn  boosters  and  the  Apollo 
spacecraft  systems  in  the  period  following  the  demon- 
stration of  the  capability  of  this  equipment  to  land  men  on 
the  moon  and  return  them  safely  to  earth. 

In  October  1967,  Cortright  joined  the  Office  of  Manned 
Space  Flight  in  NASA  Headquarters  as  Deputy  Associate 
Administrator.  In  this  position,  he  serves  as  general  man- 
ager of  NASA's  program  for  manned  space  flight. 

Cortright  joined  the  former  National  Advisory  Committee 
for  Aeronautics  (predecessor  to  NASA)  as  an  aeronautical 
research  scientist  at  the  Lewis  Flight  Propulsion  Lab- 
oratory (now  Lewis  Research  Center),  Cleveland,  Ohio, 
in  1948.  From  1949  to  li»54,  he  was  head  of  the  Small 
Supersonic  Tunnels  Section;  from  1954  to  1958,  he  was 
Chief  of  the  8-  by  6-foot  Supersonic  Wind  Tunnel  Branch. 
In  January  1958,  he  was  appointed  Chief  of  the  Plasma 
Physics  Branch. 

Upon  NASA  s  creation  on  October  1,  1958,  Cortright  be- 
came Chief  of  the  Advanced  Technology  Programs  and 
directed  initial  formulation  of  NASA's  Meteorological 
Satellite  Program,  including  Tiros  and  Nimbus.  In  Feb- 
ruary 1960,  he  became  Assistant  Director  for  lunar 
and  planetary  programs  in  the  former  Office  of  Space 
Flight  Programs.  In  this  capacity,  he  directed  the  plan- 
(Continued  on  page  3) 


Langley  trumpets  a  changing  of  the  guard,  22  March  1968. 


396 


The  Cortright  Synthesis 

associate  administrator  for  space  sciences  and  applications.  Just  before  com- 
ing to  Langley,  he  became  deputy  associate  director  of  the  Office  of  Manned 
Space  Flight  under  George  E.  Mueller. 

Throughout  his  stay  at  NASA  headquarters,  which  essentially  spanned 
the  era  of  the  spaceflight  revolution,  Cortright  had  been  involved  with  the 
management  of  an  unmanned  space  program  that,  with  the  exception  of 
the  Lunar  Orbiter  project,  did  not  involve  Langley.  To  make  matters 
worse,  as  far  as  Langley  veterans  were  concerned,  by  the  mid-1960s,  Webb's 
organization  in  Washington  had  strong  feelings  that  Langley  had  gone  its 
own  way  too  often  under  Thompson's  paternalistic  and  rather  closed  NAG  A 
style  of  management.  This  group  within  headquarters,  which  included 
young  Cortright,  felt  that  Langley  needed  to  be  brought  under  tighter 
central  control.  When  Webb  picked  Cortright  to  replace  Thompson,  he 
was  essentially  sending  him  to  the  center  to  ensure  such  waywardness  did 
not  continue — to  bring  Langley  into  the  fold. 

Webb  did  not  draft  Cortright  for  the  job;  Cortright  eagerly  applied  for  it. 
His  last  assignment  at  NASA  headquarters,  as  George  Mueller's  deputy  in 
the  Office  of  Manned  Space  Flight,  involved  "troubleshooting"  for  the  Apollo 
program;  this  basically  amounted  to  staying  in  Washington  to  stroke  the 
troops  during  day-to-day  meetings  at  headquarters  while  the  real  excitement 
took  place  elsewhere.  This  did  not  satisfy  the  ambitious  Cortright.  "It  was 
awfully  hard  to  come  in  that  late  in  Apollo  and  make  much  of  an  impact," 
he  remembers.  So  Cortright  went  directly  to  Webb  and  asked  him  for  the 
Langley  job.*  "I  wanted  my  own  ship.  I  always  felt  that  the  Center  director 
was  absolutely  the  best  job  in  NASA,  and  it  is."  To  him,  filling  Thompson's 
post  would  be  "a  career  dream  achieved."^ 

Webb  gave  the  forty-five-year-old  Cortright  the  job,  apparently  with  little 
hesitation.  In  Cortright's  view,  Webb  had  a  "gut  feeling"  that  "Langley 
needed  a  shot  in  the  arm"  and  that  it  had  gotten  "a  little  bit  sleepy,  a  little 
bit  complacent.  ...  It  needed  to  be  brought  up  to  speed."  That  was  his 
broad  charge  to  him.  Webb  gave  him  no  specifics  about  what  directorate 
or  division  needed  attention;  he  only  conveyed  the  message  that  Cortright, 
as  a  younger  person  and  an  outsider,  might  succeed  in  "putting  more  life" 
into  the  historic  "Mother  Langley."5 

Langley  employees  were  surprised  but  not  shocked  upon  hearing  the 
news  of  Cortright's  appointment.  "My  reception  was  very  cordial,  very 
courteous.  Well,  I  say  very  cordial.  It  was  certainly  friendly  but  maybe  a 
little  standoffish  because  they  didn't  know  me.  I  didn't  know  them  and  I  was 
an  unknown  factor  in  the  equation  to  some  extent."6  The  only  ones  upset  by 
Cortright's  arrival  were  Charles  Donlan's  many  close  friends  and  associates, 
who  believed  the  job  should  have  been  his,  but  not  even  they  openly  showed 


This  is  Cortright's  version.  People  at  Langley  heard  at  the  time  that  Webb,  when  he  was  about  to 
leave  NASA,  asked  Cortright  what  he  could  do  for  him.  Cortright  answered  he  wanted  Langley.  There 
are  no  authoritative  sources  to  contradict  Cortright's  version,  only  hearsay. 

397 


Space/light  Revolution 


L-68-2747 

The  old  guard  welcomed  Cortright  to  Langley  politely  but  without  real  enthusiasm. 
From  left  to  right:  Laurence  K.  Loftin,  Jr.,  Cortright,  Charles  Donlan,  T.  Melvin 
Butler,  Clifford  Nelson,  Eugene  Draley,  and  John  Duberg. 


their  disappointment.  Most  people  gave  Cortright  the  benefit  of  the  doubt. 
What  little  they  knew  about  him  and  his  reputation  at  NASA  headquarters 
gave  them  reason  to  believe  he  was  a  dynamic  leader  and  high  achiever  who 
might  be  successful  in  getting  Langley  more  resources  and  more  of  NASA's 
project  work.  But  as  we  shall  see,  the  latter  was  not  something  that  the 
research-minded  at  Langley  wanted  to  see  happen. 

When  Cortright  arrived,  he  had  no  detailed  master  plan  for  restructuring 
Langley;  thus,  the  reorganization  that  followed  did  not  happen  overnight. 
Cortright  started  his  job  as  Langley  center  director  on  1  May  1968,  but 
he  did  not  have  all  the  major  organizational  changes  formulated  and  ready 
to  be  announced  publicly  until  September  1970.  That  is  not  to  say  that 
he  did  not  hit  the  ground  running.  A  dynamic  man  with  tremendous  self- 
confidence,  Cortright  believed  that  his  project  management  experience  at 
NASA  headquarters  had  prepared  him  to  take  over  Langley's  direction.  On 
his  first  day  at  the  center,  he  met  with  the  senior  staff  and  the  mid-level 
supervisors  in  the  Langley  Morale  Activities  building  to  discuss  his  plans 
for  becoming  more  acquainted  with  the  many  aspects  of  his  new  position. 
In  his  speech  to  them,  he  praised  the  center  for  its  reputation  as  a  "well 

398 


The  Cortright  Synthesis 


On  his  first  day  as  director  (Wednesday, 
1  May  1968),  Edgar  M.  Cortright  (left) 
met  with  the  senior  staff  in  the  Morale 
Activities  building  (below).  Many  mem- 
bers of  the  senior  staff  believed  that 
Cortright  had  arrived  at  Langley  with 
preconceived  ideas  and  plans  for  the 
center. 


L-70-6003 


399 


Space/light  Revolution 

managed  and  smoothly  operating  organization"  and  expressed  "pleasure  at 
the  opportunity  to  work  with  the  Langley  staff  in  moving  forward  as  a  team 
to  continue  to  contribute  to  the  advancement  of  flight."7  Because  he  was 
an  outsider  and  largely  unknown  to  most  of  the  Langley  staff,  Cortright 
did  what  he  could  to  reassure  people  that  he  had  not  come  to  Langley  to 
make  wholesale  changes,  but  suspicions  remained.  Several  members  of  the 
senior  staff,  in  particular,  worried  that  Cortright  had  been  sent  to  Langley 
for  the  express  purpose  of  bringing  Langley  into  line  under  headquarters — a 
sentiment  not  far  from  the  truth. 

For  at  least  the  first  year,  however,  no  one  at  Langley  could  be  sure  what 
Cortright  intended.  During  his  first  six  months,  Cortright  spent  two  or  three 
days  a  week  in  meetings  in  which  every  branch  of  the  center  reported  on 
the  status  of  all  its  facilities  and  activities.  "I  had  every  group  at  the  center 
come  in  and  brief  me  in  depth  on  what  they  were  doing  so  I  wouldn't  go  off 
half-cocked.  I  think  I  got  grudging  respect  out  of  that,  for  taking  the  time 
to  learn  that  much  about  what  was  going  on."  Prom  this  thorough  review, 
Cortright  formed  a  better  picture  of  Langley's  strengths  and  weaknesses. 
"What  I  found  was  a  high  degree  of  technical  competence.  There  was  hardly 
anything  in  aeronautics  that  Langley  was  not  expert  in  and  didn't  have  at 
least  a  small  pool  of  experts."  In  aeronautics  Langley  "didn't  have  to  take 
second  place  to  anyone  in  the  country  in  almost  any  area  one  could  think  of." 
In  space,  however,  Langley  "didn't  have  quite  that  much  capability."  The 
staff  had  good  project  management  capability  from  Lunar  Orbiter  and  from 
Scout,  but  there  were  "quite  a  few  areas  of  space  technology  that  they  had 
very  little  capability  in,"  such  as  electronics  (Langley  did  have  electronics 
under  IRD),  which  Cortright  knew  was  not  Langley's  fault  because  NASA 
had  not  given  the  research  center  much  opportunity  to  develop  these  areas.8 

Overall,  Cortright  felt  that  the  organization  "could  stand  improvement 
in  terms  of  the  way  the  authorities  were  parceled  out"  inside  the  center  and 
in  the  way  Langley  interacted  with  NASA  headquarters.  On  both  counts, 
he  was  a  critic  of  the  old  NACA  way  of  doing  things.  The  NACA,  in  his 
view,  was  "a  very  loosely  structured  research  organization  where  the  power 
resided  in  the  centers,  with  the  center  directors  and  the  key  researchers." 
NACA  headquarters  was  "primarily  a  bookkeeping  type  of  organization.  .  .  . 
They  were  super  people  at  NACA  headquarters,  starting  with  Hugh  Dryden, 
but  mostly  they  went  in  and  fought  for  authorizations  and  appropriations." 
They  did  not  control  the  centers  much,  which  meant  that  the  centers  usually 
got  what  the  centers  wanted.  "What  the  centers  wanted  to  work  on,  they 
worked  on."9  In  other  words,  the  engineers  not  the  managers  were  in  charge. 

Under  the  demands  of  the  spaceflight  revolution,  NASA  had  worked  to 
invent  a  new  system.  "We  did  not  delegate  total  management  authority  to 
the  centers.  We  kept  a  rather  strong  management  team  in  NASA  headquar- 
ters." Every  project  had  "program  managers"  in  headquarters,  along  with 
a  designated  project  manager  at  the  centers.  "When  it  worked  right,  those 
two  guys  worked  hand  in  glove.  When  it  didn't  work  right,  frequently  the 

400 


The  Cortright  Synthesis 

center  directors  got  in  the  middle  and  resisted  direct  independent  interaction 
between  the  program  office  and  the  project  office."  Floyd  Thompson  had 
done  this  more  than  occasionally,  which  was  one  of  the  main  reasons  that 
Cortright  was  given  Thompson's  job.  And  he  was  just  the  man  to  correct 
the  situation.  While  working  under  Homer  Newell,  he  had  written  the  book 
on  the  relationship  between  the  program  manager  and  the  project  manager. 
Cortright  now  had  the  responsibility  of  leading  Langley  into  a  new  age  in 
which  it  would  live  by  the  book  much  more  closely  than  it  ever  had  under 
Thompson,  who  in  Cortright 's  view  was  "not  overly  responsive"  to  following 
set  procedures.10 

Cortright  also  wanted  to  keep  closer  tabs  on  how  research  was  being 
managed  at  Langley.  In  the  old  days  of  the  NACA,  research  was  allowed 
to  take  its  time  and  run  its  course,  which  meant  that  many  efforts  were 
allowed  to  go  on  indefinitely  awaiting  clear  results.  But  times  had  changed, 
and  the  uncertainty  of  such  tortoise-paced  fundamental  explorations  could 
not  be  tolerated  as  in  the  past.  About  a  year  after  coming  to  the  center, 
one  meeting  in  particular  cemented  Cortright 's  opinion  of  what  needed  to 
be  done  to  invigorate  the  center.  For  this  meeting,  he  asked  the  directors 
of  all  the  divisions  to  put  on  the  briefing  and  explain  what  their  programs 
involved  and  where  the  most  progress  could  be  made. 

That  meeting,  Cortright  remembers,  was  a  "disaster."  First,  he  discov- 
ered to  his  dismay  that  some  of  Langley 's  research  divisions  were  working 
on  technology  that  was  "95  percent  soft,"  meaning  that  the  researchers  in- 
volved were  "squeezing  the  last  ounce  out  of  it  at  a  fair  expenditure  of  time 
and  money."  More  disturbing  was  that  "it  was  very  hard  to  get  anyone  to 
tell  you  what  his  program  was  under  him.  It  was  just  a  collection  of  mis- 
cellaneous things  which  he,  the  individual,  may  have  known  how  to  justify 
informally  on  a  one-to-one  basis,  but  was  not  very  articulate  at  explaining." 
This  poor  performance  fixed  Cortright 's  feeling  that  Langley  was  indeed  a 
sleepy  place  in  need  of  new  lifeblood.  "I  was  very  disappointed  because  the 
directors  hadn't  done  their  homework  and  gave  some  very  bad  presentations. 
And  that  set  my  mind  at  work  that  there  should  be  some  changes."11 


The  Reorganization 

One  of  the  goals  of  the  major  reorganization  of  the  center  that  ensued  was 
to  make  responsibilities  and  lines  of  communication  clearer  within  Langley. 
According  to  the  NACA  way,  such  things  were  kept  informal  and  rather 
obscure.  Dr.  George  Lewis,  the  NACA's  longtime  director  of  research 
(1919-1947),  disliked  formal  organization  charts,  once  saying  to  a  young 
Langley  industrial  engineer  interested  in  developing  them  that  the  only 
thing  that  boxes  were  good  for  was  burying  people.  Henry  Reid  and  Floyd 
Thompson  agreed  wholeheartedly  with  Lewis.  The  best  way  to  set  up  an 
organization  was  to  make  things  very  fluid  so  that  "shadow  organizations" 

401 


Spaceflight  Revolution 

could  cut  horizontally  through  the  formal  boundaries.  Such  an  organization 
also  made  it  hard  for  outsiders  to  know  exactly  what  was  going  on  inside 
the  laboratory,  thereby  making  it  harder  for  them  to  exploit  or  manipulate 
its  resources.  Such  thinking  resulted  in  Floyd  Thompson's  creation  of  the 
Groups  1,  2,  and  3  organization,  which  made  it  impossible  to  tell  the  players 
inside  them  without  a  scorecard. 

Ed  Cortright  wanted  none  of  this.  In  his  own  words,  he  was  always 
"180  degrees  opposite"  from  Thompson's  shell  game.  Instead,  he  thought 
that  a  center  director  should  "show  everything  and  everyone,  right  up  to 
the  administrator,  exactly  how  you  were  organized  and  everything  you  were 
doing."  In  other  words,  "total  display"  was  his  policy.  Organization  charts 
should  be  simple  and  obviously  understandable,  with  "a  good  person  in  every 
box."  Reporting  procedures  should  be  equally  clear  so  that  no  surprises 
lurked  in  the  wings.  "If  we  were  screwing  up  someplace,  we'd  learn  it  as 
we  were  going  along — not  at  the  end  of  the  project  when  something  bad 
has  happened  and  a  [government  auditor]  comes  back  and  says  that  we've 
misperformed."  At  NASA  headquarters  earlier  in  the  1960s,  Cortright  had 
set  up  a  reporting  system  for  budgets  and  projects  that  guaranteed  such 
total  display.  With  the  same  bureaucratic  purposes  in  mind,  Cortright  now 
set  out  to  reorganize  Langley  and  its  reporting  procedures.  "We  spent  a 
lot  of  time  putting  together  our  presentations  every  year,  of  every  project, 
every  program  area,  with  pictures  and  charts  to  show  exactly  where  the 
money  was  going,  how  it  was  being  spent."12  Not  everyone  at  Langley  liked 
such  visibility,  and  some  may  not  have  abided  by  it  faithfully,  but  by  the 
early  1970s,  it  had  become  standard  at  the  center.  Those  people  who  did 
not  like  it  "lumped  it." 

By  the  end  of  1969,  Cortright  had  bided  his  time  long  enough;  he  was 
ready  to  carry  out  the  reorganization  that  had  been  germinating  in  his  mind 
since  he  arrived.  In  typical  Cortright  fashion,  he  brought  those  who  would 
be  most  affected  by  the  changes  into  the  process  of  making  them  so  that 
they  would  feel  that  they  had  at  least  a  part  in  determining  the  course  of 
their  own  future.  "I  gave  the  problem  to  my  directors,"  whether  they  liked 
it  or  not.  He  said  to  them,  "Look,  this  is  roughly  what  I  want,  and  I'm 
not  going  to  tell  you  how  to  do  it."  The  only  thing  that  he  specified  was 
that  he  did  want  basic  changes,  because  "you  get  comfortable  with  your 
mistakes,"  and  he  wanted  opportunities  made  "for  creative  young  people  to 
get  a  chance  to  lead."13 

In  the  director's  main  conference  room — the  "Brown  Room"  in  the  Ad- 
ministration building,  which  was  next  to  Cortright 's  own  office — a  small  task 
force  made  up  of  the  directors  and  a  few  handpicked  staff  assistants,  them- 
selves mostly  young,  orchestrated  the  reorganization  that  the  spaceflight 
revolution  had  made  inevitable.  On  a  large  mahogany  conference  table,  this 
task  force  laid  out  big  sheets  of  paper  and  went  to  work  reorganizing  the 
center  and  everything  and  everyone  who  belonged  to  it.  As  the  directors 
brought  in  their  ideas  and  plans,  the  task  force  coordinated  and  massaged 

402 


The  Cortright  Synthesis 

them,  keeping  Cortright 's  agenda  foremost  in  mind.  When  it  was  all  done, 
in  the  summer  of  1970,  between  80  and  90  percent  of  all  the  supervisory 
positions  at  Langley  had  new  faces  in  them.  A  few  of  the  old  supervisors 
left  NASA  through  early  retirements;  some  did  not  retire  but  became  con- 
sultants on  salary  out  of  the  line  of  command;  most  were  shuffled  into  new 
positions  and  into  new  organizations.  Only  the  people  who  really  did  not 
like  where  this  rearranging  landed  them  left. 

On  24  September  1970,  after  months  of  planning,  Cortright's  office  dis- 
tributed a  large  79-page  report  to  all  center  employees.  It  was  entitled 
"Reorganization  of  Langley  Research  Center."  The  report  detailed  all  the 
organizational  changes  and  provided  charts  outlining  the  new  overall  struc- 
ture of  the  center.  On  its  green-colored  cover,  Cortright  explained  the 
report's  purpose  and  immediate  history:  "Several  weeks  ago  I  submitted 
proposals  for  the  reorganization  of  Langley  Research  Center  to  the  Admin- 
istrator [Dr.  Thomas  O.  Paine]  and  the  Office  of  Advanced  Research  and 
Technology  for  review  and  consideration.  These  proposals  have  now  been 
approved  by  Headquarters,  and  the  organizational  and  personnel  changes 
will  be  made  effective  October  4,  1970."^  Whether  Cortright  was  aware  of 
it  or  not — and  no  evidence  exists  to  indicate  that  he  was — 4  October  was 
the  13th  anniversary  of  Sputnik.  To  Cortright,  the  day  was  just  the  start 
of  a  new  fiscal  year.  (Coincidentally,  or  perhaps  as  part  of  some  underlying 
pattern  in  the  unfolding  of  revolutions,  the  birth  of  the  Napoleonic  Code 
in  1802  dates  from  almost  exactly  13  years  after  the  outbreak  of  the  French 
Revolution.) 

In  his  letter  to  Administrator  Paine,  the  contents  of  which  he  reviewed 
in  the  "Green  Paper"  report,  Cortright  stated  that  "the  proposed  reorga- 
nization would  accomplish  a  number  of  important  steps  toward  effective 
prosecution  of  the  projected  research  program  of  this  Center  during  the 
coming  decade."  He  then  outlined  the  six  principal  advantages  of  the  re- 
alignment: first,  each  of  the  major  areas  of  research  responsibility  at  the 
center  would  have  a  clearer  "organizational  focus";  second,  all  emerging 
R&D  responsibilities  would  be  recognized  organizationally;  third,  new  su- 
pervisory assignments  would  be  tailored  to  the  future  opportunities  for  the 
center;  fourth,  personnel  and  facilities  working  on  closely  related  problems 
would  be  more  efficiently  integrated;  fifth,  the  center  would  strengthen  its 
effort  on  institutional  problems  and  work  more  effectively  with  NASA  head- 
quarters; and  sixth,  technical  and  administrative  support  of  the  research 
effort  would  be  improved.  More  specifically,  the  reorganization  would  result 
in  better  integration  of  programs;  greater  stimulation  of  new  program  efforts 
as  well  as  the  enhanced  development  of  supervisory  personnel;  more  efficient 
use  of  manpower,  equipment,  and  materials;  and  better  interface  between 
the  center  and  other  organizations.15 

The  new  organization  consisted  of  four  major  research  directorates:  Elec- 
tronics, Structures,  Aeronautics,  and  Space,  thus  eliminating  Thompson's 
mysterious  Groups  1,  2,  and  3.  Supporting  these  four  were  two  other 

403 


Space/light  Revolution 

directorates:  Systems  Engineering  and  Operations,  and  Administration.  In 
addition,  one  other  directorate  was  formed  for  center  development.  This 
final  directorate,  first  headed  by  Gene  Draley  and  later  by  Langley  veteran 
T.  Melvin  Butler,*  was  to  report  directly  to  the  center  director  on  matters 
that  encompassed  the  interest  of  all  or  most  of  the  other  directorates,  such 
as  safety,  reliability,  and  quality  assurance. 

Without  a  doubt,  the  overall  layout  was  much  cleaner  and  clearer  than 
any  during  Floyd  Thompson's  term  as  director.  Every  aspect  of  the  center 
that  could  be  rationalized  was  rationalized.  The  reorganization  drew  a 
clear  line  between  aeronautics  and  space.  Aeronautics  was  split  into  three 
basic  research  divisions:  one  for  low-speed  aircraft,  one  for  high-speed 
aircraft,  and  one  for  hypersonic  vehicles,  with  two  additional  divisions 
focusing  on  advanced  technology  transports  (most  notably  the  SST,  which 
was  not  canceled  until  the  following  year)  and  on  the  operation  of  flight 
research  vehicles.  The  Structures  Directorate  also  divided  neatly  into  three 
divisions:  loads,  structures,  and  materials.  Even  after  the  reorganization, 
directors  and  division  heads  still  exercised  some  crossover  responsibilities 
but  not  nearly  as  many  as  before. 

In  addition,  at  Cortright's  specific  request,  the  new  organization  made 
room  for  and  focused  some  of  the  center's  newest  R&D  interests.  Electronics, 
as  Cortright  had  noted  in  his  earliest  staff  briefings,  needed  beefing  up, 
so  it  became  its  own  directorate,  with  four  divisions.  Inside  the  Space 
Directorate,  a  new  division  devoted  to  the  study  of  environmental  and  space 
sciences  was  formed.  This  represented  a  new  focus  of  interest  at  Langley, 
one  stimulated  in  large  part  by  the  growing  international  interest  in  the 
problems  of  environmental  pollution  and  energy  consumption.  Under  the 
direction  of  Cortright,  who  considered  himself  a  pragmatic  environmentalist, 
the  atmospheric  sciences  ultimately  proved  to  be  one  of  the  center's  most 
productive  and  worthwhile  areas  of  research.16 

Another  noteworthy  item  on  the  organization  chart  was  the  Viking 
Project  Office.  The  goal  of  Project  Viking  was  to  land  scientific  payloads 
on  Mars;  this  mission  would  be  achieved  in  1976.  In  December  1968,  just 
a  few  months  after  Cortright  took  office,  NASA  gave  Langley  the  overall 
responsibility  for  managing  the  project.  The  assignment  made  sense  given 
Langley's  management  of  the  tremendously  successful  Lunar  Orbiter  project 
and  Cortright's  own  experience  in  project  management  and  in  the  OSSA, 
where  he  had  been  the  first  to  propose  a  manned  Mars  mission.  In  a  sense, 
Viking  came  to  Langley  with  Cortright,  and  it  went  where  he  did  even  after 
it  came  to  Langley.  As  can  be  seen  in  the  4  October  1970  organization 
chart,  the  Viking  Project  Office  reported  directly  to  Cortright.  No  one 
else,  no  other  box  on  the  chart,  intervened.  Before  the  reorganization,  the 
Viking  office  was  in  the  Office  for  Flight  Projects,  which  meant  an  assistant 


Butler,   30  years  previously,   had  been  the  young  industrial  engineer  whom  NACA  Director  of 
Research  George  Lewis  had  scolded  for  wanting  to  formalize  Langley's  organization  charts. 

404 


The  Cortright  Synthesis 


C 
OF 


QC 
HI 

I- 

LU 
O 

X 

o 

CC 

< 

UJ 
0) 
UJ 

CC 

UJ 

o 

z 


;ae 


i 

§    9 


1  z 

• 

Z 

p  5 

< 

AFFAIRS 

o  uj  O 

1 

I  o 

Z 

si 

CC  C/3 

||| 

|||| 

|l«z 

CC  ""  2 

f* 

y 

«> 

lls 

£l 

§5 

|1° 

^|P 

|p5 

0    1 

o 

2 

\ 

1 

1                1 

FOR 

UCTU 


15 


^ 

gls 

UJ  5  O 
<^« 

III 

:5yjy 

u  o< 

M  I  g 

Q.  O)  > 

w£  o 

0 

" 

1 

1               1               1 

cc      p 

LOW-SPEED 
AIRCRAFT 
DIVISION 

HIGH-SPEED 
AIRCRAFT 
DIVISION 

HYPERSONIC 
VEHICLES 
DIVISION 

RESEARCH 
AIRCRAFT 
FLIGHT 
DIVISION 

Ills 

1 

1 

I 

<z 


DIRECTOR 
FOR 

ELECTRONICS 

FLIGHT 
DYNAMICS  & 
CONTROL 

DIVISION 

ANALYSIS  & 
COMPUTATION 
DIVISION 

INSTRUMENT 
RESEARCH 
DIVISION 

FLIGHT 
INSTRUMENTATION 
DIVISION 

\      \      \ 

§ 

ON 


a 


405 


Space/light  Revolution 


L-70-7351 

Reaching  Mars  with  just  an  unmanned  spacecraft  meant  overcoming  engineering 
challenges  greater  in  some  respects  than  those  posed  by  Apollo.  Above,  a  scale 
model  of  the  Viking  Lander  is  tested  in  a  Langley  wind  tunnel  in  1970. 


director  intervened  between  Viking  managers  and  Cortright.  But  Viking  was 
Cortright's  baby,  and  with  the  reorganization  he  made  sure  it  was  pampered. 
As  the  Green  Paper  makes  clear,  "The  Viking  Project  Office  reports  to  the 
Center  Director  and  thus  occupies  a  unique  position  at  the  Center.  This 
change  recognizes  the  fact  that  Viking  has  the  highest  priority  of  any  single 
undertaking  at  the  Center  and  that  authority  has  been  delegated  to  the 
Viking  Project  Manager  [James  Martin]  to  exercise  necessary  direction  and 
control  over  Center  personnel  who  have  been  committed  to  the  support  of 
Viking  within  the  research  and  engineering  divisions."  Such  was  not  the  case 
for  the  only  other  major  project  office  at  Langley  in  1970 — Scout — which 
was  still  within  a  larger  directorate.17 

This  clear  preference  for  Project  Viking  disturbed  several  Langley  re- 
searchers who  feared  that  the  big  and  virtually  independent  project  office 
would  dominate  the  general  research  programs.  They  worried  that  Viking 
would  influence  the  center  to  a  greater  extent  than  Project  Apollo  had  in 
the  1960s.  Some  feared  a  "Viking  raid,"  and  justifiably  so. 

Cortright  was  enthusiastically  committed  to  Viking;  by  his  own  admis- 
sion, it  "probably  took  half  my  time  as  director."  With  his  help,  the  project 
office  built  up  a  full-time  staff  of  more  than  250.  Supporting  them  were  an 
equal  number  of  employees  in  systems  engineering,  the  shops,  and  in  admin- 
istration. This  did  cause  real  problems  at  Langley.  "The  researchers  felt  I 
gave  too  much  of  my  time  to  Viking  and  that  it  got  too  many  resources. 
Some  even  felt  that  it  split  the  center.  But  my  philosophy  was  simple.  We 

406 


The  Cortright  Synthesis 


Left,  a  Viking  Orbiter  1  photo  of  the  mys- 
terious "Red  Planet"  taken  in  June  1976. 
In  September  1976,  the  camera  aboard 
Viking  Lander  2  (below)  shows  a  boulder- 
strewn  field  reaching  to  the  Martian  hori- 
zon approximately  2  miles  away. 


407 


Space/light  Revolution 

had  Viking.  We  had  to  make  it  succeed,  period.  There  was  no  choice  but 
to  do  it  properly.  And  it  did  succeed,  magnificently.  It  was  the  highlight  of 
the  whole  space  program,  as  far  as  I  am  concerned.  Two  orbiters  around 
Mars;  two  landers  on  Mars.  They  all  operated  the  way  they  were  supposed 
to  do.  I  never  would  have  expected,  in  my  fondest  dreams,  that  they  would 
have  been  that  successful.  ...  If  anyone  ever  thought  [Langley]  was  a  sleepy 
center,  they  sure  didn't  while  Viking  was  going  on  or  when  it  succeeded." 18 

Viking  concerned  researchers,  but  nothing  upset  them  as  much  as  the 
personal  trauma  of  leaving  old  and  comfortable  jobs,  moving  into  new 
and  unknown  ones,  and  perhaps  even  facing  retirement.  In  concluding  his 
executive  summary  of  the  contents  of  the  reorganization  report,  Cortright 
stated:  "It  is  realized  that  the  changes  outlined  herein  affect  a  large 
number  of  employees,  many  of  whom  will  be  relocated  and  working  under 
new  supervision.  During  the  initial  transition  stages,  some  inconveniences 
will  be  experienced.  This  is  to  be  expected  and  is  a  normal  outgrowth 
of  reorganization.  However,  I  believe  that  the  immediate  and  long-range 
future  of  Langley  Research  Center  is  of  prime  concern  to  all  of  us,  and  I 
feel  confident  that  all  supervisors  and  employees  will  make  every  effort  to 
facilitate  the  transition."19 

Any  executive  officer  wants  to  put  together  his  own  team  of  subordinates, 
and  Cortright  was  certainly  no  exception.  Long  before  the  reorganization,  he 
had  made  it  clear  that  he  would  clean  house.  For  example,  when  Donlan  left 
Langley  for  Washington  in  early  1968  after  being  bypassed,  Floyd  Thompson 
had  chosen  51-year-old  Dr.  John  E.  Duberg,  a  structures  expert  and  Langley 
veteran  since  1943,  as  the  acting  associate  director.  Thompson  believed  that 
Duberg  should  stay  on  as  Cortright's  associate  director  because  he  was  just 
the  person  who  could  help  a  stranger  learn  everything  he  needed  to  know 
about  Langley.  Unfortunately  for  Duberg,  Cortright  came  to  realize  that 
Duberg  "liked  the  center  the  way  it  was,"  which  marked  him  in  Cortright's 
mind  as  "inflexible."20  Cortright  eventually  took  him  out  of  the  post  and 
made  him  chief  scientist,  a  position  in  which  Duberg  performed  well.* 

In  truth,  Duberg  just  was  not  the  kind  of  number  two  man  Cortright 
wanted.  He  was  looking  for  "a  hard-nose  driver  who  will  implement  and 
follow  through."  In  Cortright's  view,  his  own  strongest  suit  had  always  been 
"understanding  what  needs  to  be  done,  being  able  to  put  teams  together  to 
do  it,  and  providing  the  leadership,  with  the  correct  policies  and  decisions." 
He  was  not  terribly  good,  however,  at  stepping  on  toes  or  being  ruthless 
about  getting  things  done;  he  was  "too  soft  for  that."  For  the  "tough 
follow-through,"  he  needed  someone  else — a  "hard-nose"  deputy.21 

As  part  of  the  September  1970  reorganization,  Cortright  found  his  man, 
Oran  W.  Nicks,  the  acting  associate  administrator  of  the  OART  at  NASA 


Before  becoming  chief  scientist,  Duberg  served  a  short  time  as  acting  director  for  center  development, 
succeeding  Eugene  Draley,  who  wanted  no  part  of  his  new  job  under  the  Cortright  reorganization  but 
stayed  on  until  he  died  of  cancer. 

408 


The  Cortright  Synthesis 


L-70-176 

John  E.  Duberg  (left)  became  Langley's  chief  scientist  after  serving  briefly  as 
the  associate  director  for  Edgar  M.  Cortright  (right).  The  man  in  the  middle 
is  George  M.  Low,  former  head  of  the  Office  of  Manned  Spaceflight  at  NASA 
headquarters.  Low,  at  the  time  of  this  photo,  was  serving  as  acting  NASA 
administrator  following  Thomas  O.  Paine' s  resignation  in  September  1970. 


headquarters.  His  choice  was  not  a  popular  one  at  Langley,  to  put  it  mildly, 
in  large  part  because  Langley  employees  were  just  not  used  to  having  such 
a  gunslinging  honcho  around.  Thompson  had  been  a  tough  number  two 
for  Henry  Reid,  and  Donlan  had  been  tough  enough  for  Thompson,  but  no 
one  ever  cut  quite  as  sharply  into  the  flesh  of  employees  as  Nicks.  This 
was  especially  hard  to  take  from  someone  who  did  not  know  Langley.  Like 
Cortright,  Nicks  had  spent  all  his  time  up  to  then  at  NASA  headquarters 
and  mostly  in  the  management  of  the  space  sciences  arena,  and  unlike 
Cortright,  he  did  not  have  the  time  to  acquaint  himself  with  the  center,  or 
vice-versa,  before  launching  into  action.  Even  Cortright  admits  in  retrospect 
that  "people  either  loved  him  or  hated  him,  and  there  wasn't  too  much  in 
between.  He  did  step  on  toes.  He  was  impatient  with  inaction,  and  research 
centers  don't  normally  leap  into  action  every  time  you  suggest  something."22 
The  person  who  felt  the  bite  of  the  reorganization  and  Cortright 's  hard- 
nose  deputy  the  most  was  Laurence  K.  Loftin,  Jr.,  the  sage  head  of  Group  3 
and  John  Stack's  successor  as  head  of  aeronautical  research  at  the  center. 
Loftin  survived  the  reorganization  and  was  made  director  of  the  Office  of 

409 


Space/light  Revolution 


VOLUME  9,  NUMBER  20 


LANGLEY    RESEARCH    CENTER 


OCTOBER  2,  ly70 


C.  H.  Nelson 


G.  B.  Graves 


G.  W.  Brooks 


P.  J.  Grain  T.  M.  Butler  E.  C.  Draley 

AWARDS  CEREMONY  OCT.    14 


AWARDS  CEREMONY  SPEAKER:  Neil  Armstrong,  Com- 
mander of  the  Apollo  1 1  moon  mission,  will  be  guest  speak- 
er at  the  Center's  Annual  Awards  Ceremony  which  will  be 
held  October  14  at  1:30  p.m.  in  the  NASA  Hangar,  Building 
1244.  Each  year  the  Center  sets  aside  a  special  day  in  Oct. 
for  purposes  of  presenting  length  of  service  awards  to  em- 
ployees and  of  recognizing  individuals  and  groups  at  Lang- 
ley  for  outstanding  achievements  which  they  have  made  to- 
ward realizing  the  mission  of  the  Center  and  of  NASA. 


MAJOR  ORGANIZATIONAL  CHANGES 
ANNOUNCED;  EFFECTIVE  OCTOBER  4 

Major  changes  in  the  Langley  Research  Center's  organiza- 
tion, which  will  become  effective  October  4,  have  been  an- 
nounced by  Director  Edgar  M.  Cortright. 

"This reorganization,  which  is  keyed  to  the  future,"  Cort- 
right said,  "is  designed  with  several  important  objectives 
in  mind.  We  are  consolidating  our  research  talents  and 
focusingthem  on  NASA's  roles  and  missions  for  the  1970's. 
We  are  giving  increased  attention  to  the  future  growth  of  the 
Langley  Research  Center  and  to  its  relationships  with  other 
organizations.  And  we  are  moving  some  of  our  most  prom- 
ising young  men  into  positions  of  management  responsibility 
where  they  can  exercise  research  leadership." 

Four  major  research  directorates  are  being  established 
in  Aeronautics,  Space,  Electronics,  and  Structures.  They 
will  be  headed  by  Laurence  K.  Loftin,  Jr.,  Director  for 
Aeronautics;  Clifford  H.  Nelson,  Director  for  Space;  George 

B.  Graves,    Jr.,    Director  for  Electronics;  and  Dr.  George 
W.  Brooks,  Director  for  Structures. 

The  four  research  directorates  will  be  supported  by  three 
Center-wide  directorates  in  the  areas  of  Systems  Engineer- 
ing and  Operation,  Administration,  and  Center  Development. 
Percy  J.  Crain  will  be  Director  for  Systems  Engineering; 
T.  Melvin  Butler,  Director  for  Administration;  and  Eugene 

C.  Draley,  Director  for  Center  Development. 

Oran W.Nicks,  Associate  Administrator  for  Advanced  Re- 
search and  Technology  (Acting),  NASA  Headquarters,  will 
join  the  Center  as  Deputy  Director  in  early  November. 

In  a  meeting  with  Langley  senior  staff  officials,  Cortright 
outlined  some  of  the  future  NASA  goals  with  which  the  Cent- 
er is  closely  identified. 

These  include  an  expanded  aeronautics  program  with  in- 
creased emphasis  on  civil  aircraft;  a  new  era  of  manned 

(Continued  on  page  5) 


Cortright  unveils  his  sweeping  reorganization. 


410 


The  Cortright  Synthesis 

Aeronautics,  but  he  did  not  last  long.  Loft  in  and  Cortright  represented 
different  philosophies  about  research  and  how  it  could  best  be  accomplished. 
Loftin  had  gone  along  grudgingly  with  the  spaceflight  revolution  and  had 
played  a  prominent  role  in  both  the  early  planning  for  Apollo  and  the 
space  station.  In  this  regard,  he  was  typical  of  many  of  the  former  NACA 
researchers,  but  his  heart  had  always  remained  true  to  airplanes  and  the 
golden  years  of  NACA  aerodynamic  research.  In  the  twilight  of  his  career, 
he  was  not  inclined  to  bow  to  Cortright 's  plans  for  Langley — and  he  was 
even  less  inclined  to  tolerate  Oran  Nicks,  whose  technical  abilities  he  did 
not  respect. 

In  June  1971,  Cortright  replaced  Loftin,  who  moved  to  a  special  as- 
signment in  the  Office  of  the  Assistant  Secretary  for  the  U.S.  Air  Force.23 
Although  the  outward  appearance  of  Loftin's  job  change  may  have  looked 
normal,  Loftin  in  fact  felt  he  had  been  fired.  Cortright  exacerbated  the 
problem  at  the  center  when  he  gave  Loftin's  old  job  temporarily  to  Nicks. 
This  lasted  for  only  three  months,  until  September,  when  Robert  Bower, 
director  of  advanced  development  for  Grumman  Aerospace  Corporation  in 
Bethpage,  New  York  (and  Cortright's  college  fraternity  brother),  replaced 
him.24  Unfortunately,  Bower  proved  no  more  popular  at  Langley  than  Nicks. 

Part  of  what  inspired  the  reorganization  was  Cortright's  desire  for  a 
youth  movement.  "I  wanted  more  youth,"  Cortright  remembers.  "I  wanted 
people  who  had  been  in  the  same  job  for  twenty  years  to  move  and  do 
something  a  little  different.  .  .  .  That's  what  I  felt  my  charter  was  and  I 
felt  after  that  first  year  of  studying  the  center  and  its  people  that  it  was 
justified."  During  Cortright's  apprenticeship  at  NASA  in  the  early  1960s, 
Abe  Silverstein  had  told  him,  "You've  got  to  keep  the  new  blood  coming  in 
at  the  bottom  all  the  time.  If  not,  you  have  big  gaps  in  your  management 
later  on."  Cortright  launched  an  aggressive  recruitment  campaign  at  the 
same  time  he  was  encouraging  some  of  the  older  employees  to  retire  early. 
He  did  this  in  the  face  of  rather  significant  manpower  reductions  NASA- 
wide.  Even  the  heads  of  the  directorates  had  to  go  out  and  give  talks  at 
some  of  the  major  colleges  and  universities  to  recruit  talented  graduates. 
Cortright  himself  went  to  Rensselaer,  his  alma  mater,  and  to  the  California 
Institute  of  Technology  (Caltech),  which  he  knew  well  from  his  days  at 
NASA  headquarters  when  he  monitored  studies  being  done  at  JPL.25 

Cortright  changed  Langley's  recruitment  policy.  He  found  by  asking  the 
center's  personnel  department  that  Langley  had  been  recruiting  in  the  1960s 
almost  exclusively  from  a  handful  of  schools,  notably  Virginia  Polytechnic 
Institute  and  State  University,  North  Carolina  State  University,  and  Georgia 
Institute  of  Technology.  When  he  asked  why  that  was  the  case,  personnel 
officer  E.  Townsend  Johnson,  Jr.,  replied,  "Well,  Mr.  Cortright,  we're  sort 
of  partial  to  southern  boys."  That  displeased  Cortright,  and  rightfully  so, 
given  the  extent  to  which  NACA  Langley  in  the  early  years  had  built  its 
talented  staff  on  young  people  from  northeastern  and  midwestern  schools. 


411 


Spaceflight  Revolution 

Cortright  insisted  on  a  nationwide  search  from  then  until  the  time  he  left 
in  1975.26 

Cortright 's  youth  movement  was  not  designed  just  to  fill  slots  at  the 
bottom  of  the  organization;  he  also  wanted  to  give  younger  people  the 
chance  to  take  on  additional  responsibility  and  assume  leadership  positions. 
To  make  room  for  the  new  people,  however,  Langley  needed  to  get  rid  of 
some  of  the  old.  As  suggested  earlier,  Cortright  asked  for  early  retirements. 
A  number  of  senior  people  felt  that  if  anyone  had  to  go  because  of  the 
manpower  reductions,  it  ought  to  be  them  because  they  already  had  enjoyed 
good  careers  and  should  give  younger  people  a  chance.  These  younger  people 
moving  up,  of  course,  were  happy  with  the  reorganization  and  provided 
Cortright  with  a  solid  base  of  support  lasting  throughout  his  stay  at  Langley. 
Other  senior  employees,  however,  became  dissatisfied  because  of  Cortright's 
downscaling  of  many  of  the  highest  civil  service  salaries  at  the  center. 
Instead  of  trying  to  get  as  much  as  he  could  for  his  employees,  as  Thompson 
had  always  done,  Cortright  felt  that  researchers,  no  matter  how  many  years 
of  service,  should  not  make  more  money  than  the  boss. 

One  group  that  strongly  supported  Cortright  was  the  staff  toiling  in  the 
shops.  For  them,  Cortright  rehabilitated  facilities  to  provide  better  working 
conditions  and,  most  importantly,  air  conditioning  for  the  hot  and  humid 
Tidewater  summers.  In  large  buildings  like  the  huge  West  Area  airplane 
hangar  where  air  conditioning  was  impractical,  he  found  enough  money  to 
install  large  fans  that  hung  down  from  the  ceiling  and  blew  cool  air  on  the 
mechanics  and  other  workers.  "Boy,  these  shop  people  were  my  friends  from 
then  on,"  Cortright  is  proud  to  say.  "And  when  I  retired,  they  wrote  in 
their  shop  news  that  they  thought  it  was  unfortunate  that  the  best  director 
they'd  ever  had  had  seen  fit  to  retire.  ...  I  loved  those  folks.  They're  the 
salt  of  the  earth.  And  I  wasn't  going  to  have  them  working  in  sweat-shop 
conditions  while  all  the  engineers  sat  in  air  conditioning."27 

Cortright  did  other  things  to  build  support  for  his  programs.  He  initiated 
the  practice  of  giving  an  annual  "state  of  the  center"  address,  an  affair  that 
had  no  precedent  at  Langley.  "I  was  a  big  one  for  getting  people  on  my  side 
and  pumped  up,"  he  states.  "Langley  did  so  many  great  things  before  I  got 
there  under  Reid  and  Thompson,  that  I  just  wanted  to  get  more  of  the  same. 
I  didn't  particularly  want  to  do  any  better  than  they  did,  just  to  make  sure 
that  I  did  that  well,  at  least."  Besides  giving  this  formal  overview  of  the 
center's  current  programs  and  future  directions  to  the  Langley  employees, 
Cortright  also  presented  a  simplified  version  of  his  state  of  the  center  address 
to  the  employees'  spouses.  "I  gave  them  about  an  hour  and  a  half  lecture  on 
what  the  center  was  doing,  why  their  husbands  were  coming  home  late,  and 
what  they  were  accomplishing.  That  'coffee  and  briefing  program'  worked 
out  real  well.  It  was  just  part  of  my  style."28 

Cortright  also  moved  to  stimulate  NASA's  popular  appeal  in  the  com- 
munity at  large  by  creating  an  official  NASA  Visitors'  Center,  which  opened 
on  site  at  the  center  in  1971.  Such  an  attraction  contrasted  sharply  with 

412 


The  Cortright  Synthesis 

the  old  NACA  days  prior  to  the  spaceflight  revolution  when  the  low-key 
Langley  installation  was  largely  invisible  to  the  American  public.  The  lead- 
ers of  the  NACA  wanted  to  keep  it  that  way,  but  Cortright,  following  his 
philosophy  of  total  display,  wanted  the  center  opened  to  visitors — as  many 
of  them  as  would  come.  "I  built  the  Visitors'  Center.  I  initiated  a  major 
program  to  communicate  the  good  works  of  the  lab.  I  got  headquarters' 
approval  for  it  and  we  designed  it  internally,  put  up  all  the  displays,  and 
opened  it  up  in  less  than  two  years'  time.  It  became  the  place  for  having  out- 
side groups  and  hosting  receptions."  (Axel  Mattson,  Langley's  former  secret 
agent  in  Houston,  actually  put  the  visitors'  center  together,  along  with  many 
of  its  exhibits.)  In  fact,  until  its  official  closing  in  1992  and  its  move  to  the 
multimillion-dollar  Virginia  Air  and  Space  Center  on  the  riverfront  in  down- 
town Hampton,  this  small  visitors'  center  on  site  at  Langley  Research  Cen- 
ter attracted  (at  least  according  to  NASA  statistics)  more  tourists  than  any 
other  single  attraction  on  the  Peninsula,  which  includes  the  many  historic 
and  entertaining  sites  in  the  Jamestown- Williamsburg-Yorktown  triangle.29 
In  a  similar  vein,  Cortright  brought  Langley  and  its  surrounding  com- 
munity together  by  joining  with  the  city  of  Hampton  in  the  building  of  a 
unique  steam-generating  trash  burner.  "We  saw  where  we  could  burn  trash 
and  get  all  of  Langley  Field's  steam  requirement  and  help  solve  the  landfill 
problem."  Put  into  operation  in  1978,  this  unique  facility  worked  well  except 
for  the  release  of  some  nitrogen  oxides  into  the  air,  a  problem  that  the  plant 
engineers  eventually  solved.  The  incinerator,  which  remains  in  operation  to 
this  day,  is  a  testimony  to  the  kind  of  community  program  that  Cortright 
sponsored.30 


New  Directions 


Ed  Cortright  was  not  an  unpopular  center  director.  Nor  was  he  unac- 
complished. During  his  seven  years  at  the  center,  Langley  again  built  a  solid 
record  of  R&D  achievements.*  Topping  the  list,  of  course,  were  the  success- 
ful landings  of  the  Viking  spacecraft,  Cortright 's  pride  and  joy,  on  Mars 
in  1976.  Nothing  in  the  history  of  space  exploration,  except  perhaps  the 
manned  lunar  landing,  was  a  more  amazing  feat  or  more  difficult  to  carry 
out.  Under  Langley's  expert  project  management  and  Cortright's  general 
oversight,  Viking  accomplished  its  mission  almost  flawlessly.31 

One  of  the  technological  developments  that  had  made  the  spaceflight 
revolution  possible  was  the  arrival  of  powerful  new  computers  with  new 
integrated  circuitry.  Langley  had  made  good  use  of  this  new  technology 


After  Langley,  Cortright  moved  to  executive  jobs  in  industry,  first  with  a  $4-billion  glass,  plastics, 
and  paperboard  packaging  company  in  Owen,  Illinois,  and  then  as  president  of  Lockheed's  California 
company. 

413 


Spaceflight  Revolution 

through  Paul  Fuhrmeister's  ACD,  but  little  had  been  done  programmati- 
cally  during  Floyd  Thompson's  term  as  director  to  formalize  and  consoli- 
date the  opportunities  made  possible  by  the  development  of  new  electronic 
capabilities.  During  Cortright's  directorship,  Langley's  computational  skills 
grew  much  more  sophisticated  and  organized.  Early  in  his  term  as  director, 
he  sponsored  the  development  of  the  Institute  for  Computer  Applications 
in  Science  and  Engineering  (ICASE)  at  the  center,  which  enabled  leading 
professors  from  around  the  world  to  use  Langley's  Star  computer  to  explore 
ways  of  applying  their  mathematical  abilities  to  the  development  of  new 
aerospace-related  technologies.  At  about  the  same  time,  Langley's  compu- 
tational specialists  refined  a  system  called  "NASTRAN,"  which  stood  for 
NASA  Structural  Analysis.  This  innovative  computer  system,  first  con- 
ceived at  NASA  Goddard  in  the  mid-1960s,  expedited  analysis  of  many 
types  of  structures,  including  large  launch  vehicles,  aircraft,  bridges,  and 
even  automobiles.  As  part  of  the  Cortright  reorganization,  a  NASTRAN 
Systems  Management  Office  evolved  inside  the  Structures  Division  of  the 
Structures  Directorate.  The  office  continued  until  1978. 32 

Cortright  sponsored  the  development  of  a  computer  program  that  was 
not  as  successful — IPAD,  or  the  Integrated  Program  for  Aerospace- Vehicle 
Design.  IPAD  engineers  were  to  help  optimize  aircraft  design  by  creat- 
ing interactive  computer  programs  relating  aerodynamics,  propulsion,  and 
structures.  The  computer  program  was  to  be  set  up  with  subroutines  that 
would  interact  through  a  master  central  computer.  At  the  master  com- 
puter, in  Cortright's  words,  "a  group  of  whizzes  can  design  airplanes  in  a 
quarter  of  the  time"  that  it  normally  took.  Ultimately,  IPAD  would  give 
Langley  researchers  important  insights  into  such  questions  as  how  a  change 
in  wing  shape  would  affect  the  weight  of  an  aircraft  or  how  changing  engine 
placement  would  influence  the  craft's  aerodynamics.33 

Unfortunately,  the  optimizing  of  an  aircraft  through  computers  was  not 
an  easy  matter.  Moreover,  veteran  Langley  researchers  were  not  accustomed 
to  being  put  into  the  role  of  actual  aircraft  designers.  By  formal  NACA 
policy,  researchers  were  to  provide  the  fundamental  information  that  aircraft 
designers  in  industry  would  need  to  do  their  jobs;  they  were  not  to  design 
particular  airplanes.  At  senior  staff  meetings  in  the  early  1970s,  Cortright 
had  it  "thrown"  at  him  that  "we  were  not  a  design  center"  but  a  place  where 
researchers  worked  on  the  problems  that  those  who  were  designing  would 
encounter.  Old-timers  like  Larry  Loftin  and  NACA  veteran  and  assistant 
director  of  the  Structures  Directorate  Richard  Heldenfels  warned  Cortright 
about  IPAD:  "That's  too  elaborate.  The  companies  design;  we  do  basic 
research."  To  which  Cortright  replied,  "Well,  how  do  you  know  you're  doing 
the  right  basic  research  if  you  don't  know  anything  about  design?  And  who 
is  going  to  do  this  type  of  advanced  thinking  of  getting  computers  to  play 
with  each  other?"  After  all,  at  that  time  computers  were  just  becoming 
important  in  the  nation's  research  centers.34 


414 


The  Cortright  Synthesis 

In  spite  of  the  resistance  from  his  senior  aerodynamicists  (who  argued 
that  they  indeed  did  know  a  lot  about  design  because  they  conducted 
fundamental  research),  Cortright  went  ahead  with  IPAD  and  eventually 
contracted  Boeing  for  the  system  for  just  under  $40  million.  IPAD  never 
achieved  everything  it  set  out  to  do,  partly  because  its  original  goals  were 
so  elaborate,  and  NASA  eventually  canceled  it  when  further  developments 
in  computer  technology,  including  the  development  of  personal  computers 
(PCs)  and  individual  work  stations,  made  IPAD  obsolete.  Still,  IPAD  did 
have  some  positive  results:  it  taught  engineers  in  the  aircraft  industry  and 
NASA  how  to  begin  conceptualizing  an  integrated  aircraft  design. 

A  more  noteworthy  Cortright  program  that  was  to  incorporate  new  com- 
putational capabilities  was  the  Terminal  Configured  Vehicle  and  Avion- 
ics (TCVA)  program,  which  NASA  co-sponsored  with  the  FA  A.*  In  1973, 
NASA  bought  Boeing's  737-100  prototype,  a  twin-jet  standard-body  air- 
liner, at  a  cost  of  only  $2.2  million  (the  market  value  of  a  used  737  in  1972 
was  about  $3.5  million)  and  outfitted  the  plane  for  research.  The  goal  of 
the  TCVA  program  was  "to  provide  improvements  in  the  airborne  systems 
(avionics  and  air  vehicle)  and  operational  flight  procedures  for  reducing  ap- 
proach and  landing  accidents,  reducing  weather  minima,  increasing  air  traf- 
fic controller  productivity  and  airport  and  airway  capacity,  saving  fuel  by 
more  efficient  terminal  area  operations,  and  reducing  noise  by  operational 
procedures."  More  specifically,  the  program  was  to  investigate  the  best  ap- 
proaches of  airliners  for  noise  abatement  and  improved  airport  acceptance 
rates,  cockpit  displays  of  traffic  information,  and  computer-aided  navigation 
for  improved  fuel  efficiency  and  closer  spacing  of  aircraft  landings.  Other 
experiments  eventually  carried  out  as  part  of  the  program  explored  such 
things  as  high-speed  runway  turnoffs  and  use  of  the  new  microwave  landing 
system  that  the  FA  A  was  developing.35 

In  summary,  the  task  of  the  TCVA  program  was  to  improve  air  opera- 
tions in  crowded  airport  areas,  which  was  a  goal  in  some  ways  as  formidable 
(and  certainly  as  worthwhile)  as  landing  humans  on  the  moon.  As  this 
program  evolved,  TCVA  became  concerned  with  the  quest  for  "the  cockpit 
of  the  future."  This  cockpit  would  incorporate  the  most  advanced  digi- 
tal avionics,  a  subfield  that  proved  to  be  among  the  most  ambitious  aero- 
nautical applications  of  computer  power.  The  program  also  was  one  way 
for  the  center  to  stay  involved  with  technologies  relevant  to  the  operation 
of  the  supersonic  transport,  even  after  Congress's  cancellation  of  the  SST 
in  1971.  In  particular,  TCVA  funding  allowed  NASA  researchers  to  continue 
investigating  the  advanced  digital  electronic  displays  and  automatic  guid- 
ance and  control  systems  that  Boeing,  the  former  SST  prime  contractor, 
had  been  considering  for  its  supersonic  airliner. 


The  word  "Avionics"  was  later  dropped,  leaving  the  acronym  TCV.    In  1982  the  entire  name  was 
changed  to  the  Advanced  Transport  Operating  System,  or  ATOPS  program. 

415 


Space/light  Revolution 


A  I/ '11- scale  model  of  the  737  air- 
plane is  prepared  for  testing  in  Langley's 
Anechoic  Antenna  Test  Facility  as  part 
of  an  effort  to  develop  a  collision  avoid- 
ance system  for  commercial  airliners. 


"I  got  the  center  into  avionics,"  Cortright  claims  with  pride,  "which  was 
totally  new."  Langley  insiders,  however,  argue  that  veteran  NACA/NASA 
test  pilot  John  P.  "Jack"  Reeder  and  other  researchers  actually  deserve  the 
credit.  Whoever  was  most  responsible,  the  results  were  impressive,  not 
just  for  Langley,  which  became  perhaps  NASA's  foremost  research  center 
specializing  in  the  application  of  microelectronics  to  aviation,  but  also  for 
the  FA  A  and  the  world  of  civil  aviation.  For  Boeing  specifically,  TCVA  led 
to  sophisticated  avionics  systems  incorporated  into  the  company's  new  757 
and  767  jet  airliners.36 

Cortright  also  believes  that  he  kept  Langley  active  in  supersonic  and 
hypersonic  aerodynamics,  which  were  two  fields  of  high-speed  performance 
that  had  lost  national  support  by  the  early  1970s.  "There  was  a  huge  effort 
to  cancel  hypersonics,"  he  recalls,  "but  I  fought  for  it  all  the  eight  years  I 
was  here,  and  kept  it  pretty  much  alive."  The  same  was  true  for  supersonics, 
which  became  something  of  a  dirty  word  around  the  country  following  the 
controversies  over  sonic  booms,  noise,  and  environmental  pollution,  which 
prompted  Congress  to  kill  the  national  SST  program  in  1971.  "We  pushed 
supersonics  even  when  we  were  told  to  drop  all  of  it."  In  both  these  cases, 
Cortright  had  to  back  off  his  philosophy  of  "total  display,"  because  the  only 
way  to  keep  such  politically  sensitive  work  going  in  the  early  1970s  was  by 
"bootlegging"  it.  "Sometimes  you  just  have  to  survive.  So  we  found  ways 
to  keep  research  going — critical  research  on  supersonics  and  hypersonics — 
even  when  Congress  wanted  anything  labeled  supersonic  transport  or  the 


416 


The  Cortright  Synthesis 


Cortright  won  initial  funding  in  1975 
for  what  became  the  National  Transonic 
Facility,  but  his  successor,  Donald  P. 
Hearth  (director  1975-1985),  would 
spend  considerably  more  time,  energy, 
and  money  than  expected  to  build  the 
innovative  new  wind-tunnel  design  and 
establish  its  regular  operation.  Left, 
Hearth 's  successor,  Richard  H.  Petersen 
(1985-1991),  stands  next  to  a  model  of 
the  Boeing  767  prototype  mounted  in 
the  NTF.  Below,  a  tunnel  technician 
checks  the  nozzles  that  inject  nitrogen 
gas  into  the  airflow. 


L-84-13,632 


417 


Space/light  Revolution 

like,  struck."37     Cortright's  critics,  however,  underplay  the  importance  of 
his  role  in  keeping  these  research  areas  alive. 

Cortright  initiated  one  last  project  before  leaving  office  in  1975,  and 
this  project  proved  to  be  critical  to  Langley's  future:  he  won  NASA's 
approval  for  $250,000  to  build  a  small  pilot  facility  to  prove  the  feasibility 
of  an  in-house  concept  for  a  large  cryogenic,  high  Reynolds  number  wind 
tunnel.  The  pilot  facility  did  prove  the  large  tunnel  feasible.  In  1983  after  a 
long  and  tortuous  developmental  period  under  Cortright's  successor,  Donald 
P.  Hearth,  this  major  new  facility,  the  National  Transonic  Facility  (NTF  or 
"The  Big  Cold  One"),  opened  at  Langley  to  rave  reviews,  not  to  mention 
the  audible  sighs  of  relief.38 

Critique  from  the  Old  Guard 

Not  everyone  was  happy  with  the  Cortright  reorganization  or  with  the 
way  the  spaceflight  revolution  had  been  changing  the  character  of  Langley 
Research  Center.  Many  of  these  critics  were  members  of  the  old  guard  from 
the  days  of  the  NACA  when  the  engineers  were  in  charge  and  fundamental 
research  was  their  bread  and  butter.  This  group  was  concerned  that  NASA 
Langley  was  heading  in  a  dangerous  direction:  toward  project  management 
work,  toward  bureaucracy,  toward  contracting  out,  and  toward  less  in-house 
capability.  Some  of  the  disenchanted  were  so  frustrated  and  depressed  by 
the  changes  taking  place  around  them  that  they  resigned  and  moved  to  new 
jobs  or  retired.  Others  stayed  on  to  fight  for  their  vision  of  Langley's  future. 

To  nurture  communication  among  his  supervisory  personnel,  Ed 
Cortright  had  initiated  the  practice  of  holding  brief  "retreats"  from  the 
center.  The  first  of  these  included  three  dozen  people  and  took  place  in 
the  summer  of  1969  at  Airlie  House,  a  quiet  resort  lodge  in  the  mountains 
of  western  Virginia.  A  second  and  third  retreat  followed  at  Airlie  in  1970 
and  1971,  followed  by  annual  retreats  to  Williamsburg,  Virginia,  which  was 
much  closer  to  Langley.  On  these  occasions,  some  of  the  Langley  veterans 
spoke  up  and  exchanged  opinions  about  the  proper  balance  of  research  and 
project  work.  One  of  the  retreat  participants  most  concerned  about  the 
decay  of  the  research  effort  at  Langley  was  Macon  C.  Ellis,  Jr.,  head  of  the 
MPD  Branch,  which  Cortright  would  abolish  as  part  of  his  September  1970 
reorganization.  (See  chapter  5.) 

Mike  Ellis,  however,  was  unhappy  even  before  the  dissolution  of  the 
MPD  Branch.  In  January  1969,  just  months  after  Cortright  took  over, 
Ellis  began  putting  down  on  paper  what  he  called  "Very  Rough  Thoughts 
on  the  Degeneration  of  Research-Mindedness  at  Langley  and  the  Need 
for  a  'Research  Renaissance.'"  He  eventually  shared  the  polished  version, 
"Rough  Thoughts  on  Status  and  Future  of  Research  at  Langley,"  with  only 
a  few  sympathetic  colleagues.  At  the  heart  of  his  critique  was  the  rather 
categorical  conclusion  that  "the  Center  is  no  longer  research-oriented;  it  is 

418 


The  Cortright  Synthesis 

project-oriented."  Driven  by  the  urgencies  of  the  space  program,  Langley 
management  had  grown  "much  more  concerned  with  projects  and  their 
problems  almost  to  the  exclusion  of  concern  and  action  related  to  the  status 
of  our  future  research."  NASA  management  was  only  giving  "lip  service" 
to  "the  concept  that  projects  require  a  research  cadre  at  the  Center  to  draw 
upon,"  but  in  reality  "most  of  management's  attention  is  to  projects,  taking 
for  granted  the  necessary  input  from  research  people."39 

In  his  confidential  paper,  Ellis  identified  several  circumstances  and  trends 
that  he  believed  were  indicative  of  what  he  called  "the  second-best  present 
position  of  research."  For  instance,  "In  former  years  of  almost  solid  research 
orientation,  we  had  monthly  meetings  of  research  'committees'  and  an 
overall  monthly  research-oriented  meeting  called  'the  department  meeting.' " 
But  in  recent  years,  the  monthly  department  meeting  had  been  replaced  by 
"a  very  large  meeting  open  not  only  to  the  research  and  engineering  staffs 
but  to  the  administrative  and  technical  staffs."  This  change  has  "pulled  the 
level  of  presentation  down  to  the  level  of  Scientific  American  or  even  Life 
magazine."  Such  a  meeting  "where  most  of  the  scientific  flavor  is  lost"  was 
fine  for  fiscal  and  other  administrative  personnel,  but  it  was  "demeaning 
to  research  personnel,"  especially  so  when  "no  alternate  forum"  for  the 
researchers  existed  at  any  level  inside  the  center.40 

Ellis  cited  more  evidence  for  the  deteriorating  status  of  research  at 
Langley:  "Awards  and  rewards  are  now  being  made  for  lesser  and  lesser 
accomplishments"  and  with  "no  apparent  recognition  of  the  diminution 
of  awards  for  research."  In  Ellis's  opinion,  NASA  was  even  giving  out 
some  of  its  exceptional  scientific  awards  for  achievements  that  were  "clearly 
not  scientific."  This,  he  lamented,  was  "demoralizing  to  the  hardworking 
successful  researcher  aspiring  to  true  scientific  excellence."  What  was  worse, 
NASA  management  was  not  keeping  up,  "even  to  a  cursory  degree,"  with  the 
details  of  the  various  specific  research  areas,  such  as  his  own  fields  of  space 
science  and  plasma  physics.  "How  often  does  management  above  the  division 
level  examine,  criticize,  and  evaluate  research  activities  and  objectives,"  he 
asked  pointedly,  "in  areas  other  than  those  very  close  to  applied  problems?" 
Furthermore,  why  was  it  that  in  nearly  all  activities  requiring  the  support  of 
administrative  or  technical  services,  "the  top  priority  and  attention"  always 
went  to  the  projects?  Pointing  to  two  recent  cases  of  personnel  transfers 
"out  of  research,"  Ellis  emphasized  that  the  prime  reason  given  by  both 
individuals  involved  was  that  "all  types  of  support  for  research,"  especially  in 
terms  of  help  from  the  shops  and  from  laboratory  technicians,  have  "declined 
to  such  a  point  that  research  activities  do  not  have  management  priority."41 

In  October  1969,  Ellis  sat  as  a  member  of  a  Space  Sciences  and  Tech- 
nology Steering  Committee,  which  Cortright  had  put  together  to  evaluate 
the  status  of  space  sciences  research  at  the  center.  At  this  meeting,  Ellis 
expressed  very  strong  opinions  about  the  deterioration  of  the  overall  sta- 
tus of  research  at  Langley  and  recommended  that  Langley's  various  efforts 
in  space  science  be  consolidated  by  creating  a  new  space  sciences  division. 

419 


Space/light  Revolution 

His  MPD  Branch  would  become  a  part  of  this  new  division.  In  a  typed 
statement  that  he  planned  to  present  to  the  committee's  chairman,  Herbert 
A.  Wilson,  Jr.,  but  apparently  never  did,  Ellis  wrote: 

It  is  believed  that  the  Center's  reputation  as  a  Research  Center  has  eroded  consider- 
ably over  the  past  10  years.  The  trend  has  been  toward  a  Center  primarily  concerned 
with  projects  and  contract  administration,  with  increasingly  scattered  and  smaller 
groups  of  research  excellence.  Good  research  people,  including  our  brightest  young 
ones,  who  see  the  situation  going  in  this  direction  with  no  apparent  positive  action  to 
preserve  a  strong  research  role,  do  not  stay.  A  degenerative  process  results  in  which, 
not  only  research  itself,  but  projects  which  depend  on  the  research  cadre,  may  also 
diminish  ultimately  leaving  only  management  activities  and  contract  administration. 
What  is  needed  at  this  point  is  a  hard  examination  of  the  research  situation  at  Lang- 
ley,  and  the  initiation  of  positive  steps  to  preserve  and  upgrade  the  status  of  research. 
To  the  researcher,  the  present  situation  appears  to  be  that  leadership  of  a  successful 
project  is  the  acme  of  accomplishment  at  Langley.  For  equal  recognition,  research 

accomplishment  must  be  so  truly  outstanding  and  highly  visible  as  to  discourage 

42 
many  competent  researchers. 

Unfortunately,  the  ideas  Ellis  did  express  on  these  matters  were  mostly 
neglected,  or  at  least  he  thought  so.  In  a  handwritten  set  of  notes  to  himself 
dated  10  November  1969  and  entitled  "So-called  Minutes  of  10/27/69  Mtg. 
of  Sp.  Sci.  &  Tech.  Steer.  Comm.,"  he  wrote,  "I  begin  to  detect  a  lack  of 
candor  in  reporting  what  actually  was  said  in  the  discussions.  There  were 
some  very  forthright  statements  made  (by  MCE  among  others)  that  are  very 
much  unreported.  ["MCE"  stands  for  himself,  Macon  C.  Ellis.]  Everybody 
seems  to  see  and  hear  what  he  wants  to,  regardless  of  what  is  presented."43 

Ironically,  Cortright,  in  his  reorganization  the  following  year,  did  create 
an  Environmental  and  Space  Sciences  Division  within  the  new  directorate 
for  space  as  Ellis  had  recommended.  Mike  Ellis,  however,  would  not  be 
a  part  of  the  new  division,  although  most  of  his  MPD  scientists  would. 
Instead,  Ellis  became  associate  chief  of  the  Hypersonic  Vehicles  Division 
under  John  V.  Becker — a  position  that  he  did  not  want.  As  for  his  ideas 
on  the  degeneration  of  research-mindedness  at  the  center,  nothing  in  the 
historical  record  suggests  that  the  Cortright  administration  tried  to  reverse 
the  trend  by  moving  away  from  projects  toward  basic  research  as  Ellis 
recommended;  the  administration  seems  only  to  have  pushed  the  center 
further  in  the  direction  of  projects. 

Another  individual  who  spoke  up  about  the  erosion  of  Langley  research 
was  John  V.  Becker,  Ellis's  boss  both  when  he  worked  in  the  MPD  Branch 
in  the  Aero-Physics  Division  and  now  in  his  position  in  the  new  Hypersonic 
Vehicles  Division.  Although  Becker  had  started  working  in  the  field  of 
high-speed  aerodynamics  at  NACA  Langley  under  John  Stack's  tutelage 
in  1936,  he  was  no  dinosaur.  When  it  came  to  keeping  up  with  the  times  and 
adapting  to  changing  circumstances,  Becker  always  proved  to  be  thoughtful, 

420 


The  Cortright  Synthesis 

pragmatic,  and  adept.  He  understood  NASA's  reasons  for  naming  Ed 
Cortright  Langley's  new  center  director.  As  Becker  has  written  in  the 
notes  to  his  private  papers,  "to  a  shrewd  aggressive  administrator  like  James 
E.  Webb,  who  had  studied  and  written  on  modern  management  techniques, 
Langley's  director  F.  L.  Thompson  and  his  staff  of  aging  Langley  employees 
must  have  appeared  old-fashioned  and  in  need  of  managerial  rejuvenation. 
Langley's  long-standing  division  structure  established  along  disciplinary 
lines  and  deliberately  provided  with  vague  names  like  'Full-Scale'  or  'Aero- 
Physics'  was  an  obvious  target  for  change.  Thus  in  1968  Webb  chose 
a  man  from  his  HQ  staff,  Ed  Cortright,  to  succeed  Thompson,  passing 
over  Langley  heir  apparent  C.  Donlan."  In  Becker's  opinion,  Cortright 
"undoubtedly  came  with  a  clear  mandate  to  shake  up  and  modernize  the 
Langley  operation."^ 

Becker  was  one  of  a  small  minority  of  Langley  managers  who  believed 
that  changes  were  definitely  needed.  "Communications  both  within  and 
among  the  research  divisions  were  clearly  inadequate  and  in  fact  entirely 
lacking  in  many  cases,"  he  admitted.  "Mechanisms  to  insure  accountabil- 
ity were  virtually  non-existent  at  all  levels  from  researchers  on  up  through 
division  offices."  What  bothered  Becker  specifically  was  that  researchers  in 
some  of  the  "aging  divisions"  were  pursuing  research  in  areas  completely 
unrelated  to  NASA's  aerospace  charter,  such  as  cancer-cell  biology,  inter- 
stellar mechanics,  and  even  personalized  rocket  propulsion  units  for  U.S. 
soldiers.  In  his  own  division,  he  saw  the  need  for  improvement.  Although 
he  had  supported  the  esoteric  pursuits  of  MPD  for  many  years,  Becker  now 
felt  the  branch  needed  to  be  discarded  so  that  its  staff  could  move  on  to 
more  promising  research.45 

The  Cortright  reorganization  was  not  exactly  the  change  Becker  had 
in  mind,  however.  "After  a  few  months  of  review  Cortright  concluded 
that  a  part  of  the  research  talent  should  be  reorganized  into  dedicated 
divisions  focusing  specifically  on  the  prime  NASA  systems  of  the  immediate 
future,  i.e.,  the  Space  Shuttle  and  the  Space  Station."  Becker  and  others 
on  the  senior  staff,  notably  Larry  Loftin,  disagreed  with  this  approach. 
"Loftin  and  I  argued  strongly  against  Ed's  concept  of  a  Space  Systems 
Division."  This  division  was  formed  by  Cortright  in  1969  and  named  the 
Space  Systems  Research  Division;  it  was  to  serve  as  the  lead  division  for 
support  of  NASA's  major  manned  space  systems.  In  Becker's  view,  "such 
a  management-favored  division  having  the  bulk  of  industry  contacts,  liberal 
funding,  shop  priorities,  etc.,"  would  "seriously  degrade  the  disciplinary 
divisions."  Rather  than  a  Space  Systems  Research  Division,  Becker  and 
Loftin  advocated  using  ad  hoc  task  groups  with  memberships  drawn  as 
required  from  the  disciplinary  divisions;  this  method  had  been  used  by 
Thompson  and  others  at  Langley  with  great  success.  Specifically,  the  duo 
suggested  an  ad  hoc  Shuttle  task  group  to  be  headed  by  Becker's  longtime 
associate,  NAG  A  veteran  Eugene  S.  Love,  then  the  associate  chief  of  the 
Aero-Physics  Division.  In  the  end,  Cortright  made  Eugene  Love  the  first 

421 


Space/light  Revolution 

head  of  the  new  Space  Systems  Research  Division,  the  unit  Becker  and 
Loftin  had  fought  against.46 

Clearly,  Cortright  was  "intent  on  making  large  breaks  with  former 
Langley  practices,"  Becker  believes.  His  Space  Systems  Research  Division 
"soon  became  the  centerpiece"  of  his  new  Space  Directorate.  This  direc- 
torate included  the  new  Environmental  and  Space  Sciences  Division,  which 
Becker  also  opposed  because  it  was  an  "obviously  earthbound"  area  of  re- 
search related  to  aerospace  only  through  the  use  of  satellite  photography 
and  satellite  navigational  aids.  Most  of  Becker's  former  MPD  researchers, 
with  the  exception  of  Mike  Ellis  and  a  few  others,  were  moved  into  these  two 
new  divisions;  the  work  in  these  divisions,  in  Becker's  opinion,  fell  outside 
Langley's  essential  charter.  Most  of  the  other  old  disciplinary  divisions  "con- 
tributed heavily"  to  the  new  nondisciplinary  divisions  as  well;  the  Applied 
Materials  and  Physics  Division,  the  historic  PARD,  was  "virtually  wiped 
out."  "Perhaps  to  ease  the  pain  Cortright  chose  to  make  these  reorganiza- 
tion moves  piecemeal  over  a  three-year  period."47 

Although  he  did  not  agree  with  many  components  of  the  Cortright 
reorganization,  Becker,  like  many  other  Langley  employees,  eventually  made 
his  peace  with  what  appeared  to  be  inevitable.  However,  this  did  not 
stop  him  from  speaking  out.  Over  and  above  the  specific  details  of  the 
center's  structural  reorganization,  what  concerned  Becker  most  deeply  in 
the  late  1960s  and  early  1970s  was  the  decline  of  the  general  research  culture 
at  NASA  Langley.  On  13  March  1970,  at  the  second  Airlie  House  retreat, 
he  presented  a  stinging  appraisal  of  what  had  been  happening  to  Langley 
in  the  years  since  Sputnik.  This  critique  captures  the  essence  of  how  the 
spaceflight  revolution  had  changed  Langley. 

Becker  began  his  talk  by  distinguishing  between  "research"  and  "pro- 
jects." He  defined  the  former  as  "the  whole  spectrum  of  Langley  Research 
Center  activity,  basic  and  applied,"  and  the  latter  as  "any  highly  focused 
ad  hoc  endeavor,  involving  usually  the  development  of  a  flight  vehicle 
system."  He  then  reviewed  the  interaction  of  research  and  projects  during 
the  NACA  era.  Becker  did  concede  that  the  entire  research  program  had 
always  been  "shaped  and  characterized  by  aircraft  and  missile  projects  and 
their  demands  for  R&D."  Langley  research  (even  during  the  NACA  period), 
although  "fundamental"  and  "basic,"  was  always  meant  to  be  "applied"  and 
"practical."  Indeed,  the  NACA  worked  on  projects,  but  these  projects  were 
driven  by  research  needs  such  as  the  lack  of  transonic  wind-tunnel  data,  not 
by  operational  flight  requirements.48 

After  Sputnik  such  research-related  projects  continued  to  be  "key  factors 
in  NASA  research"  as  evidenced  by  space  projects  Fire,  RAM,  the  hyper- 
sonic research  engine  (HRE),  the  barium  cloud  experiment,  and  continued 
involvement  in  the  development  of  specific  aircraft,  including  the  F-14,  the 
SST,  and  the  V/STOL.  "But  with  NASA  came  a  new  type  of  project  in- 
volvement: project  management,"  Becker  argued.  Because  of  the  space- 
flight  revolution,  Langley  became  an  "operating  and  management  agency 

422 


The  Cortright  Synthesis 

rather  than  a  research  agency."  Less  than  10  percent  of  its  budget  went  to 
in-house  use,  whereas  in-house  research  received  more  than  90  percent  of 
NACA  funding.49 

Given  the  natural  diversion  of  NASA  center  management's  interests  to 
"the  high-dollar  areas  of  project  operation  rather  than  research,"  it  was  "re- 
markable," in  Becker's  view,  that  "research  has  continued  to  survive  in  this 
atmosphere  to  the  extent  it  has."  Nonetheless,  despite  this  tenacious  ability 
to  survive,  Becker  felt  that  research  was  "losing  its  identity  in  the  NASA 
management/operations  jungle."  An  increasing  number  of  researchers  were 
saying  that  "NASA  is  a  management  agency  and  we  should  be  getting  into 
management."  Even  their  spouses  asked,  "What  projects  are  you  working 
on?"  or  "Shouldn't  you  get  into  a  project  to  get  ahead?"50 

In  his  talk  before  the  senior  staff  at  the  Airlie  House,  Becker  made  the 
interesting  point  that  research  involvement  with  projects  usually  hit  a  high 
mark  early  in  the  project;  as  the  project  progressed,  research  dropped  off 
significantly,  thus  leaving  only  project  management.  He  then  cited  three 
examples:  the  Apollo  program,  for  which  the  years  1959  to  1961  (before 
President  Kennedy  committed  the  country  to  the  lunar  landing)  were  "the 
big  years  from  the  research  view";  the  Viking  program,  which  in  1968 
occupied  eight  of  his  crew  in  the  Aero-Physics  Division  but  in  1970  involved 
only  one;  and  the  proposed  manned  Mars  program,  for  which,  in  his  opinion, 
"our  principal  involvement  in  research  will  be  over  before  a  project  office  is 
set  up."  These  examples  led  him  to  a  keen  observation:  "Any  one  LaRC 
[Langley  Research  Center]  project,  no  matter  how  large,  will  provide  only  a 
small  and  short-lived  part  of  the  total  project  stimulus  desirable  for  LaRC 
research."51  Research  demanded  not  more  projects  but  more  commitment  to 
the  development  of  a  broad  spectrum  of  technical  and  scientific  disciplines. 

"Ten  percent  of  Langley  research  personnel  are  now  in  Project  Offices," 
Becker  told  his  colleagues,  which  amounted  to  only  the  tip  of  the  iceberg 
given  how  much  additional  support  projects  required.  "Much  more  than 
10  percent  of  Center  management  is  involved  due  to  the  heavy  concentration 
of  dollars,  management  problems,  and  pressure  to  meet  schedules."  Becker 
wanted  to  know,  "Where  will  the  growth  end?"  In  the  eyes  of  many 
researchers,  Becker  advised,  project  management  appears  to  be  "primarily 
a  service  function,  not  a  research  function."52 

Becker  then  asked  the  key  question:  "Is  it  desirable  for  a  Research 
Center  to  develop  and  maintain  a  corps  of  project  management  people  whose 
services  are  available  for  successive  space  projects,  many  of  which  do  not 
spring  from  and  do  not  relate  strongly  to  the  main  research  activities  of  the 
Center?  If  it  is  desirable,  there  must  be  some  clear  benefits."  Try  as  hard  as 
he  could,  however,  Becker  could  not  find  any.  Did  Langley's  management  of 
nonresearch- related  projects  such  as  Scout  benefit  research  simply  by  being 
close  to  it,  the  "benefit  of  proximity"?  It  did  not,  according  to  Becker.  Did 
projects  people  return  to  research  after  the  termination  of  a  project  with 
"recharged  interests  and  new  ideas"?  Becker  saw  no  evidence  of  that.  Could 

423 


Space/light  Revolution 

projects  be  used  as  "a  haven  for  research  misfits"?  Yes,  but  this  was  really 
an  organizational  disadvantage  because  it  was  a  quick  fix  for  a  problem, 
not  a  long-term  resolution.  Do  effective  project  management  practices  and 
techniques  "rub  off  on  research  and  improve  it"?  No,  Becker  answered,  such 
methods  were  "actually  incompatible  with  research."53 

Becker  then  suggested  that  project  management  may  simply  be  the  "easy 
way  out  for  NASA  Headquarters."  Those  who  are  responsible  for  managing 
a  research  center  like  Langley  get  their  "brownie  points"  for  cooperating  with 
NASA  headquarters  in  its  pet  projects.  And  given  the  contemporary  social 
apathy  (and  even  growing  antipathy)  for  basic  science  and  technology,  it  was 
easier  for  NASA  to  "impress  local  politicians  and  public  with  management  of 
$800  million"  in  projects  than  it  was  to  explain  the  benefits  of  basic  research 
to  them  as  the  lower-key  NACA  officials  had  always  done.54 

Becker's  Airlie  House  presentation  struck  a  responsive  chord  with  most 
of  the  veteran  NACA  researchers  present,  and  several  of  them  asked  him 
afterward  for  copies  of  his  talk.  Naturally,  Cortright  and  those  involved 
primarily  with  project  management  made  their  counterpoints.  These  in- 
cluded the  notion  that  Langley  needed  projects  to  remain  in  the  public  eye. 
Becker  and  some  of  the  other  skeptics  interpreted  that  to  mean  that  the 
most  important  goal  was  not  to  keep  Langley  alive  and  prosperous  doing 
what  it  was  designed  to  do  best  but  simply  to  keep  it  alive  at  all  costs. 
Cortright  also  responded  with  the  suggestion  that  research  division  chiefs 
simply  would  have  to  do  a  better  job  of  "selling"  the  benefits  of  research 
to  the  center,  to  NASA  headquarters,  and  to  the  American  public  at  large. 
This  would  make  research  more  competitive  with  projects  and  keep  it  from 
being  pushed  aside.55 

Such  was  the  mindset  of  the  man  who  synthesized  the  spaceflight 
revolution  at  NASA  Langley.  Basic  research,  the  NACA's  reason  for  being, 
had  to  be  explained  not  only  to  the  country's  body  politic  as  in  the  past 
but  also  to  those  leading  the  organization  in  which  it  was  done.  As  NASA 
Administrator  Thomas  Paine  had  been  quoted  in  an  August  1969  newspaper 
article  entitled  "Future  Operation  of  LRC  Assured,"  the  fate  of  Langley 
Research  Center  "will  rise  or  fall"  depending  upon  the  fortunes  of  the  space 
program.56 

To  John  Becker,  the  administrator's  remark  epitomized  all  that  was 
wrong  with  NASA's  attitude  toward  research.  In  a  calculated  act  of  defiance, 
Becker  sent  a  copy  of  the  newspaper  article  quoting  Paine  to  Cortright  with 
a  two-page  "unsolicited  opinion."  Becker  pointed  out  to  the  director  that 
Langley  was  a  research  center  and,  therefore,  its  fate  should  not  depend 
on  the  "uncertain,  short-lived  projects  of  the  space  program."  Becker 
lamented  that  "the  current  movement  of  Langley  to  glorify  and  expand 
project  involvement  at  the  expense  of  research,  if  it  succeeds  eventually 
in  replacing  the  old  research  and  technology  development  program,  will 
put  Langley  in  the  'rise  or  fall'  predicament  described  by  Dr.  Paine."  He 
concluded,  "No  reply  necessary  or  expected."57 

424 


The  Cortright  Synthesis 

Cortright  responded  immediately  (in  what  he  called  an  "Unsolicited 
Reply")  with  the  single  comment  that  "Paine  did  not  make  the  statement 
during  his  talk — if  at  all."58  This  was  a  rather  lame  rejoinder  to  the  serious 
matters  raised  by  Becker's  cutting  memo.  Cortright,  however,  would  hear 
such  criticisms  again.  Although  he  and  the  other  consolidators  of  the 
spaceflight  revolution  would  prevail,  the  fundamental  tensions  caused  by  the 
essential  incompatibility  of  basic  research  and  project  management  would 
always  remain  at  Langley.  The  clarity  of  purpose  that  Langley  had  enjoyed 
during  its  history  as  an  NAG  A  laboratory  was  gone  forever. 


425 


Epilogue 


Either  space/light  will  be  proven  a  successful  revolu- 
tion that  opened  the  heavens  to  human  use  and  habi- 
tation, or  it  will  be  proven  an  unsuccessful  revolution 
that  demonstrated  in  its  failure  the  limits  of  techno- 
logical advance  .  ...  If  spaceflight  does  fail,  its  aban- 
donment will  represent  a  technological  counterrevo- 
lution of  great  consequence,  symbolizing  the  end  of 
progress  as  it  is  understood  today. 

—William  Sims  Bainbridge 
The  Spaceflight  Revolution: 
A  Sociological  Analysis 

America's  space  program  was  once  a  source  of  pride 
and  inspiration.  Now  it  is  a  shambles,  its  mission 
unclear,  and  its  very  existence  at  hazard. 

— Stuart  F.  Brown 
"Space  after  the  Race" 


By  the  early  1970s,  the  spaceflight  revolution  was  already  gasping  for 
breath.  Born  in  the  violence  of  the  cold  war,  driven  by  the  Sputnik  hysteria, 
perverted  by  the  specter  of  atomic  apocalypse,  and  ultimately  fed  by  the 
need  to  demonstrate  in  some  dramatic — maybe  even  primal  or  mythic- 
way  the  essential  superiority  of  the  American  democratic  system  over  its 
supposedly  entrenched  and  black-hearted  communist  enemies,  the  U.S. 
civilian  space  program  of  the  1960s  represented  an  extraordinary  national 
movement  that  could  be  sustained  only  for  a  brief  but  intense  period  when 
Americans  were  deathly  afraid  of  the  end  of  life  as  they  knew  it.  Kennedy's 
moon  shot  was  meant  to  alleviate  deeply  embedded  fear  and  paranoia.  By 
design,  his  Apollo  program  put  America  back  on  course  as  the  world's 
indisputable  technological  leader  and  the  guardian  of  a  safer  and  more 
hopeful  future. 

Until  the  United  States  beat  the  Soviets  in  space  and  answered  the 
enemy's  challenge,  nothing  dear  seemed  secure.  It  did  not  take  a  half  dozen 
lunar  landings  to  reassure  the  American  people  and  the  rest  of  the  world 

427 


Space/light  Revolution 

of  America's  virtues;  the  Apollo  11  mission  alone  was  enough.  With 
the  lunar  landing  accomplished,  the  country  had  faced  down  the  rival 
superpower  and  had  won  the  space  race;  the  public  could  feel  safe  once 
again.  The  threshold  of  terror  that  our  national  psyche  had  crossed  through 
because  of  Sputnik  receded  to  a  more  comfortable  distance.  In  turn,  the 
United  States  no  longer  felt  the  need  to  support  the  adventurous  and 
expensive  space  program  of  the  1960s.  That  venture  cost  Americans  as 
much  as  $5  billion  annually — or  4.5  percent  of  the  national  budget  per 
year.1  The  country  retreated  from  NASA's  expansive  plans  for  setting  off 
as  regular  travelers  through  the  Solar  System.  Instead,  the  country  looked 
to  accomplish  other,  more  earthly  objectives.  These  included  ending  the 
war  in  Vietnam,  cleaning  up  the  polluted  natural  environment,  and  finding 
a  better  way  to  get  along  with  the  communist  world  through  detente.2  This 
"new  age"  agenda  moved  the  American  people  back  to  the  earth,  not  away 
from  it,  and  whether  spaceflight  enthusiasts  realized  it  or  not  (and  for  the 
most  part,  they  did  not),  the  extraordinary  flurry  of  technological  activity  to 
get  humans  off  the  planet  and  on  their  way  to  other  worlds  far,  far  away  was 
over — at  least  for  the  time  being,  until  external  circumstances  would  once 
again  come  together  to  spur  the  inner  disquiet  that  launches  such  odysseys. 

"In  a  fundamental  way,  Apollo  was  about  leaving."  This  is  what 
Apollo  11  astronaut  Mike  Collins  said  shortly  after  the  end  of  the  lunar  land- 
ing program.  "It  was  our  first  move  outward,  off  the  home  planet."3  For 
a  variety  of  reasons,  a  lot  of  people  had  been  eager  to  go.  In  July  1969, 
Thomas  Paine,  the  NASA  administrator  at  the  time  of  Apollo  11,  predicted 
that  a  $5000  lunar  vacation  would  be  available  by  1990.  A  future  president, 
Ronald  Reagan,  was  one  of  thousands  who  joined  the  waiting  list  in  the 
summer  of  1969  for  the  first  commercial  flight  to  the  moon.  Pan  American 
World  Airways  actually  booked  reservations  for  lunar  flights  scheduled  in 
the  year  2000 — plans  that  today  look  more  than  a  little  premature  techno- 
logically, not  to  mention  commercially  because  Pan  Am  is  now  defunct.4 

Apollo  may  have  been  about  leaving  as  Collins  has  suggested,  but  not 
everybody  wanted  to  go.  In  truth,  Americans  of  the  1960s  were  ambivalent 
about  the  moon  shot.  Not  everyone  felt  the  urgency  or  saw  the  sense  in 
Jack  Kennedy's  goal.  Many  were  skeptical  or  outraged  at  the  billions  of 
dollars  spent.  Maybe  once,  only  once,  did  a  vast  majority  of  Americans 
approve  of  the  Apollo  program — on  20  July  1969,  the  day  Armstrong  and 
Aldrin  stepped  onto  the  Sea  of  Tranquility.  But  July  20  was  just  one  day 
in  a  very  turbulent  year.  The  year  1969  was  the  time  of  Woodstock  and 
the  Age  of  Aquarius,  the  movie  Easy  Rider,  the  bridge  at  Chappaquidick, 
campus  unrest,  anti- Vietnam  War  marches,  and  the  trial  of  the  Chicago 
Seven.  Cities  still  smoldered  from  race  riots.5 

"I  don't  know  whether  it's  the  noblest  expression  of  the  twentieth  century 
or  a  statement  of  our  lunacy,"  Norman  Mailer  thought  to  himself  while 
watching  the  first  lunar  landing  from  the  spectator's  gallery  at  Mission 
Control  and  later  wrote  in  his  offbeat  account  of  Apollo  11,  Of  a  Fire  on  the 

428 


Epilogue 

Moon.6  Even  more  offbeat  was  the  declaration  made  by  rock  critic  Greil 
Marcus  that  "Janis  Joplin's  new  album  is  more  important  than  landing  on 
the  moon."  Others  were  less  impressed.  "It  means  nothing  to  me,"  artist 
Pablo  Picasso  said.  Some  were  downright  hostile.  The  historian  and  social 
critic  Lewis  Mumford  called  the  moon  landing  "a  symbolic  act  of  war"  by  a 
"megatechnic  power  system  in  the  lethal  grip  of  the  myth  of  the  machine." 
He  predicted  that  "the  very  triumphs  of  technology"  would  soon  turn  "the 
planet  into  a  lunatic  asylum  or  a  crematorium."7  That  dire  prediction  was 
perhaps  misguided  in  several  respects,  but  it  certainly  overestimated  the 
passing  influence  of  Apollo. 

The  year  1969  simply  may  not  have  been  a  good  time  for  bold  predictions 
of  any  kind.  Wernher  von  Braun,  the  deus  ex  machina  of  the  Apollo 
program,  predicted  that  by  the  year  2000  "we  will  undoubtedly  have  a 
sizable  operation  on  the  moon;  we  will  have  achieved  a  manned  Mars 
landing;  and  it's  entirely  possible  we  will  have  flown  with  men  to  the 
outer  planets."8  "Undoubtedly"  and  "entirely  possible"  were  phrases  that 
reflected  NASA's  wishful  thinking  but  not  the  present  mood  of  most 
Americans. 

If  Apollo  was  about  leaving,  the  period  after  Apollo  was  to  be  about 
staying  home;  that  is  what  the  predictors  should  have  been  forecasting 
in  1969.  Having  visited  our  nearest  neighbor  half  a  dozen  times  between 
July  1969  and  December  1972,  and  having  brought  back  all  the  souvenirs 
our  travel  bags  could  carry,  we  were  ready  to  return  to  a  more  normal 
routine.  On  a  much  larger  scale,  the  American  space  program  seems  to  have 
gone  through  something  comparable  to  the  aftermath  of  a  family  vacation 
to  the  Grand  Canyon;  it  was  an  exciting  and  valuable  adventure  but  also 
an  exhausting  and  expensive  trek  from  which  Americans  needed  to  recover. 

This  was  certainly  the  attitude  of  the  general  public  and  their  president. 
Just  a  few  months  after  Apollo  11,  an  opinion  poll  showed  that  50  percent 
of  Americans  thought  the  country  should  "do  less"  in  space;  only  20  percent 
thought  the  country  should  "do  more."  As  for  President  Nixon's  view, 
only  four  days  after  flying  out  to  the  USS  Hornet  in  the  middle  of  the 
Pacific  Ocean  to  personally  welcome  back  the  crew  of  Apollo  11  from 
its  lunar  sojourn,  he  hit  NASA  with  the  first  of  several  sharp  blows  to 
its  ambitious  plans  for  the  future.  On  28  July  1969,  his  director  of  the 
Office  of  Management  and  Budget,  Robert  Mayo,  sent  a  letter  to  Thomas 
Paine,  the  NASA  administrator  who  had  predicted  lunar  vacations  by  1990; 
this  letter  stated  that  the  space  program's  budget  would  be  frozen  at 
$3.5  million  for  the  remainder  of  Nixon's  term.9  A  second  blow  came 
shortly  before  Christmas  1969:  President  Nixon  rejected  NASA's  appeal 
for  enough  additional  money  to  keep  production  of  the  Saturn  V  going. 
This  news  devastated  NASA's  long-range  plans.  A  few  weeks  later,  buoyed 
by  this  victory  for  fiscal  responsibility,  Nixon's  budget  director  along  with 
other  members  of  the  White  House  staff  informed  NASA  that  "there  is  no 


429 


Space/light  Revolution 

commitment,  implied  or  otherwise,  for  development  starts  for  either  the 
space  station  or  the  shuttle  in  FY  [fiscal  year]  '72. "10 

Barely  a  scintilla  of  hope  remained  for  the  continuation  of  the  spaceflight 
revolution.  That  small  chance  was  a  report  that  in  early  1970  sat  on  a 
desk  in  the  White  House  awaiting  a  reply.  Vice-President  Spiro  Agnew's 
Space  Task  Group  had  submitted  the  report  to  Nixon  for  consideration 
in  September  1969.  The  report,  issued  after  a  lengthy  seven-month  study 
made  in  response  to  Nixon's  own  request  for  a  "definitive  recommendation 
on  the  direction  the  U.S.  space  program  should  take  in  the  post- Apollo 
period,"  called  for  a  bold  continuation  of  the  aggressive  space  program 
of  the  1960s:  a  permanent  lunar  base  and  a  space  station  serviced  by 
a  completely  reusable  space  shuttle — technologies  long  envisioned  by  the 
pioneers  of  space  exploration  as  the  springboard  to  Mars  and  the  other 
planets.11  The  Agnew  team's  report  essentially  ratified  NASA's  own  vision 
of  where  the  space  program  needed  to  go  after  Apollo.12  But  the  glimmer 
of  hope  associated  with  the  fate  of  this  report  was  soon  squelched.  In  early 
March  1970,  President  Nixon  issued  a  policy  statement  entitled  "The  Future 
of  the  United  States  Space  Program,"  which,  although  conceding  that  "space 
activities  will  be  part  of  our  lives  for  the  rest  of  time,"  sounded  the  deathbell 
of  the  spaceflight  revolution  by  cautioning  that  "we  must  also  realize  that 
space  expenditures  must  take  their  proper  place  within  a  rigorous  system 
of  national  priorities.  What  we  do  in  space  from  here  on  in  must  become  a 
normal  and  regular  part  of  our  national  life  and  must  therefore  be  planned 
in  conjunction  with  all  of  the  other  undertakings  which  are  also  important  to 
us."  The  country  could  not  afford  more  "separate  leaps"  like  the  moon  shot, 
another  "massive  concentration  of  will,"  or  another  "crash  timetable."13  As 
one  scholar  who  has  examined  the  post- Apollo  retrenchment  in  detail  has 
said,  NASA  would  now  "have  to  get  down  on  the  ground  and  scratch  for 
seed  with  all  the  other  government  chickens."14  The  revolution  was  over. 
Times  were  again  to  be  "normal"  and  "regular." 

Congress  agreed  with  Nixon's  conservative  space  policy,  in  large  part 
because  the  country's  legislators  were  in  the  same  stay-at-home  mood  as 
the  president.  With  opinion  polls  clearly  demonstrating  a  declining  interest 
in  space  (poll  data  had  actually  shown  a  significant  decline  in  support  for 
the  space  program  long  before  the  achievement  of  the  first  moon  landing) , 
NASA  was  lucky  to  win  the  few  political  concessions  it  did  in  the  early 
1970s,  including  the  one  in  the  summer  of  1970  that  ensured  just  enough 
funding  to  launch  an  orbital  workshop  called  Skylab  inside  the  upper  stage 
of  an  already  assembled  Saturn  V.15 

By  1971  even  NASA  leadership  was  waving  a  white  flag.  James  Fletcher, 
who  took  over  as  NASA  administrator  in  April  1971,  understood  when  he 
accepted  the  Nixon  appointment  that  "there  is  no  way"  to  do  a  space  station 
and  a  space  shuttle  at  the  same  time.16  The  best  NASA  could  do  was 
to  approach  its  long-term  objectives  incrementally,  by  first  requesting  a 
shuttle — and  not  even  the  fully  reusable  vehicle  it  wanted — and  later  asking 

430 


Epilogue 

for  a  space  station  that  the  shuttle  could  eventually  service.  This  made  much 
less  sense  technologically,  but  it  made  sense  fiscally  and  politically,  and  that 
was  what  counted.  In  early  1972,  the  year  that  saw  the  last  two  Apollo  flights 
(Apollo  16  and  17),  President  Nixon  finalized  the  country's  retreat  from  the 
spaceflight  revolution  with  a  key  policy  decision:  the  space  station  was 
again  to  be  leapfrogged  and  postponed  until  some  indeterminate  time  in  the 
future.  For  the  time  being,  the  country  indeed  would  have  only  a  shuttle — 
and  a  scaled-down,  partially  reusable  version  of  the  vehicle  that  NASA  really 
wanted.  This  way  initial  development  costs  could  be  minimized.  The  fact 
that  operational  costs  for  such  a  shuttle  would  eventually  skyrocket  was 
something  for  a  future  president  and  NASA  administrator  to  worry  about. 
Thus,  NASA  and  the  space  program  entered  a  20- year  period  of  incremental 
politics  and  the  less-than-optimum  technologies  produced  by  such  politics, 
in  which  the  development  of  a  limited  shuttle  mission  sufficed  and  the  fight 
to  build  an  affordable  (and  almost  perennially  redesigned)  space  station 
dragged  on.17  Even  though  the  space  budget  would  approximately  double 
between  the  early  1970s  and  the  early  1990s,  the  result  would  nonetheless 
be  a  rather  pathetic  space  exploration  program  locked  in  earth  orbit.  Of 
the  155  U.S.  spacecraft  launched  between  1984  and  1994,  only  7  left  earth 
orbit.18 

One  could  argue  that,  even  with  this  reversal  of  fortune,  the  spaceflight 
revolution  never  really  ended — and,  for  that  matter,  Americans  did  not 
really  stay  home.  Pioneer  10,  launched  in  1972,  became  the  first  man- 
made  object  to  exit  the  Solar  System.  In  1976,  Viking  took  our  robots  to 
Mars.  And  beginning  in  1977,  the  Voyager  spacecraft  began  a  phenomenal 
"Grand  Tour"  of  the  outer  planets.  In  taking  us  into  earth  orbit,  even  the 
scaled-down  Space  Shuttle,  which  made  its  first  operational  flights  in  1981, 
took  us  someplace  we  needed  to  go  routinely  if  we  were  ever  to  become 
real  space  travelers.  Thus,  the  spaceflight  revolution  continued,  only  with 
a  temporarily  different  emphasis.  From  this  point  of  view,  what  Sputnik 
started  was  permanent.  It  was  only  a  matter  of  time  before  humankind 
again  spurted  ahead  on  target  to  some  distant  star. 

A  fundamental  change  in  the  ulterior  makeup  of  American  society  and 
government  also  survived  Apollo,  some  would  say.  As  historian  Walter 
McDougall  has  argued  in  The  Heavens  and  the  Earth:  A  Political  History 
of  the  Space  Age,  the  launch  of  Sputnik  in  1957  gave  birth  to  a  state  of 
"perpetual  technological  revolution"  through  technocracy,  which  McDougall 
defines  as  "the  institutionalization  of  technological  change  for  state  purposes, 
that  is,  the  state-funded  and  -managed  R&D  explosion  of  our  time." *  J  Once 
institutionalized,  technocracy  would  not  go  away;  this  is  McDougall's 
message.  The  Soviet  Union  was  the  world's  first  technocracy,  and  it  would 
stay  in  power  even  if  that  required  murderous  purges.  The  United  States 
was  the  world's  second  technocracy,  created  by  Sputnik,  and  although  not 
a  brutal  oppressor  like  Soviet  militaristic  totalitarianism,  America's  civilian 
technocracy,  which  would  stay  in  power  through  the  federal  bureaucracy  and 

431 


Space/light  Revolution 

military-industrial  complex,  would  eventually  prove  as  oppressive  in  terms 
of  the  social  engineering  it  directed  as  the  Soviet  system  it  was  designed  to 
fight.  Thus,  in  McDougall's  view,  the  spaceflight  revolution  did  not  end  in 
the  early  1970s,  and  it  never  will,  short  of  a  fundamental  spiritual  upheaval — 
a  conversion  that  would  place  far  less  faith  in  the  passing  material  values 
of  scientific  and  technological  progress  and  more  faith  in  the  transcendent. 
"When  that  day  arrives,"  McDougall  imagines,  "the  technocratic  pump  may 
cavitate,  the  human  heart  have  a  meltdown,  and  science  become  again  a 
branch  of  moral  philosophy."20 

But  nothing  about  human  history  is  perpetual,  and  such  days  of 
cavitation — or  some  less  cataclysmic  version  of  them — can  and  do  hap- 
pen. McDougall's  controversial  book  was  published  in  1985,  a  year  before 
the  Challenger  accident,  four  years  before  Tiananmen  Square,  six  years  be- 
fore the  fall  of  the  Berlin  Wall  and  the  collapse  of  the  Soviet  Union,  and 
nearly  a  decade  before  the  Republican  party  regained  majority  control  in  the 
U.S.  Congress  for  the  first  time  since  1954.  During  the  period  McDougall 
was  researching  and  writing  his  book,  President  Reagan,  the  hopeful  moon 
traveler,  strongly  endorsed  the  building  of  Space  Station  Freedom,  gave  his 
wholehearted  support  to  the  Space  Defense  Initiative  or  SDI  (better  known 
to  its  critics  as  "Star  Wars" ) ,  and  called  for  the  development  of  a  National 
Aero-Space  Plane  (NASP),  which  was  a  futuristic  hypersonic  vehicle  capable 
of  flying  in  and  out  of  the  atmosphere.  It  was  misnamed  by  Reagan's  speech- 
writers  "The  Orient  Express,"  because  such  a  spaceplane  might  someday  be 
able  to  fly  nonstop  from  New  York  to  Tokyo  in  a  few  hours.21  Although 
NASA  had  lost  many  major  battles  by  then,  agency  leadership  still  oper- 
ated in  the  1980s  according  to  the  momentum  model  of  the  1960s.  Because 
the  United  States  seemed  to  be  locked  in  mortal  competition  with  intran- 
sigent communism,  this  model  deemed  it  necessary  to  have  an  expansive 
space  program  doing  whatever  it  took  for  the  country  to  remain  number 
one  technologically. 

The  ambition  of  many  well-meaning  supporters  of  the  space  program 
was,  therefore,  to  secure  another  dramatic  presidential  commitment  for 
a  bold  new  initiative.  This  initiative  was  on  the  scale  of  Kennedy's 
lunar  commitment  in  May  1961,  which  energized  the  American  people 
so  successfully,  or  Reagan's  commitment  to  Space  Station  Freedom  in 
January  1984,  which  failed  to  do  anything  of  the  sort.  Even  critical  and 
scholarly  observers  of  the  U.S.  space  program,  including  those  who  lamented 
the  bureaucratization  of  NASA  and  the  devaluation  of  its  technical  culture 
since  Apollo,  have  made  that  suggestion.  For  example,  political  scientist 
Howard  McCurdy,  professor  of  public  affairs  at  American  University  in 
Washington,  D.C.,  suggested  in  his  article  "The  Graying  of  Space"  that 
NASA  could  best  rejuvenate  itself  by  aiming  at  a  big  new  goal  such  as  a 
trip  to  Mars.22  In  essence,  however,  such  suggestions  implied  that  American 
space  policy  would  be  best  served  if  the  momentum  model  could  be  refueled. 
This  analysis  missed  the  point  that  the  momentum  model  has  actually  been 

432 


Epilogue 

the  major  reason  for  the  failure  of  American  space  policy  since  Apollo  11 
and  that  it,  too,  has  been  the  cause  for  much  of  the  technical  decline  in  the 
NASA  organization.  If  NASA's  long-term  competency  and  productivity  in 
basic  R&D  are  a  main  concern,  U.S.  presidents  need  to  stop  making  such 
bold  pronouncements,  not  make  more  of  them. 

The  collapse  of  the  Berlin  Wall  in  1991  and  the  fall  of  the  Soviet  Union 
proved,  if  proof  were  necessary,  that  the  unimaginable  really  does  happen.  In 
a  very  short  span  of  time,  these  megahistorical  events,  which  were  certainly 
in  the  league  of  Sputnik  in  terms  of  unraveling  the  expected,  pulled  the  plug 
on  the  momentum  model  that  had  been  driving  the  American  space  program 
since  Sputnik.  U.S.  leadership  saw  that  the  ground  rules  of  what  had  been  a 
space  race  had  changed  fundamentally  almost  overnight.  President  George 
Bush  appointed  Daniel  Goldin  as  the  head  of  NASA  in  early  1992.  Goldin 
had  worked  the  prior  two  decades  in  U.S.  weapons  factories  on  top-secret 
military  space  hardware  projects,  but  as  the  NASA  administrator  he  proved 
so  innovative  that  President  Bill  Clinton  decided  to  keep  him  on  after  the 
election.  Goldin  understood  that  the  space  program  of  the  mid-1990s  and 
beyond  could  not  operate  on  the  basis  of  an  obsolete  world  order  of  fear, 
violence,  and  competition  with  an  enemy  who  was  no  longer  there.  The  key 
words  for  the  future  were  cooperation  and  international  partnerships.  The 
Soviet  Union  no  longer  existed;  now  there  was  only  Russia.  The  United 
States  would  still  be  concerned  about  Russian  cosmonauts,  but  not  as 
competitors  or  as  instruments  of  the  Soviet  propaganda  machine,  rather  as 
fellow  crew  members  in  an  international  space  program  working  alongside 
not  just  NASA  astronauts,  but  those  of  Germany,  Canada,  France,  Japan, 
and  other  nations.  The  world  had  turned  topsy-turvy  again.  Up  was  down; 
down  was  up.  Space  explorers  had  to  adjust. 

The  political  and  technological  imperatives  that  had  driven  the  American 
space  program  through  its  revolutionary  period  after  Sputnik  and  that  had 
tried  with  diminishing  effectiveness  to  move  it  after  Apollo  had  evaporated. 
With  their  disappearance,  much  about  the  space  program  had  to  change. 
In  Goldin's  pet  phrase,  NASA  needed  to  do  everything  "smaller,  faster, 
cheaper,  better."23  In  keeping  with  tough  economic  times  and  the  trend 
toward  downsizing  in  American  business  and  industry,  the  American  space 
program  had  to  learn  how  to  do  more  with  less.  To  meet  a  directive  from 
President  Clinton  in  1993,  NASA  moved  to  "reinvent"  itself  and  find  ways 
to  reduce  its  spending  by  30  percent  over  five  years,  which  resulted  in  a 
$30  billion  budget  cut.  (In  contrast,  in  the  summer  of  1990,  a  NASA 
advisory  panel  still  operating  from  the  momentum  model  and  chaired 
by  systems  engineering  guru  Norman  Augustine  of  the  Martin  Marietta 
Corporation,  had  called  for  a  10-percent  budget  increase  for  NASA  each 
and  every  year  into  the  early  twenty-first  century  just  to  meet  the  "critical 
national  needs"  in  space.24)  The  nation  would  still  move  boldly  out  into 
space,  but  in  cooperation  with  the  rest  of  the  world,  not  alone,  not  facing 
down  enemies,  and  not  making  great  leaps.  Instead  of  placing  such  a 

433 


Space/light  Revolution 

predominant  emphasis  on  government-directed  spending,  the  privatization 
and  commercialization  of  the  enterprise  gained  momentum.  Instead  of  the 
prospect  of  an  ever-increasing  technocracy,  there  was  a  growing  possibility 
that  space  exploration  might  move  ahead  with  less  overriding  government 
involvement  and  direction.  The  U.S.  space  program  would  still  be  part  of  the 
total  equation  of  national  strength  and  security,  but,  because  that  equation 
had  changed  so  dramatically  with  the  end  of  the  cold  war,  NASA  had  a 
chance  to  escape  from  the  old  mindset,  be  much  less  bureaucratic,  return 
to  basic  operating  principles  more  like  those  of  the  former  NACA  before 
Sputnik,  and  be  part  of  the  solution,  not  part  of  the  problem. 

All  of  this  augurs  significant  and  positive  change  for  the  American 
space  program  and  NASA.  This  may  be  especially  true  for  NASA  Langley 
Research  Center,  a  place  that  has  done  its  best,  not  only  to  abide  by,  but 
to  encourage  the  prevailing  momentum  model  through  the  1970s  and  1980s, 
even  though  it  meant  the  slow  but  sure  dissipation  of  the  traditional  research 
values  that  had  been  the  creative  heart  and  soul  of  the  center  since  its 
formative  days  in  the  1920s  under  the  NACA.  Langley,  perhaps  more  than 
any  other  NASA  facility,  needed  to  move  away  from  the  modus  vivendi  of 
the  cold  war  and  refocus  on  what  it  has  always  done  best:  basic  applied 
research.  The  cold  war,  perhaps  war  of  any  kind,  ultimately  used  up  more 
good  ideas  than  it  replenished. 

Following  Apollo  11,  NASA  Langley  had  moved  beyond  the  waning 
aspirations  of  moonflight  to  manage  the  Viking  missions  to  Mars — a  planet 
some  65  million  kilometers  (40.39  million  miles)  away.  Although  not  nearly 
as  celebrated  as  the  manned  lunar  landings,  Project  Viking  rated  as  one 
of  the  greatest  technological  triumphs  of  all  time,  in  some  respects  even 
surpassing  Apollo.  "To  that  day — maybe  to  this  day — Viking  was  the  most 
difficult  unmanned  space  project  ever  taken,"  a  justifiably  proud  Edgar 
Cortright  has  declared.  "It  brought  Langley  to  the  forefront  of  spacecraft 
technology,"  which  was  where  Cortright  thought  Langley  should  be  although 
many,  like  John  Becker  and  Mike  Ellis,  would  have  preferred  a  return  to 
the  ways  of  the  NACA  and  were  calling  for  a  "research  renaissance." 25  In 
Cortright 's  mind  and  that  of  other  avid  participants  in  the  spaceflight 
revolution,  the  Viking  robots  set  the  stage  for  a  manned  mission  to  the 
mysterious  red  planet,  an  adventure  that  they  felt  was  sure  to  come  before 
the  end  of  the  twentieth  century.  What  Viking  accomplished  at  Langley, 
however,  was  not  altogether  positive.  Although  it  did  keep  the  center 
focused  on  the  technologies  necessary  for  another  successful  space  mission, 
the  attention  given  to  Viking  continued  a  dangerous  trend  dating  from  the 
Apollo  period;  this  trend  marginalized  research  in  favor  of  managing  big 
projects,  and  thus  was  something  that  could  not  continue  forever  if  the 
center  was  to  live  up  to  its  original  charter  and  serve  the  nation  well  over 
the  long  haul. 

After  Viking,  Langley  gave  all-out  support  for  the  Space  Shuttle. 
Langley's  contribution  was  a  herculean  engineering  effort  that  was  absolutely 

434 


Epilogue 

necessary  for  the  Shuttle's  success  but  that  prevented  Langley  researchers 
from  seeking  breakthroughs  across  the  broadest  possible  technological  front. 
Because  NASA  was  staking  so  much  of  its  future  on  just  this  one  vehicle, 
every  NASA  facility  was  expected  to  provide  support  for  the  new  STS,  and 
Langley,  relying  on  in-house  capabilities  developed  largely  during  its  NACA 
years,  was  still  well  equipped  to  help.  The  problems  facing  a  hybrid  reusable 
vehicle  that  was  to  fly  into,  through,  and  back  from  space,  from  blastoff  to 
a  landing  on  a  runway,  were  unusually  extreme.  During  any  mission,  the 
Shuttle  had  to  pass  through  three  distinctly  different  flight  regimes:  hy- 
personic, supersonic,  and  subsonic.  Over  the  years,  dating  to  the  NACA 
period,  Langley  engineers  had  pioneered  research  in  each  one  of  these  speed 
regimes  and  knew  that  each  regime  by  itself  posed  difficult  flight  problems. 
Together,  these  problems  made  the  Shuttle  program  into  one  of  the  biggest 
challenges  Langley's  wind-tunnel  complex  ever  faced.  Just  to  confirm  the 
basic  aerodynamic  soundness  and  structural  integrity  of  the  Shuttle  required 
more  than  60,000  "occupancy  hours"  of  testing  in  several  tunnels  over  a  pe- 
riod of  an  entire  decade.  Thousands  of  hours  were  necessary  to  optimize  the 
design  of  the  Shuttle's  thermal  protection  system — a  unique  arrangement  of 
ceramic  tiles  protecting  the  reusable  vehicle  from  the  intense  heat  of  reen- 
try. A  team  of  over  300  Langley  engineers  and  technicians  put  together  by 
Director  Donald  P.  Hearth  took  on  this  sticky  problem  and  eventually  came 
up  with  a  practical  way  of  certifying  the  performance  of  the  hundreds  of 
glue-on  tiles.26 

Although  the  Shuttle  program  predominated,  many  studies  were  con- 
ducted at  Langley  that  explored  the  feasibility  of  advanced  spaceflight  tech- 
nologies for  future  U.S.  civil  and  military  operations — many  of  them,  but 
not  all,  were  for  manned  missions.  These  investigations  focused  on  new 
concepts  like  the  "aerobrake,"  which  was  a  technology  with  important  ap- 
plications for  space  transfer  vehicles  taking  payloads  and  crews  from  a  space 
station  to  the  moon,  to  Mars,  or  even  to  other  destinations  in  the  inner  Solar 
System.27  Langley  researchers  also  rehabilitated  a  few  neglected  older  con- 
cepts such  as  the  "lifting  body,"  a  promising  technology  that  had  been 
left  in  the  lurch  after  Sputnik  because  powerful  enough  rockets  did  not  yet 
exist  to  boost  such  a  heavy  vehicle  to  orbit.  In  the  1970s  and  1980s,  en- 
gineers at  Langley  helped  to  revitalize  the  lifting-body  concept,  and  they 
eventually  identified  a  highly  innovative  configuration  called  the  HL-20  that 
offered  a  less  expensive  alternative  to  the  Shuttle  for  direct  manned  access 
to  orbit.28  With  NASA's  new  emphasis  on  "smaller,  faster,  cheaper,  better" 
in  the  mid-1990s,  Langley's  persistent  interest  in  the  lifting  body  may  yet 
pay  off;  perhaps  it  will  be  a  convenient  way  of  traveling  to  and  from  the 
international  Space  Station  Alpha. 

Something  that  will  surely  pay  off  in  the  future  is  Langley's  sustained 
level  of  basic  R&D  in  hypersonics,  a  field  that  has  been  one  of  the  center's 
strengths  since  the  early  1950s  when  NACA  researchers  pioneered  concepts 
that  were  incorporated  in  the  North  American  X-15.  So  solid  was  Langley's 

435 


Space/light  Revolution 

reputation  in  hypersonics  that  when  the  NASP  program  began  under  joint 
NASA/DOD  sponsorship  early  in  President  Reagan's  second  term,  Langley 
became  the  project's  "lead  center."  By  the  early  1990s,  although  many 
critical  engineering  problems  had  yet  to  be  solved  before  such  a  radically 
new  technology  (the  prototype  was  labeled  X-30)  could  become  operational, 
Langley's  involvements  in  NASP  had  proved  the  value  of  not  only  the 
revolutionary  supersonic  combustion  ramjet  engine  (nicknamed  "scramjet") 
but  also  of  the  novel  integrated  lifting-body  aerothermal  shape  as  the  best 
suited  for  transatmospheric  flight.29 

Like  everything  else  that  originated  before  the  fall  of  the  Berlin  Wall, 
however,  the  NASP  program  was  driven  by  cold  war  thinking.  The  DOD 
had  planned  to  use  the  X-30  as  a  reconnaissance  vehicle,  a  weapon  against 
enemy  satellites,  and  possibly  as  a  nuclear  bomber.  Thus,  with  the  collapse 
of  the  Soviet  Union,  the  NASP  program  (as  well  as  the  SDI  with  which  it 
was  in  certain  ways  associated)  had  to  change  fundamentally.  Faced  with 
a  much  smaller  budget,  NASP  officials  developed  a  plan  in  1994  to  close 
out  work  on  the  X-30  and  transform  the  project  from  the  development  of 
a  prototype  military  vehicle  to  the  development  of  more  generic  hypersonic 
flight  technologies.30  Such  a  change,  although  it  meant  less  money,  signified 
a  reorientation  of  R&D  priorities  more  suited  to  an  open  intellectual 
environment  wherein  creative  thinking  about  a  wide  range  of  technological 
possibilities  can  be  better  cultivated  than  in  the  so-called  "black"  world  of 
top-secret  military  hardware  development. 

Another  important  effort  that  Langley  managed  to  continue  in  the  trou- 
bled post- Apollo  era  was  its  work  on  the  cutting  edge  of  wind-tunnel  technol- 
ogy; this  research  was  essential  to  the  center's  ability  to  stay  vitally  relevant 
in  the  twenty-first  century.  From  the  time  of  the  Wright  brothers,  the  wind 
tunnel  had  proved  to  be  the  essential  piece  of  versatile  experimental  machin- 
ery on  which  much  about  the  progressive  evolution  of  aircraft  depended. 
Without  a  new  state-of-the-art  tunnel,  especially  one  that  simulated  the 
vexatious  conditions  of  flight  at  transonic  speeds  (a  regime  that  both  high- 
speed aircraft  and  rockets  had  to  fly  through),  Langley  might  have  ceased 
being  Langley,  which  had  been  for  decades  a  mecca  for  aerodynamicists  from 
around  the  world. 

In  the  mid-1970s,  NASA  assured  Langley's  future  by  pursuing  funding  for 
what  became  the  National  Transonic  Facility  (NTF).  The  NTF  was  a  radi- 
cally new  wind  tunnel.  It  was  to  operate  cryogenically  with  liquid  nitrogen 
replacing  air  as  the  test  medium.  Conceptually,  the  NTF  had  roots  going 
back  to  Langley's  own  Variable-Density  Tunnel  (VDT)  of  the  early  1920s, 
which  was  a  revolutionary  machine  that  provided  a  pressurized  airflow  for 
higher  Reynolds  numbers  (the  parameter  representing  the  accuracy  of  test 
results  using  scale  models  in  a  wind  tunnel).  With  the  VDT,  the  NACA  had 
obtained  more  realistic  aerodynamic  data  than  anyone  had  yet  been  able  to 
obtain  from  any  previous  tunnel.  Like  the  VDT,  which  put  the  struggling 
early  NACA  on  the  world  map,  the  NTF  also  represented  breakthrough 

436 


Epilogue 

technology.  By  using  a  spray  of  liquid  nitrogen  to  refrigerate  the  tunnel's 
airstream,  thereby  dramatically  decreasing  its  viscosity,  the  all-important 
Reynolds  numbers  could  be  increased  up  to  five  times  with  no  increase  in 
model  loads  or  required  drive  power.31 

Authorized  by  Congress  in  1975,  the  incredibly  complex  NTF  design, 
which  was  meant  to  serve  the  transonic  R&D  needs  of  both  NASA  and  the 
DOD,  became  the  major  preoccupation  of  Langley  for  the  rest  of  the  decade 
and  beyond.  It  gave  Langley  Director  Don  Hearth  and  his  associates  endless 
headaches  before  all  the  kinks  were  worked  out  in  1982.  The  total  cost  of  the 
NTF  was  $85  million,  much  more  than  NASA  could  have  afforded  for  such 
a  general  research  tool  during  the  1960s  when  such  large  amounts  of  money 
could  not  have  been  spent  without  directly  tying  the  project  to  the  lunar 
landing.  But  the  result  was  formidable.  The  most  ambitious  and  expensive 
wind  tunnel  ever  built  anywhere,  the  NTF  was  quickly  put  to  use,  not 
only  testing  models  of  the  Space  Shuttle  (the  STS  had  to  pass  through  the 
transonic  speed  range  both  to  and  from  space)  but  also  evaluating  advanced 
subsonic  transports,  fighter  aircraft,  and  attack  submarines.  Beginning  in 
the  late  1980s,  NASP  designs  also  underwent  hundreds  of  hours  in  "The  Big 
Cold  One,"  as  it  came  to  be  known.32 

The  NTF  incorporated  the  most  modern  automated  equipment,  which 
included  four  separate  high-speed  digital  computers  capable  of  handling  up 
to  50,000  data  points  per  second.  These  computers  controlled  virtually 
everything  about  the  NTF,  from  monitoring  the  facility's  environment  to 
maneuvering  the  scale  models  in  the  test  sections.  When  the  NTF  first 
began  running  in  1982,  its  builders  felt  that  it  would  give  the  United  States 
a  full  five-year  technological  lead  over  other  countries  and  would  "provide 
aspiring  young  aerodynamicists  the  wherewithal  to  design  the  aerospacecraft 
of  the  future."33  Occasional  tunnel  mishaps  and  breakdowns  led  a  few 
critical  observers  to  question  whether  the  NTF  lived  up  to  its  billing,  but 
as  the  fits  and  starts  of  Langley's  earlier  wind-tunnel  history  demonstrate, 
it  sometimes  takes  years  of  refinement  before  such  a  basic  research  facility 
hits  its  full  stride. 

In  important  respects,  the  NTF  helped  to  pave  the  way  to  Langley's 
future  by  recalling  its  strong  NACA  roots.  Like  the  closed-loop  circuitry 
of  its  NTF,  NASA  Langley  by  the  mid-1990s  seemed  to  be  on  the  verge  of 
coming  full  circle,  not  returning  to  the  exact  place  where  it  had  been  40  years 
earlier — history  never  really  repeats  itself — but  arriving  at  a  familiar  spot 
closer  to  the  one  it  occupied  during  the  NACA,  before  Sputnik  changed 
history.  In  1993,  NASA  Langley  officials  estimated  that  60  percent  of  its 
employees  worked  on  aeronautics  and  only  40  percent  on  space.  By  the 
following  year,  the  margin  had  widened  to  70/30. 34 

At  Langley,  this  ratio  points  to  a  rather  curious  trend.  Thirty-five 
years  earlier,  it  was  the  aeronautical  enthusiasts  who  were  losing  ground 
and  becoming  disgruntled  with  the  transition  to  space  at  Langley.  In  the 
mid-1990s,  it  is  the  space  enthusiasts  who  worry  about  the  possibility  that 

437 


Spaceflight  Revolution 

the  rug  might  be  pulled  from  under  them.  Aeronautics  is  coming  back 
strong  at  Langley;  space  work  is  being  taken  away  and  assigned  to  other 
centers.  The  center  has  never  forgotten  that  the  first  "A"  in  NASA  stands  for 
"Aeronautics."  Indeed,  although  its  many  achievements  in  aeronautics  after 
Sputnik  were  usually  overshadowed  (as  they  are  in  this  book)  by  the  hype 
over  spaceflight,  the  center  still  managed  somehow  to  maintain  its  position 
as  a  world  leader  in  aeronautical  R&D.  Langley 's  contributions  included  the 
X-15,  VTOL  aircraft,  the  variable-sweep  wing,  the  supercritical  airfoil,  SST- 
related  technology,  sonic  boom  studies,  winglets,  runway  riblets,  composite 
materials,  laminar-flow  control  systems,  and  flight  avionics.  Still,  NASA  for 
its  entire  history  had  been  the  "space  agency,"  and  aeronautics,  although 
important,  had  played  second  fiddle. 

True  to  the  symbols  in  the  original  NASA  logo  (the  "meatball"),  which 
signify  that  the  organization  was  about  both  aeronautics  and  space,  Langley 
will  continue  to  play  a  viable  role  in  space.*  Although  Langley  probably 
will  not  be  conducting  any  space  missions,  it  still  stays  active,  in  less  vis- 
ible but  nonetheless  critical  ways,  in  space-related  research.  For  example, 
researchers  at  the  center  continue  to  help  design  the  space  station  and  the 
next-generation  space  shuttle,  repair  and  test  space-bound  instruments  and 
materials,  and  conduct  research  on  the  effects  of  clouds  and  pollution  on 
the  earth's  atmosphere.  As  organized  in  early  1994,  there  will  be  four  space 
"thrusts"  at  Langley:  first,  the  Spacecraft  Technology  Thrust,  which  will  en- 
tail the  continued  analysis  of  the  space  station  design  and  a  joint  effort  with 
NASA  Goddard  researchers  to  launch  a  spacecraft  called  the  EOS- AMI, 
an  unmanned  earth-observing  space  vehicle,  scheduled  for  launch  in  1998; 
second,  the  Remote  Sensing  Thrust,  which  will  involve  development  of  in- 
struments such  as  the  Lidar  In-Space  Technology  Experiment  (LITE),  which 
uses  a  laser  system  to  probe  the  atmosphere  and  provide  new  information 
about  the  biosphere;  third,  the  Space  Transportation  Technology  Thrust, 
which  will  encompass  the  design  of  the  next-generation  space  shuttle  and 
the  testing  of  spacecraft  materials  and  structures  to  see  how  they  hold  up 
to  extreme  heat,  sound,  air  pressure,  and  collisions  with  micrometeorites; 
and  fourth,  the  Atmospheric  Sciences  Thrust  (a  research  division  created 
by  Ed  Cortright  in  the  early  1970s),  which  will  focus  on  NASA's  "Mission 
to  Planet  Earth"  that  was  started  in  the  1980s  and  involves  monitoring  the 


*  Shortly  after  becoming  NASA  administrator,  Daniel  Goldin  replaced  the  NASA  "worm,"  which  had 
become  the  logo  of  the  space  agency  in  the  1970s,  with  the  aesthetic  blue  "meatball"  of  NASA's  early 
days.  Goldin  decided  to  do  this  after  visiting  NASA  Langley,  seeing  the  handsome  old  NASA  logo  on 
the  outside  of  the  West  Area  flight  hangar,  and  hearing  from  Langley's  center  director,  Paul  Holloway, 
that  a  return  to  the  old  symbol  would  help  improve  NASA  morale.  Goldin  agreed,  thinking  that  it 
might  be  a  good  way  of  signaling  NASA's  return  as  a  high-quality,  less  bureaucratic  organization.  The 
decision,  however,  gave  NASA's  graphics  department  and  public  affairs  offices  a  fit,  because  it  meant 
that  all  NASA  letterhead  and  promotional  materials  had  to  be  redone. 

438 


Epilogue 

ozone  hole  and  measuring  the  carbon  monoxide  and  aerosol  gases  in  the 
atmosphere.35 

Despite  the  general  reorientation  in  favor  of  aeronautics,  space  will  not 
be  forgotten  completely  at  Langley.  As  a  local  NASA  beat  reporter  has 
written,  "When  man  first  set  foot  on  the  moon  25  years  ago,  NASA  Langley 
Research  Center  was  a  star  player  in  the  space  program,  training  astronauts 
and  developing  the  lunar  lander.  But  in  the  future,  the  Hampton  research 
center  will  play  a  behind-the-scenes  role  in  space  research,  like  the  football 
lineman  who  blocks  for  the  star  running  back."36  This  general  trend  is  as 
it  should  be — and  as  it  was  at  Langley  before  Sputnik.  Revolutions  come 
and  go,  but  basic  research  needs  to  stay. 

Langley  Research  Center  survived  the  spaceflight  revolution  and  its 
aftermath.  It  helped  the  country  accomplish  some  wonderful  things  along 
the  way.  With  the  changes  in  orientation  that  are  now  taking  place  inside 
NASA,  the  center  is  likely  to  remain  productive  well  into  the  twenty-first 
century,  if  not  beyond.  To  do  so,  however,  Langley  does  not  need  to  turn 
back  the  clock  to  before  1957  or  necessarily  return  to  the  ways  of  the  NACA, 
but  it  does  need  to  refocus  on  what  it  does  best.  During  the  NACA  years,  it 
had  a  clarity  of  purpose  that  served  it  extremely  well.  Like  the  rest  of  NASA, 
it  needs  to  regain  much  of  its  lost  in-house  capabilities  and  not  depend  so 
heavily  upon  contractors.  If  the  staff  has  to  become  a  little  smaller  and  the 
overall  organization  a  little  leaner,  that  can  work  out  to  be  a  great  advantage. 
The  small  size  and  more  intimate  collegiality  of  the  NACA  staff  before  World 
War  II  stimulated  a  high  level  of  creativity  within  the  organization.  And 
above  all,  the  bureaucracy  in  Washington  must  be  streamlined  and  able  to 
give  a  wide  berth  to  people  who  want  to  follow  through  with  bright  ideas. 
Langley  during  the  NACA  years  operated  with  maximum  independence  from 
NACA  headquarters;  somehow  the  best  elements  of  that  kind  of  freedom 
need  to  be  revived  in  NASA. 

Someday,  the  country  may  see  another  extraordinarily  tempestuous  time 
like  the  12  years  following  Sputnik.  Hopefully,  if  that  time  does  come, 
Langley  will  again  be  in  a  general  state  of  good  health  as  it  was  when  Sputnik 
first  orbited  the  earth  and  will  be  able  to  make  the  tremendous  contributions 
it  made  during  the  first  spaceflight  revolution.  The  ways  and  means  of  the 
old  NACA  are  now  mostly  dead  and  gone,  and  Apollo  anniversaries  are  a 
time  to  celebrate  what  now  seems  an  impossible  achievement.  If  Langley 's 
future — and  NASA's  future — is  to  be  as  productive  as  the  NACA  and  early 
NASA  past,  the  spark  of  greatness  in  the  present  will  have  to  come  from 
something  else  besides  the  momentum  of  a  satellite  launched  nearly  40  years 
ago  by  a  country  that  no  longer  exists;  it  will  have  to  come  from  a  fresh  new 
stream  of  historical  development,  only  now  at  its  headwaters. 

Contrary  to  what  many  space  enthusiasts  inside  and  outside  of  NASA 
might  still  prefer  to  believe,  the  spaceflight  revolution  of  the  1960s  was 
something  of  an  aberration.  We  should  not  look  to  Apollo  as  a  model  to 
emulate.  That  is  what  the  children  of  the  twenty-first  century,  who  will  not 

439 


Space/light  Revolution 

remember  when  there  even  was  a  Soviet  Union  or  a  Khrushchev  toy  called 
Sputnik,  perhaps  need  most  to  understand  about  this  history.  The  nervous 
energies  that  took  us  on  our  wonderful  trips  to  the  moon  produced  little  in 
terms  of  a  creative  infrastructure  that  the  country  could  hope  to  sustain  for 
long;  rather,  they  created  a  more  expansive  and  plodding  bureaucracy  that 
has  sustained  itself  quite  well  for  the  past  three  decades  but  only  at  great 
cost  to  what  was  worth  saving  in  the  first  place. 

Hopefully,  Langley  Research  Center  and  the  rest  of  NASA  will  continue  to 
find  ways  through  the  dawn  of  the  next  millennium  to  recover  the  best  about 
what  they  have  been,  and  still  should  be,  no  matter  what  convulsion — or 
realized  impossible  dream — any  future  revolution  might  bring. 


440 


Abbreviations 


AAS  American  Astronomical  Society 

ABL  Allegheny  Ballistics  Laboratory 

ABMA  Army  Ballistic  Missile  Agency 

AC  alternating  current 

ACD  Analysis  and  Computation  Division 

ACLOPS  Agena-Class  Lunar  Orbiter  Project 

AFB  air  force  base 

AGARD  Advisory  Group  for  Aeronautical  Research  and 
Development 

AIAA  American  Institute  for  Aeronautics  and  Astronautics 

AMIS  Advanced  Man  in  Space 

AORL  Apollo  Orbital  Research  Laboratory 

ARPA  Advanced  Research  Projects  Agency 

AT&T  American  Telephone  and  Telegraph 

ATOPS  Advanced  Transport  Operating  System 

AVRO  A.V.  Roe  Aircraft  Corporation 

BellComm  Bell  Communications 

CAN  Nike-Cajun 

CIA  Central  Intelligence  Agency 

CM  command  module 

CNN  Cable  News  Network 

comsat  communication  satellite 

ComSatCorp         Communications  Satellite  Corporation 

DAN  Nike-Deacon 

DC  direct  current 

DOD  Department  of  Defense 


441 


Space/light  Revolution 


EDT 

EOR 

ESRO 

F 

FAA 

FRC 

FY 

G.E. 

Heos 

HRE 

IAF 

IAS 

IAU 

ICASE 

ICBM 
IEEE 
IGY 
ILSS 

Intelsat 

IPAD 

IRD 

JPL 

K 

LCF 

LEM 

LHA 

LITE 

LOLA 

LOPO 


eastern  daylight  time 

earth-orbit  rendezvous 

European  Space  Research  Organization 

Fahrenheit 

Federal  Aviation  Administration 

Flight  Research  Center 

fiscal  year 

General  Electric 

Highly  Eccentric  Orbiting  Satellite 

hypersonic  research  engine 

International  Astronautical  Federation 

Institute  of  Aeronautical  Sciences 

International  Astronomical  Union 

Institute  for  Computer  Applications  in  Science  and 
Engineering 

intercontinental  ballistic  missile 

Institute  of  Electrical  and  Electronic  Engineers 

International  Geophysical  Year 

Integrative  Life  Support  System 

International  Telecommunications  Satellite 
Consortium 

Integrated  Program  for  Aerospace- Vehicle  Design 

Instrument  Research  Division 

Jet  Propulsion  Laboratory 

kelvin 

Langley  Central  Files 

lunar  excursion  module 

Langley  Historical  Archives 

Lidar  In-Space  Technology 

Lunar  Orbit  and  Letdown  Approach  Simulator 

Lunar  Orbiter  Project  Office 


442 


Abbreviations 


LOR  lunar-orbit  rendezvous 

LORL  Large  Orbiting  Research  Laboratory 

LRC  Langley  Research  Center 

LTV  Ling-Temco-Vought 

MALLAR  Manned  Lunar  Landing  and  Return 

MALLIR  Manned  Lunar  Landing  Involving  Rendezvous 

MHD  magnetohydrodynamics 

MISS  Man-In-Space-Soonest  Project 

MIT  Massachusetts  Institute  of  Technology 

MOL  Manned  Orbiting  Laboratory 

MORAD  Manned  Orbital  Rendezvous  and  Docking 

MORL  Manned  Orbiting  Research  Laboratory 

MPD  magnetoplasmadynamics 

MSC  Manned  Spacecraft  Center 

MSFC  Marshall  Space  Flight  Center 

MTSS  Military  Test  Space  Station 

MWDP  Mutual  Weapons  Defense  Program 

NACA  National  Advisory  Committee  for  Aeronautics 

NASA  National  Aeronautics  and  Space  Administration 

NASA  HQ  NASA  headquarters 

NASA  HQA  NASA  Headquarters  Archives 

NASA  NP  NASA  Publication 

NASA  RP  NASA  Reference  Publication 

NASA  SP  NASA  Special  Publication 

NASA  TN  NASA  Technical  Note 

NASA  TR  NASA  Technical  Report 

NASP  National  Aero-Space  Plane 

NASTRAN  NASA  Structural  Analysis 

NATO  North  Atlantic  Treaty  Organization 

NHFRF  National  Hypersonic  Flight  Research  Facility 

NOSS  National  Orbiting  Space  Station 


443 


Space/light  Revolution 


NRL 

NTF 

OART 

OHC 

OSSA 

Pageos 

Parasev 

PARD 

PC 

PCC 

PSAC 

R&D 

RA 

RAM 

RAP 

RCA 

rpm 

RRA 

Saint 

SCAT 

SDI 

SEPC 

SETI 

SM 

SMD 

SNAP 

SST 

STG 

STL 

STS 

TAGIU 


Naval  Research  Laboratory 

National  Transonic  Facility 

Office  of  Advanced  Research  and  Technology 

Oral  History  Collection 

Office  of  Space  Sciences  Applications 

Passive  Geodetic  Earth- Orbiting  Satellite 

Paraglider  Research  Vehicle 

Pilotless  Aircraft  Research  Division 

personal  computer 

Peninsula  Chamber  of  Commerce 

President's  Science  Advisory  Committee 

research  and  development 

Research  Authorization 

Radio  Attenuation  Measurements 

Research  Airplane  Project 

Radio  Corporation  of  America 

revolutions  per  minute 

Resident  Research  Associate 

Satellite  Interceptor  project 

Supersonic  Commercial  Air  Transport 

Space  Defense  Initiative 

Space  Exploration  Program  Council 

Search  for  Extraterrestrial  Intelligence 

service  module 

Space  Mechanics  Division 

Systems  for  Nuclear  Auxiliary  Power 

supersonic  transport 

Space  Task  Group 

Space  Technologies  Laboratories 

Space  Transportation  System 

Tracking  and  Ground  Instrumentation  Unit 


444 


Abbreviations 


TCVA  Terminal  Configured  Vehicle  and  Avionics 

Texas  A&M  Texas  Agricultural  and  Mechanical  University 

TRW  Thompson-Ramo-Wooldridge 

TWGEP  Technical  Working  Group  for  Electric  Propulsion 

UHF  ultra-high  frequency 

USNC/IGY  U.S.  National  Committee/International  Geophysical 
Year 

VDT  Variable-Density  Tunnel 

V/STOL  vertical  and  short  takeoff  and  landing 

VHF  very  high  frequency 

WS-110A  Weapons  System-llOA 


445 


Notes 


A  Caveat  to  Chapter  Notes 

Many  of  the  following  notes  refer  to  a  piece  of  correspondence  by  file  number  (e.g.,  A32-1, 
LCF  [Langley  Central  Files]).  The  NACA/early  NASA  correspondence  files  have  been 
accessioned  by  the  National  Archives  (NASA  Record  Group).  Any  reference  that  appears 
as  above  is  now  in  the  National  Archives.  This  caveat  applies  only  to  correspondence  by  file 
number  in  the  Langley  Central  Files.  Other  references  to  documents  in  the  Langley  Central 
Files  and  Langley  Historical  Archives  are  correct  as  of  publishing  date. 

Prologue 

1.  George    Will,     "What    Paths    Would    the    Nation    Have    Taken    Had    Taylor    Lived?" 
Washington  Post,  20  June  1991. 

2.  Ward  Moore,  Bring  the  Jubilee  (London:  Heinemann,  1955). 

3.  Will,  "What  Paths." 

4.  John  Noble  Wilford,  "A  Spacefaring  People:  Keynote  Address,"  in  A  Spacefaring  People: 
Perspectives  on  Early  Spaceflight,  ed.  Alex  Roland,  NASA  SP-4405  (Washington,  1985), 
p.  69. 

5.  See  Walter  A.   McDougall,    The  Heavens  and  the  Earth:    A  Political  History  of  the 
Space  Age  (New  York:    Basic  Books,   1985),  p.  123;   also  Frederick  I.  Ordway  III  and 
Mitchell  R.  Sharpe,  The  Rocket  Team  (New  York:    Crowell,  1979),  pp.  377-378. 

6.  See  chap.  5,  "The  Satellite  Decision,"  pp.  112-134,  in  McDougall,  Heavens  and  the  Earth. 

7.  McDougall,  Heavens  and  the  Earth,  pp.  153-154;  on  Project  Vanguard,  see  Constance  M. 
Green  and  Milton  Lomask,  Vanguard:  A  History,  NASA  SP-4202  (Washington,  1970). 

8.  Wilford,  "A  Spacefaring  People,"  pp.  69-70.  For  a  persuasive  account  of  the  responses  of 
American  policymakers  to  the  launching  of  Sputniks  1  and  2,  see  Robert  A.  Divine,  The 

"Sputnik"  Challenge:  Eisenhower's  Response  to  the  Soviet  Satellite  (New  York:  Oxford 
University  Press,  1993);  for  an  interesting  revisionist  interpretation,  see  Rip  Bulkeley, 

The  Sputniks  Crisis  and  Early  United  States  Space  Policy  (Bloomington,  Ind.:  Indiana 
University  Press,  1991).  Bulkeley  argues  that  no  U.S.  president  could  have  quieted  the 
alarm  aroused  in  1957-1958  by  the  Soviet  demonstration  of  ICBM  capability.  If  the  Soviet 
Union  had  been  only  the  second  country  in  the  world  to  orbit  a  satellite,  the  American 
electorate  would  have  still  perceived  the  ICBM  demonstration  as  a  cataclysmic  threat 
to  its  strategic  security  and  as  a  turn  of  events  demanding  an  instant  and  dramatic 
response  from  the  country's  leadership.  Although  Dr.  Bulkeley's  thesis  has  some  merit,  I 
remain  convinced  that  the  post-Sputnik  media  riot  and  the  country's  generally  hysterical 

447 


Notes  for  Prologue 

reaction  could  have  been  generated  only  in  response  to  the  Soviets  having  the  first  space 
launch.  If  the  Soviets  had  been  second  into  space,  I  do  not  believe  that  the  NACA  would 
have  been  abolished  and  NASA  established  before  a  year  had  passed. 

9.  For  an  analysis  of  this  close  election,  see  pp.  350-365  of  Theodore  H.  White,  The  Making 
of  the  President,  1960  (New  York:  Atheneum,  1961). 

10.  For  a  critical  interpretation  of  the  history  of  the  NACA,   NASA's  predecessor,   see 
Alex  Roland,    Model  Research:     The  National  Advisory   Committee  for  Aeronautics, 
1915-1958,   2  vols.,   NASA  SP-4103  (Washington,   1985).     For  more  sympathetic  in- 
terpretations focusing  on  the  operations  of  the  NACA's  various  research   laborato- 
ries, see  Richard  P.  Hallion,   On  the  Frontier:   Flight  Research  at  Dryden,   1946-1981, 
NASA  SP-4303   (Washington,    1984);    Elizabeth  A.   Muenger,    Searching  the  Horizon: 
A  History  of  Ames  Research  Center,  1940-1976,  NASA  SP-4304  (Washington,  1985); 
James  R.  Hansen,  Engineer  in  Charge:  A  History  of  the  Langley  Aeronautical  Labora- 
tory, 1917-1958,  NASA  SP-4305  (Washington,  1987);  and  Virginia  P.  Dawson,  Engines 
and  Innovation:  Lewis  Laboratory  and  American  Propulsion  Technology,  NASA  SP-4306 
(Washington,  1991).  For  a  brief  survey  treatment,  see  Roger  E.  Bilstein,  Orders  of  Mag- 
nitude: A  History  of  the  NACA  and  NASA,  1915-1990,  NASA  SP-4406  (Washington, 
1989). 

11.  James  R.  Hansen,  "Transition  to  Space:  A  History  of  Space  Plane  Concepts  at  Langley 
Aeronautical  Laboratory,  1952-1957,"   Journal  of  the  British  Interplanetary  Society  40 
(Feb.  1987):  67-80.  For  additional  historical  and  engineering  analysis  of  the  technologies 
pertinent  to  the  development  of  hypersonic  boost-gliders  and  other  "space  planes,"  see 
The  Hypersonic  Revolution:  Eight  Case  Studies  in  the  History  of  Hypersonic  Technology, 
ed.  Richard  P.  Hallion  (Wright-Patterson  AFB,  Ohio:  Special  Staff  Office,  Aeronautical 
Systems  Division,  1987). 

12.  On  the  many  twists  and  turns  of  NASA's  space  station  development  since  the  late 
1950s,  see  Howard  E.  McCurdy,  The  Space  Station  Decision:  Incremental  Politics  and 
Technological  Choice  (Baltimore  and  London:  Johns  Hopkins  University  Press,  1990). 

13.  Eric  Hoffer,  Before  the  Sabbath  (New  York:  Harper  and  Row,  1979),  p.  55. 

14.  On  NASA's  controversial  decision  to  go  to  the  moon  via  lunar-orbit  rendezvous,  see  John 
Logsdon,  "Selecting  the  Way  to  the  Moon:  The  Choice  of  the  Lunar-Orbit  Rendezvous 
Mode,"  Aerospace  Historian  18  (June  1971):  66-70.  On  the  demise  of  the  American  SST 
program,  see  Mel  Horwitch,  Clipped  Wings:    The  American  SST  Conflict  (Cambridge, 
Mass.,  and  London:  MIT  Press,  1982).   On  the  space  station,  see  McCurdy,  The  Space 
Station  Decision,  pp.  30-31.  The  books  we  have  on  the  Space  Shuttle  Challenger  accident 
came  out  in  the  immediate  aftermath  of  the  tragedy  and  really  do  not  adequately  explore 
either  the  causes  of  the  explosion  or  the  character  of  the  follow-up  investigation  by  the 
presidential  commission  chaired  by  former  Secretary  of  State  William  Rogers.  Essentially, 
they  are  little  more  than  examples  of  sensational  journalism.  For  the  best  treatment  of 
the  forces  at  work  in  the  Space  Shuttle  Challenger  disaster,  see  Michael  Collins,  Liftoff: 
The  Story  of  America's  Adventure  in  Space  (New  York:    NASA/Grove  Press,   1988), 
pp.  222-242.    As  a  former  NASA  astronaut  (Gemini  X  and  Apollo  XI),  Collins  is,  of 
course,  sympathetic  to  NASA's  interests;  however,  he  provides  a  keen  analysis  of  what 
went  wrong  with  the  Space  Shuttle  program. 


448 


Notes  for  Chapter  1 

15.  Lewis  Carroll,  Alice's  Adventures  in  Wonderland  and  Through  the  Looking-Glass  (New 
York:  Macmillan  Co.,  1963),  bk.  2,  p.  43. 

16.  Thomas  S.  Kuhn,  The  Structure  of  Scientific  Revolutions  (University  of  Chicago  Press, 
1962). 

17.  Ibid.,  p.  91. 

18.  For  an  extensive  and  interesting  application  of  Kuhn's  theory  of  revolution,  see  Edward 
Constant  II,  Origins  of  the  Turbojet  Revolution  (Baltimore:   Johns  Hopkins  University 
Press,  1980). 

19.  William  Sims  Bainbridge,  The  Space/light  Revolution:  A  Sociological  Analysis  (New  York: 
John  Wiley  &  Sons,  1976),  p.  1. 

20.  On  the  Committee  for  the  Future,  see  Bainbridge,  The  Spaceflight  Revolution,  pp.  12-14 
and  158-197.  On  the  deviancy  of  science-fiction  fandom,  see  pp.  227-233. 

21.  Ibid.,  p.  4. 

Chapter  1 
The  Metamorphosis 

1.  In  the  wake  of  Sputnik,  Eisenhower  created  a  President's  Science  Advisory  Committee 
(PSAC)  from  an  existing  science  advisory  board  that  was  helping  the  Office  of  Defense 
Mobilization;    chairing  the  new  PSAC  was  James  R.  Killian,   Jr.,   president  of  MIT 
and  new  special  assistant  to  the  president  for  science  and  technology.     Eisenhower 
asked  the  committee  to  identify  American  objectives  in  space  and  advise  him  on  how 
those  objectives  could  best  be  met.    On  22  Feb.  1958,  the  PSAC  recommended  in  a 
preliminary  staff  document  entitled  "Organization  for  Civil  Space  Programs"  (copy  in 
NASA  HQA)  that  a  new  space  agency  be  established  with  the  NACA  as  its  nucleus. 
NACA  leaders  learned  about  this  recommendation  and  passed  informal  word  of  it  to 
their  laboratories  soon  thereafter.  On  2  Apr.  1958,  all  NACA  Main  Committee  members 
received  a  confidential  letter  from  the  NACA's  executive  secretary  informing  them  that 
the  president,  following  the  PSAC  recommendation,  had  just  sent  a  plan  for  a  new  space 
agency  to  Congress. 

See  John  F.  Victory's  letter  to  Floyd  L.  Thompson,  Langley  associate  director,  with 
attached  memorandum  from  D wight  D.  Eisenhower  to  the  secretary  of  defense  and  the 
chairman  of  the  NACA,  2  Apr.  1958,  plus  the  attached  White  House  press  release, 
marked  "strictly  confidential"  and  dated  2  Apr.  1958,  12:00  noon  EST,  with  the  text 
of  Eisenhower's  message  to  Congress.  The  letter  from  Victory  to  Thompson  is  in  the 
folder  "Space  Material,"  Floyd  L.  Thompson  Collection,  LHA. 

For  a  historical  analysis  of  Killian's  and  the  PSAC's  role  in  the  establishment  of  NASA, 
see  Walter  A.  McDougall's  Pulitzer  Prize-winning  book,  The  Heavens  and  the  Earth:  A 
Political  History  of  the  Space  Age  (New  York:  Basic  Books,  1985),  pp.  170-172. 

2.  For  Eisenhower's  reaction  to  Sputnik  and  his  concerns  about  the  militarization  of  space, 
see  McDougall,  Heavens  and  the  Earth,  pp.  158-176.    The  reader  needs  to  pay  special 
attention,  however,  to  McDougall's  interpretation  of  the  NACA  in  these  pages.  On  page 
164,  for  instance,  McDougall  argues  that  "by  the  mid-1950s  the  venerable  NACA  was 
slumping.  It  was  the  best  equipped  aeronautical  research  organization  in  the  world,  but 
institutional  conservatism  and  financial  strictures  rendered  its  future  very  dubious."   In 

449 


Notes  for  Chapter  1 

the  opinion  of  other  historians,  including  myself,  McDougall's  portrayal  of  the  NACA 
in  its  last  years  is  off  the  mark.  The  NACA  was  not  in  any  slump.  In  the  years  just 
preceding  1958,  the  NACA  was  involved  in  some  of  the  most  productive  work  it  had  ever 
done.  McDougall's  interpretation  implies  that  the  NACA  died  of  old  age.  An  alternative 
interpretation,  which  I  support,  would  suggest  that  the  NACA  was  killed  in  its  prime. 
It  was  not  Sputnik  per  se  that  killed  it;  it  was  the  hysterical  American  reaction  to  the 
threat  posed  by  the  Soviet  satellite.  For  criticism  of  McDougall's  interpretation  of  the 
late  NACA  similar  to  my  own,  see  Richard  P.  Hallion's  review  of  Heavens  and  the  Earth 
in  Technology  and  Culture  28  (Jan.  1987):  130-132,  and  Virginia  P.  Dawson,  Engines 
and  Innovation:  Lewis  Laboratory  and  American  Propulsion  Technology,  NASA  SP-4306 
(Washington,  1991),  p.  162. 

3.  Alex  Roland,  Model  Research:  The  National  Advisory  Committee  for  Aeronautics,  1915- 
1958,  2  vols.,  NASA  SP-4103  (Washington,  1985),  1:1-25.    On  the  NACA's  inaugural 
meeting,  see  the  "Minutes  of  Meeting  of  National  Advisory  Committee  for  Aeronautics 
held  in  the  War  Department  of  Washington,  D.C.,  April  23,  1915,"  cited  in  Roland,  Model 
Research,  1:27  and  326  n.  2. 

4.  Charles  J.  Donlan  interview  with  author,   Hampton,   Va.,    17  Aug.   1990,   transcript, 
pp.  34-35,  OHC,  LHA. 

5.  For  the  full  text  of  the  NACA  charter,  see  app.  A  to  James  R.  Hansen,  Engineer  in 
Charge:  A  History  of  the  Langley  Aeronautical  Laboratory,  1917-1958,  NASA  SP-4305 
(Washington,  1987),  p.  399. 

6.  Tom  D.  Crouch  offers  an  excellent  and  fair  appraisal  of  Samuel  P.  Langley's  mixed 
contributions   to   aeronautics   in   A   Dream   of   Wings:     Americans    and   the   Airplane, 
1875-1905  (New  York  and  London:  W.  W.  Norton  &  Co.,  1981).  See  especially  chap.  7, 
"S.  P.  Langley:    The  Scientist  as  Engineer,"   pp.  127-156,  and  chap.  12,   "The  Great 
Aerodrome,"  pp.  255-283.  A  1988  paperback  version  of  Crouch's  book  is  available  from 
the  Smithsonian  Institution  Press. 

7.  On  the  NACA's  annual  Aircraft  Manufacturers'  Conferences  (later  called  NACA  inspec- 
tions), see  Roland,  1:111-114  and  232-233,  and  my  own  Engineer  in  Charge,  pp.  148-158. 

8.  This  opinion  was  expressed  by  British  aviation  journalist  C.  G.  Grey  in  an  obituary  for 
NACA  Director  of  Research  George  W.  Lewis,  published  in  The  Aeroplane,  27  Aug.  1948. 
More  recently,  distinguished  American  aviation  historian  Richard  K.  Smith  wrote  in  the 
Smithsonian  Institution's  Milestones  of  Aviation  (New  York:  Hugh  Lauter  Levin,  1989) 
that  "in  the  years  1928-1938  no  other  institution  in  the  world  contributed  more  to  the 
definition  of  the  modern  airplane  than  the  Langley  Laboratory  of  the  U.S.  National 
Advisory  Committee  for  Aeronautics"  (p.  240). 

For  the  history  of  Ames  laboratory,  see  Edwin  P.  Hartmann,  Adventures  in  Research: 
A  History  of  Ames  Research  Center,  1940-1965,  NASA  SP-4302  (Washington,  1970), 
and  Elizabeth  A.  Muenger,  Searching  the  Horizon:  A  History  of  Ames  Research  Center, 
1940-1985,  NASA  SP-4304  (Washington,  1985);  and  for  a  history  of  Lewis,  see  Virginia  P. 
Dawson,  Engines  and  Innovation.  On  the  High-Speed  Flight  Station  at  Muroc  Dry  Lake 
(later  renamed  Dryden  Flight  Research  Center),  see  Richard  P.  Hallion,  On  the  Frontier: 
Flight  Research  at  Dryden,  1946-1981,  NASA  SP-4303  (Washington,  1984).  The  best 
work  so  far  on  the  history  of  Wallops  Island  is  Joseph  A.  Shortal,  A  New  Dimension: 
Wallops  Flight  Test  Range,  The  First  Fifteen  Years,  NASA  RP-1028  (Washington,  1978). 

450 


Notes  for  Chapter  1 

9.  Jerome  C.  Hunsaker  interview  with  Walter  Bonney,  2  Nov.  1971,  transcript,  p.  1,  OHC, 
LHA. 

10.  For  all  the  details,  plus  a  critical  analysis,  of  the  NACA  committee  system,  consult 
Roland,  Model  Research,  esp.  app.  B,  2:423-465.    For  a  rough  breakdown  of  the  roles 
played  by  the  various  committees,  see  Hansen,  Engineer  in  Charge,  pp.  5-9. 

11.  A  biography  of  Hugh  Dryden  is  badly  needed.  The  place  to  start  is  Richard  K.  Smith,  The 
Hugh  L.  Dryden  Papers,  1898-1965:   A  Preliminary  Catalogue  of  the  Basic  Collection 
(Baltimore:     Milton  S.   Eisenhower  Library,   Johns  Hopkins  University  Press,    1974). 
Dr.  Smith,  who  compiled  and  edited  this  collection  of  Dryden's  papers,  informs  me  that 
the  collection  is  in  truth  a  rather  meager  potpourri. 

Our  picture  of  George  W.  Lewis  is  a  little  more  complete.  See,  for  example,  my  chapter, 
"George  W.  Lewis  and  the  Management  of  Aeronautical  Research,"  in  William  P.  Leary's 
biographical  anthology,  Aviation's  Golden  Age:  Portraits  from  the  1920s  and  1930s  (Iowa 
City:  University  of  Iowa  Press,  1989),  pp.  93-112.  Those  interested  in  source  materials 
about  George  Lewis  should  consult  my  short  bibliographical  essay  about  him  at  the  end 
of  Leary's  book  (pp.  186-187). 

12.  Some  of  the  "young  Turks"  came  from  Langley  (for  example,  Robert  Gilruth  and  Max 
Faget,   future  leaders  of  the  NASA  Space  Task  Group,   which  put  together  Project 
Mercury),  but  the  leading  group  developed  at  NACA  Lewis.    This  group  was  led  by 
Lewis's  dynamic  associate  director,  Abe  Silverstein,  and  included,  among  others,  such 
promising  young  men  as  George  Low,  who  before  too  many  years  would  be  heading 
NASA's  Manned  Lunar  Landing  Task  Force,  and  Edgar  M.  Cortright,  Jr.,  future  director 
of  NASA  Langley  (1968-1976).    To  be  a  "young  Turk"  in  1957-1958  meant  to  be  an 
NACA  employee  who,  in  historian  Alex  Roland's  words,  "wanted  the  NACA  to  campaign 
for  a  broad  new  role  in  space"   (Model  Research,  1:292).    The  title  apparently  dated 
from  an  informal  NACA  dinner  meeting  hosted  by  NACA  Chairman  Jimmy  Doolittle  in 
Washington  on  18  Dec.  1957.     During  the  dinner,  an  NACA  generation  gap  became 
obvious  over  the  issue  of  what  the  NACA's  role  should  be  in  space,   and  Langley's 
intemperate  John  Stack  called  Hugh  Dryden  "an  old  fogey."    To  qualify  the  inferences 
drawn  from  Roland's  portrayal,   I  must  add  that  Stack's  passion  rested  squarely  in 
aeronautics;  he  was  hardly  a  "space  cadet"  as  Roland  (and  McDougall)  seems  to  suggest. 
A  few  months  after  the  Washington  dinner,  in  the  spring  of  1958,  Silverstein  reported 
to  Washington,  at  Dr.  Dryden's  request,  to  take  on  the  huge  job  of  inventing  NASA's 
spaceflight  development   program.      He  eventually  brought   with  him  to  Washington 
several  other  Lewis  people.   On  this  subject,  see  also  Dawson,  Engines  and  Innovation, 
pp.  163-166.   In  Heavens  and  the  Earth,  McDougall  refers  to  the  "young  Turks"  as  the 
NACA's  "frontier  faction"  (p.  200).  However,  the  NACA  actually  employed  two  frontier 
factions:  one  devoted  to  space  and  one  devoted  to  aeronautics.  The  spaceflight  revolution, 
in  key  respects,  left  members  of  the  second  group  behind. 

13.  "Biographical  Sketch  of  Dr.  T.  K.  Glennan,"  Langley  Air  Scoop,  15  Aug.  1958.  On  the 
front  page  of  the  same  issue  was  a  news  item,  "President  Announces  Nominees  to  Head 
NASA." 

14.  Both  Roland  and  McDougall  refer  to  the  occasion  when  Hugh  Dryden  told  the  House 
Space  Committee  in  the  spring  of  1958  that  some  of  the  ideas  being  proposed  for  putting 
Americans  into  space  have  "about  the  same  technical  value  as  the  circus  stunt  of  shooting 

451 


Notes  for  Chapter  1 

a  young  lady  from  a  cannon."  Sensibly  cautious  remarks  like  this  did  not  endear  Dryden 
to  congressmen  who  were  in  a  rush  to  see  Americans  in  space  and  made  his  selection  as 
the  first  NASA  administrator  politically  impossible.  Model  Research,  1:299;  Heavens  and 
the  Earth,  pp.  195-196. 

15.  A  seven-page  carbon  transcript  of  "Glennan  Message  to  NACA  Employees,"    dated 
22  Sept.  1958,  can  be  found  in  the  folder  marked  "Space  Material"  in  the  Thompson 
Collection,  LHA.     Glennan  did  not  make  his  first  trip  to  Langley  until  Wednesday, 
7  Jan.  1959.  Accompanying  the  administrator  was  Hugh  Dryden,  deputy  administrator, 
as  well  as  a  number  of  other  NASA  headquarters  officials.  The  group  arrived  at  Langley 
around  11:30  a.m.,  after  a  brief  stop  to  see  research  facilities  at  Wallops  Island,  and 
they  returned  to  Washington  in  the  late  afternoon.     Glennan  and  his  staff  toured  a 
number  of  Langley  wind  tunnels  in  which  various  tests  were  being  conducted  in  support 
of  NASA's  nascent  space  programs.    See  the  article  "Glennan  Views  Facilities,  Hears 
Research  Reports,"  in  NASA  Langley  Air  Scoop,  9  Jan.  1959. 

16.  "Glennan  Message,"  pp.  1-2. 

17.  T.  Keith  Glennan,  "The  First  Years  of  the  National  Aeronautics  and  Space  Administra- 
tion," unpublished  diary,  1964,  Eisenhower  Library,  Abilene,  Kans.,  1:6-7.   Subsequent 
to  the  drafting  of  this  book,  NASA  published  an  edited  version  of  Glennan's  diary,  The 
Birth  of  NASA:  The  Diary  of  T.  Keith  Glennan,  ed.  J.  D.  Hunley,  NASA  SP-4105 
(Washington,  1993). 

18.  Donlan  interview,  17  Aug.  1990,  p.  9. 

19.  "Glennan  Message,"  pp.  2-3. 

20.  Ibid.,  pp.  3-4. 

21.  Quoted  in  "Glennan  Message,"  pp.  6-7. 

22.  Floyd  L.  Thompson  interview  with  Walter  Bonney,  Hampton,  Va.,  28  July  1975,  tran- 
script, p.  25,  OHC,  LHA;  on  the  history  of  NACA  Wind  Tunnel  No.  1,  see  Hansen, 
Engineer  in  Charge,  pp.  69-72  and  442. 

23.  For  detailed  inventories  of  the  history  of  the  NACA's  wind  tunnels,  see  Roland.  Model 
Research,  2:507-528,  and  Hansen,  Engineer  in  Charge,  pp.  441-478. 

24.  Flight  Log  Book,  1958,  7  Oct.  1958,  Langley  Flight  Log  Book  Collection,  LHA;  next  to 
the  F-101A  entry  are  the  words  "Hubbard's  Sound  Flight."  See  also  Domenic  J.  Maglieri, 
Harvey  H.  Hubbard,  and  Donald  L.  Lansing,  "Ground  Measurements  of  the  Shock- Wave 
Noise  From  Airplanes  in  Level  Flight  at  Mach  Numbers  to  1.4  and  at  Altitudes  to 
45,000  Feet,"  NASA  TN  D-48,  Sept.  1959.    For  information  on  the  variable-electronic 
control  system,  see  Siguard  A.  Sjoberg,   "Flying  Qualities  Associated  with  Electronic 
Flight  Control  Systems,"  NASA  TN-66282,  5-6  Nov.  1958. 

25.  Quoted  in  Charles  Murray  and  Catherine  Bly  Cox,  Apollo:  Race  to  the  Moon  (New  York: 
Touchstone  Books,  1989),  p.  31. 

26.  Ira  H.  Abbott,  "A  Review  and  Commentary  of  a  Thesis  by  Arthur  L.  Levine,  Entitled 
'A   Study  of  the  Major   Policy   Decisions   of  the  National   Advisory   Committee   for 
Aeronautics,'  Dated  1963,"  NASA  HQA,  HHN-35,  Apr.  1964,  p.  195.    The  LHA  also 
preserves  a  copy  of  Abbott's  manuscript. 

27.  H.  J.  E.  Reid  to  Dr.  Hugh  L.  Dryden,   "Transmittal  of  Report  on  Review  of  Langley 
Research  Effort,"  31  Jan.  1958,  A200-1A,  LCF.   A  photocopy  of  this  report  is  also  in  the 
Thompson  Collection,  LHA,  in  the  folder  "Space  Material." 

452 


Notes  for  Chapter  1 

28.  Langley  was  a  strong  supporter  of  the  Dyna-Soar  project.  Inaugurated  by  the  U.S.  Air 
Force  in  Nov.  1957,  the  concept  was  for  the  design  and  operation  of  an  experimental 
manned  hypersonic  glider,  designated  the  X-20,  that  could  fly  out  of  the  atmosphere, 
bounce  in  and  out  of  shallow  earth  orbit  for  at  least  a  large  part  of  a  trip  around 
the  planet,  and  return  home  to  a  runway  landing.     Even  before  the  NACA's  formal 
agreement  in  May  1958  to  support  the  air  force's  manned  military  space  project,  Langley 
had  been  active  in  hypersonic  and  boost-glider  research.    Some  of  its  researchers  and 
facilities  stayed  active  in  support  of  Dyna-Soar  until  the  DOD  canceled  the  project  in 
Dec.   1963.     For  firsthand  insights  into  Langley's  involvement  in  Project  Dyna-Soar, 
see  John  V.   Becker,    "The  Development  of  Winged  Reentry  Vehicles,"    unpublished 
manuscript,  May  1983,  pp.  35-51,  copy  in  the  LHA.  For  a  brief  critical  examination  of  the 
Dyna-Soar 's  controversial  history,  see  McDougall,  Heavens  and  the  Earth,  pp.  339-341. 

29.  In  midsummer  1958,  Stack  visited  the  Vickers  organization  in  Weybridge,  England,  to 
review  a  British  proposal  for  the  Swallow  arrow-wing  aircraft.  The  proposal  consisted  of 
a  plan  for  a  25,000-pound  research  airplane  derived  from  the  Swallow  configuration  that 
would  be  the  progenitor  of  a  supersonic  commercial  transport.  Stack  made  several  trips 
to  Europe  in  the  late  1950s  and  early  1960s  as  an  NACA/NASA  representative  to  NATO 
meetings  concerning  the  Mutual  Weapons  Defense  Program  (MWDP).    For  information 
on  the  Swallow  and  Stack's  trips  to  Europe,  see  the  miscellaneous  material  in  the  folders 
marked  "MWDP,"  in  section  6  of  the  John  Stack  Collection,  LHA. 

30.  James  R.  Hansen,  "Transition  to  Space:  A  History  of  'Space  Plane'  Concepts  at  Langley 
Aeronautical  Laboratory  1952-1957,"   Journal  of  the  British  Interplanetary  Society  40 
(1987):  67-80.  On  the  X-15,  see  Richard  P.  Hallion,  On  the  Frontier,  pp.  101-129,  and 
Milton  O.  Thompson,  At  the  Edge  of  Space:  The  X-15  Flight  Program  (Washington  and 
London:  Smithsonian  Institution  Press,  1992). 

31.  Thompson  interview  with  Bonney,  28  July  1975,  p.  20. 

32.  Ibid. 

33.  Murray  and  Cox,  Apollo,  p.  28. 

34.  Roland,  Model  Research,  1:135-137;  Hansen,  Engineer  in  Charge,  pp.  145-146. 

35.  Caldwell  Johnson  quoted  in  Murray  and  Cox,  Apollo,  p.  28. 

36.  For  an  appraisal  of  Langley's  research  environment  in  the  early  years,   see  Hansen, 
Engineer  in  Charge,  pp.  33-40.  For  an  insightful  autobiographical  look  into  the  freedoms 
of  NACA  laboratory  research,  read  John  V.  Becker,    The  High-Speed  Frontier:    Case 
Histories  of  Four  NACA  Programs,    1920-1950,   NASA  SP-445  (Washington,   1980). 
Becker,  who  came  to  work  at  Langley  in  1936,  did  pioneering  research  in  high-speed 
aerodynamics  culminating  in  major  contributions  to  the  X-15  program. 

37.  Murray  and  Cox,  Apollo,  p.  27. 

38.  In  essence,    the  deal  with  the  Selective  Service  System  called  for  the  induction  of 
eligible  Langley  employees  into  the  Army  Air  Corps  Enlisted  Reserves.  However,  NACA 
inductees  received  no  military  training,  never  wore  a  uniform,  and  spent  absolutely  no 
time  on  active  duty.  For  the  details  of  this  arrangement,  see  Hansen,  Engineer  in  Charge, 
pp.  203-205. 


453 


Notes  for  Chapter  2 

Chapter  2 
The  First  NASA  Inspection 

1.  As  had  been  the  case  for  the  NACA's  annual  manufacturers'  conferences,  Langley's  plan- 
ning for  NASA's  First  Anniversary  Inspection  was  complete.    The  typically  exhaustive 
thoroughness  of  preinspection  planning  is  evident  in  both  the  abundance  and  the  detail  of 
the  extant  archival  material.  This  material  includes  several  relevant  internal  memoranda, 
including  an  "Operation  and  Policy"  statement;  all  purchase  requests;  complete  lists  of 
the  invited  guests  plus  copies  of  the  invitations;  all  programs  and  booklets;  representative 
visitor  identification  badges;  luncheon  menus;  maps  of  tour-bus  routes;  as  well  as  group 
photos,  sheets  with  daily  attendance  figures,  selected  press-coverage  clippings,  and  all 
NASA  press  releases.  In  the  LHA  are  five  large  folders  of  material  for  the  1959  inspec- 
tion. They  are  located,  along  with  the  records  for  the  1964  NASA  inspection  (also  held 
at  Langley),  in  a  special  cabinet  drawer  labeled  "1959  Inspection-1964  Inspection."  Ma- 
terial on  the  1959  inspection  can  also  be  found  in  E6-1E,  LCF.   Langley's  correspondence 
files  are  located  in  the  LHA. 

2.  "The  Nice  NASA  Show  for  the  People,"  Newport  News  (Va.)  Daily  Press,  27  Oct.  1959. 

3.  "Nearly  20,000  at  Open  House;  Public  Pays  First  Visit  to  NASA,"  Newport  News  Daily 
Press,  25  Oct.  1959. 

4.  "Nice  NASA  Show,"  Daily  Press. 

5.  Axel  T.   Mattson,    "Synopsis  of  the  NASA  First   Anniversary  Inspection,    Technical 
Presentations,"    6  July   1959,   p.   1,   in   1959  inspection  folder  labeled    "NASA   'First 
Anniversary  Inspection'  LRC-1959  Policy  Statement,"    in  the   "1959  Inspection-1964 
Inspection"  drawer. 

6.  T.  Melvin  Butler,  Langley  administrative  management  officer,   "Memorandum  for  All 
Concerned;  Subject:  Special  Duties  for  the  1959  NASA  Inspection  at  Langley  Research 
Center,"  5  Oct.  1959,  and  "Information  for  Group  Leaders  and  Assistant  Group  Leaders, 
NASA  1959  Inspection,"  9  Oct.  1959.  Both  memoranda  are  in  the  "1959  Inspection  File" 
in  the  LHA.  For  in-house  news  stories  (with  photographs)  on  the  1959  inspection,  see 
the  16  Oct.  1959  issue  of  the  Langley  Air  Scoop.  For  outside  technical  reviews,  consult 
the  19  Oct.,  26  Oct.,  and  2  Nov.  1959  issues  of  Aviation  Week.  An  editorial  team  from 
Aviation  Week  covered  all  four  days  of  the  NASA  inspection  and  published  several  articles 
on  NASA's  research  highlights.    Most  of  these  articles,  however,  were  based  largely  on 
NASA's  press  releases  prior  to  the  inspection. 

7.  See  folder  entitled  "NASA  'First  Anniversary  Inspection'  LRC-1959  Policy  Statement" 
for  correspondence  between  Langley  and  the  various  NASA  centers  regarding  their 
participation  in  the  1959  inspection.    See  also  the  relevant  NASA  press  releases  in  the 
"Press  Releases"  folder  for  the  1959  inspection.  On  the  cancellation  of  the  Vega  program, 
see  "NASA  Kills  Vega,  Adopts  USAF  Agena."  Aviation  Week,  21  Dec.  1959,  and  "Vega 
Out,  Hopes  Pinned  on  Centaur,"  Missiles  and  Rockets,  21  Dec.  1959. 

8.  Axel  Mattson  interview  with  author,  14  Aug.  1989,  Hampton,  Va.,  transcript,  pp.  9-10, 
OHC,  LHA.  At  the  time,  Mattson  was  the  assistant  chief  of  Langley's  Full-Scale  Research 
Division. 

9.  Mattson  interview,  14  Aug.  1989,  pp.  9-10. 
10.  Ibid. 

454 


Notes  for  Chapter  2 

11.  Ibid.,  pp.  2-5  and  11.    See  also  Axel  Mattson  to  the  associate  director.    The  Langley 
correspondence  files,  E6-1E,  LCF,  contain  several  letters  from  Mattson  to  Langley  and 
NASA  headquarters  concerning  his  visits  to  JPL,  Goddard,  and  other  NASA  centers  in 
preparation  for  the  1959  inspection. 

12.  Mattson  interview,  14  Aug.  1989,  p.  5. 

13.  Charles  J.  Donlan  interview  with  author,  Hampton,  Va.,  17  Aug.  1990,  transcript,  p.  16, 
OHC,  LHA. 

14.  Mattson  interview,  14  Aug.  1989,  p.  4. 

15.  Smith  J.  DePrance  to  Dr.  H.  J.  E.  Reid,  29  Oct.  1959,  E6-1E,  LCF. 

16.  H.  J.  E.  Reid  to  Dr.  Smith  J.  DeFrance,  30  Oct.  1959,  E6-1E,  LCF. 

17.  Englishman  John  Hodge  had  joined  the  STG  in  the  spring  of  1959  along  with  30  other 
engineers  who  had  been  handpicked  by  the  STG  after  they  had  been  laid  off  by  the 
AVRO  (A.  V.  Roe)  aircraft  corporation  in  Canada;  a  new  Conservative  government  in 
Ottawa  had  canceled  a  project  for  the  building  of  a  technologically  innovative  supersonic 
interceptor,  the  CF-105  Arrow,  and  some  swift  action  by  NASA  had  allowed  the  space 
agency  to  acquire  some  exceedingly  talented  people.    Hodge  became  the  second  NASA 
flight  director — Christopher  C.  Kraft's  deputy — at  Mission  Control  in  Houston.   It  was 
his  misfortune  to  be  the  flight  director  on  duty  during  the  tragic  Apollo  1  fire  that  killed 
astronauts  Roger  Chaffee,  Edward  White,  and  Virgil  Grissom  in  Jan.  1967.  Subsequent 
to  the  Apollo  program,  Hodge  served  as  director  of  NASA's  Space  Station  Task  Force. 
Windier  also  became  a  Houston  flight  director,  and  Huss,  too,  served  in  Mission  Control. 
On  the  recruitment  of  the  Canadians  into  NASA's  space  program,  see  Loyd  S.  Swenson, 
Jr.,  James  M.  Grimwood,  and  Charles  C.  Alexander,   This  New  Ocean:    A  History  of 
Project  Mercury,  NASA  SP-4201  (Washington,  1966),  p.  153,  as  well  as  Charles  Murray 
and  Catherine  Ely  Cox,  Apollo:  The  Race  to  the  Moon  (New  York:  Touchstone  Books, 
1989),  pp.  21-23.  See  also  Donlan  interview,  17  Aug.  1990,  p.  21. 

18.  Among  the  Langley  documents  relevant  to  the  establishment  of  the  STG,  see  especially 
Paul  E.  Purser  to  Robert  R.  Gilruth,   "Initial  Hiring  of  Personnel  for  Space  Center," 
3  Sept.  1958;  and  Robert  R.  Gilruth  to  Langley  Associate  Director  Floyd  L.  Thompson, 
"Space  Task  Group,"  3  Nov.  1958.  Both  letters  can  be  found  in  the  Floyd  L.  Thompson 
Collection,   LHA.  On  the  formation  of  the  STG,  see  also  Gilruth's   "Memoir:    From 
Wallops  Island  to  Mercury,  1945-1958,"  unpublished  manuscript  presented  at  the  Sixth 
International  History  Symposium,  Vienna,  Austria,  13  Oct.  1972,  pp.  39-42,  copy  in 
LHA.   For  an  excellent  history  of  NASA's  Project  Mercury,  see  Swenson,  Grimwood,  and 
Alexander,  This  New  Ocean. 

19.  STG  script  for  "NASA  1959  Inspection:   Project  Mercury,"  Oct.  1959,  p.  1.   All  of  the 
scripts  for  presentations  given  at  the  1959  inspection  are  in  the  folder  labeled  "Press 
Releases,  1959  Inspection,"  in  the  1959-1964  inspection  drawer  in  the  LHA. 

20.  Donlan  interview,  17  Aug.  1990,  pp.  29-30. 

21.  Swenson,  Grimwood,  and  Alexander,   This  New  Ocean,  pp.  237-238.    See  also  John  A. 
(Shorty)  Powers'  memorandum  for  STG  files,  "Response  to  queries  about  Astronauts' 
personal  stories,"  27  Aug.  1959,  C136-1,  LCF. 

22.  STG  script,  "Project  Mercury,"  pp.  1-2. 


455 


Notes  for  Chapter  2 

23.  Gilruth,  "Memoir,"  p.  45.  On  NASA's  selection  of  the  Mercury  astronauts,  see  Swenson, 
Grimwood,  and  Alexander,   This  New  Ocean,  pp.  129-131;  also  Mae  Mills  Link,  Space 
Medicine  in  Project  Mercury,  NASA  SP-4003  (Washington,  1965),  pp.  44-46. 

24.  By  this  time  Bob  Champine,  one  of  Langley's  most  talented  test  pilots,  had  already 
experienced  the  high  G- forces  of  the  navy's  centrifuge  at  Johnsville,  Pa.,  and  had  also 
ridden  several  of  the  Project  Mercury  simulators,  so  he  knew  what  NASA  astronauts 
must  endure.  See  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  pp.  43  and  94. 
For  some  thoughts  on  the  tortures  of  the  wheel  from  the  perspective  of  a  Gemini  and 
Apollo  astronaut,  see  Michael  Collins,  Liftoff:  The  Story  of  America's  Adventure  in  Space 
(New  York:  NASA/Grove  Press,  1988),  pp.  30-31. 

25.  Yeager  quoted  in  Collins,  Liftoff,  p.  46. 

26.  In  the  LHA,  I  have  left  a  folder  with  photocopies  of  dozens  of  local  newspaper  stories 
about  the  Mercury  astronauts.     The  chief  local  reporter  covering  the  astronauts  was 
Virginia  Biggins,  columnist  for  the  Newport  News  Times  Herald.  In  her  column,  "Point 
of  View,"   Ms.  Biggins  often  covered  the  personal  lives  of  the  astronauts,  in  spite  of 
the  notoriously  restrictive  Time-Life  contract.  For  some  fascinating  recollections  of  the 
Project  Mercury  astronauts  and  their  days  in  the  Hampton  Roads  area,  see  the  transcript 
of  my  interview  with  Ms.  Biggins,  dated  1  Aug.  1990,  OHC,  LHA;  for  comments  on  how 
she  was  able  to  work  around  the  Time-Life  restrictions,  see  pp.  4-5.  A  Langley  videotape 
of  a  presentation  made  by  Ms.  Biggins  in  July  1991  at  the  former  Langley  Visitors'  Center 
on  her  recollections  of  the  days  when  the  astronauts  were  training  at  Langley  is  available 
through  the  Langley  Office  of  Public  Services. 

27.  See  Robert  Voas,  "Project  Mercury  Astronaut  Training  Program,"  unpublished  paper 
presented   at  the  Symposium   on   Psychophysiological   Aspects  of  Space   Flight,    San 
Antonio,  Tex.,  26-27  May  1960,  copy  in  Langley  Technical  Library.    Dr.  Voas  (a  navy 
lieutenant  with  a  Ph.D.  in  psychology)  served  on  the  NASA  astronaut  selection  committee 
and  became  a  formal  member  of  the  organization  within  the  STG  that  evaluated,  tested, 
and  trained  the  Mercury  astronauts.    For  a  virtually  complete  bibliography  of  Voas's 
contemporary  papers  and  presentations  on  Mercury  astronaut  training,   consult  the 
NACA/NASA  card  file  in  the  Langley  Technical  Library.    For  a  historical  analysis  of 
Mercury  astronaut  selection  and  evaluation,  as  well  as  the  role  of  Dr.  Voas  and  others  in 
it,  see  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  pp.  159-165. 

28.  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  pp.  237-238. 

29.  See  STG  script,  "Project  Mercury,"  pp.  3-5.   John  Glenn  is  quoted  in  Collins,  Liftoff, 
p.  31. 

30.  Some  readers  may  be  surprised  by  this  reference  to  a  "sidearm  controller"  for  the  X-15 
"space  plane"  and  Mercury  spacecraft  simulator  if  they  associate  the  original  development 
of  sidearm-control  technology  with  the  much  ballyhooed  and  supposedly  altogether  new 
electronic  flight  control  systems  of  the  1980s.  The  history  of  the  sidearm  controller  goes 
back  at  least  to  German  research  into  new  flight  control  systems  for  their  advanced  fighter 
aircraft  of  World  War  II.     For  those  interested  in  this  specialized  technical  topic,  see 
the  short  bibliography  compiled  by  engineer  Paul  E.  Hunt  of  Langley's  8-Foot  Tunnels 
Branch  in  1959,  on  early  (post-World  War  II)  American  research  into  sidearm  flight 
controls,  NACA/NASA  card  file,  Langley  Technical  Library. 

31.  STG  script,  "Project  Mercury,"  pp.  1-4. 

456 


Notes  for  Chapter  3 

32.  Ibid.,  pp.  1  and  3. 

33.  The  phrase  that  the  STG  treated  the  Mercury  astronauts  as   "active  and  valuable 
participants"   is  from  Collins,   Liftoff,   p.  48.     Mike  Collins  was  not  part  of  Project 
Mercury,  but  as  an  astronaut  in  the  follow-up  Gemini  and  Apollo  programs,  he  became 
very  familiar  with  the  experiences,  good  and  bad,  of  the  original  seven  astronauts.  For 
Gilruth's  experiences  in  flight  testing,  see  the  early  sections  of  his  "Memoir"  as  well  as  the 
analysis  in  my  Engineer  in  Charge,  NASA  SP-4305  (Washington,  1987),  pp.  262-270  and 
275-278.    Gilruth's  systematic  and  quantified  approach  to  aircraft  flight  testing  is  also 
featured  prominently  in  chap.  3  of  Walter  G.  Vincenti,  What  Engineers  Know  and  How 
They  Know  It:  Analytical  Studies  from  Aeronautical  History  (Baltimore:  Johns  Hopkins 
University  Press,   1990);   chap.  3  is  entitled,   "Establishment  of  Design  Requirements: 
Flying-Quality  Specifications  for  American  Aircraft,  1918-1943." 

34.  "Capsule  Built  at  Langley  and  Lewis  Launched  Atop  Atlas  in  Mercury  Test,"  Langley  Air 
Scoop,  11  Sept.  1958.  On  the  history  of  Big  Joe,  see  Swenson,  Grimwood,  and  Alexander, 
This  New  Ocean,  pp.  125-128. 

35.  STG  script,  "Project  Mercury,"  pp.  4-5. 

36.  Gilruth,   "Memoir,"  p.  46.    On  the  history  of  Little  Joe,  see  Swenson,  Grimwood,  and 
Alexander,  This  New  Ocean,  esp.  pp.  124-126,  208-213,  and  291-294. 

37.  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  pp.  208-209. 

38.  Ibid.,  pp.  210-213.    For  a  contemporary  news  magazine  story,  see  "Little  Joe  Launch 
Tests  Mercury  Capsule  Escape,"  Aviation  Week,  14  Dec.  1959.  For  a  former  astronaut's 
thoughtful  summary  of  the  role  played  by  chimpanzees  and  other  primates  in  Project 
Mercury,  see  Collins,  Liftoff,  pp.  48  and  51-52. 

39.  Gilruth,  "Memoir,"  p.  43. 

Chapter  3 
Carrying  Out  the  Task 

1.  Robert  R.  Gilruth,  "Memoir:  From  Wallops  Island  to  Mercury,  1945-1958,"  unpublished 
manuscript  presented  at  the  Sixth  International  History  Symposium,  Vienna,  Austria, 
13  Oct.  1992,  p.  38,  copy  in  LHA.  In  the  OHC,  LHA,  is  a  barely  audible  copy  of  my 
tape-recorded  interview  with  Gilruth  at  his  home  at  Kilmarnock,  Va.,  10  July  1986.  This 
interview,  which  lasted  for  over  three  hours,  covers  important  aspects  of  Gilruth's  entire 
NACA/NASA  career. 

On  Presidential  Science  Adviser  Kistiakowsky  and  ARPA's  contrasting  positions  on 
the  early  space  program,  see  Walter  A.  McDougall,  The  Heavens  and  the  Earth:  A 
Political  History  of  the  Space  Age  (New  York:  Basic  Books,  1985),  pp.  202-205.  On 
ARPA  specifically,  see  pp.  195-198  and  201-206. 

2.  Gilruth,  "Memoir,"  pp.  34-39;  Loyd  S.  Swenson,  Jr.,  James  M.  Grimwood,  and  Charles  C. 
Alexander,  This  New  Ocean:  A  History  of  Project  Mercury,  NASA  SP-4201  (Washington, 
1966),  pp.  110-116;  Michael  Collins,  Liftoff:  The  Story  of  America's  Adventure  in  Space 
(New  York:  NASA/Grove  Press,  1988),  p.  28. 

3.  Charles  Zimmerman  interview  with  author,   1  Aug.   1990,  Hampton,  Va.,  transcript, 
pp.  5-6,  OHC,  LHA. 

4.  Gilruth,  "Memoir,"  pp.  38-39. 

457 


Notes  for  Chapter  3 

5.  The  key  technical  paper  was  H.  J.  Allen  and  A.  J.  Eggers,  Jr.,  "A  Study  of  the  Motion 
and  Aerodynamic  Heating  of  Ballistic  Missiles  Entering  the  Earth's  Atmosphere  at  High 
Supersonic  Speeds,"  NACA  Conf.  Research  Memo.  A53D28,  Aug.  1953.  Five  years  later, 
in  1958,  the  NACA  issued  updated  versions  of  the  original  report  as  NACA  TN-4047 
and  NACA  TR-1381.    For  historical  analysis  of  the  blunt-body  concept,  see  Edwin  P. 
Hartmann,  Adventures  in  Research:   A  History  of  Ames  Research  Center,   1940-1965, 
NASA  SP-4302  (Washington,  1970),  p.  218;  Elizabeth  A.  Muenger,  Searching  the  Hori- 
zon:   A  History  of  Ames  Research  Center,   1940  1976,  NASA  SP-4304  (Washington, 
1985),  pp.  66-68;  Alex  Roland,  Model  Research:   The  National  Advisory  Committee  for 
Aeronautics,  1915-1958,  2  vols.,  NASA  SP-4103  (Washington,  1985),  1:285-286;  Collins, 
Liftoff,  p.  24;  and  James  R.  Hansen,  Engineer  in  Charge:  A  History  of  Langley  Aero- 
nautical Laboratory,  1917-1958,  NASA  SP-4305  (Washington,  1987),  pp.  349-350  and 
358-361. 

6.  PARD  engineers  J.  Thomas  Markley  and  Max  Faget  are  quoted  in  Charles  Murray  and 
Catherine  Ely  Cox,  Apollo:  The  Race  to  the  Moon  (New  York:  Touchstone  Books,  1989), 
pp.  39-40.    Faget  analyzed  the  Mercury  capsule  design  in  "Mercury  Capsule  and  Its 
Flight  Systems,"  Aero/Space  Engineering  19  (Apr.  1960):  48-53.  Charles  Donlan  speaks 
highly  of  Faget  and  his  contributions  to  the  Mercury  spacecraft  design:    "Max  was  a 
very  ingenious  fellow.    He  had  some  novel  and  sound  ideas,  for  the  most  part,  a  good 
engineer.  The  original  ballistic  capsule  approach  came  out  of  Harvey  Allen's  work.  But  as 
a  simple,  quick  way  to  get  the  job  done,  I  think  the  Mercury  capsule  came  [from]  Max,  and 
[from]  Caldwell  Johnson."   Donlan  interview  with  author,  Hampton,  Va.,  17  Aug.  1990, 
transcript,  p.  28,  OHC,  LHA.  Langley  veteran  Charles  Zimmerman  seconds  Donlan's 
appraisal  of  Faget:    "I  don't  think  Max  Faget  has  ever  gotten  the  credit  he  deserved  in 
this  project."    Zimmerman  interview  with  author,  1  Aug.  1990,  p.  14.   Many  references 
to  Faget's  contributions  to  Project  Mercury  may  be  found  in  Swenson,  Grimwood,  and 
Alexander,  This  New  Ocean. 

7.  The  NACA  published  the  Faget-Garland-Buglia  paper  on  nonlifting,  wingless  satellites  in 
NACA  Conference  on  High-Speed  Aerodynamics,  Ames  Aeronautical  Laboratory,  Moffett 
Field,  Calif.,  March  18,  19,  and  20,  1958:  A  Compilation  of  Papers  Presented  (NACA: 
Washington,  1958),  p.  19ff. 

8.  Gilruth,  "Memoir,"  p.  32. 

9.  Ibid.,  p.  33.    It  is  easy  to  confuse  Matthews'  concept  with  the  winged  reentry  vehicle 
concepts  championed  by  Langley's  hypersonics  specialist  John  V.  Becker,  as  the  two 
researchers'  theoretical  findings  were  combined  for  various  analytical  purposes  beginning 
in  1957.   For  detailed  information  on  Becker's  early  space  plane  concepts,  see  his  auto- 
biographical "The  Development  of  Winged  Reentry  Vehicles,  1952-1963,"  23  May  1983. 
A  copy  of  this  manuscript  is  preserved  in  the  LHA.    Becker's  pioneering  ideas  for  hyper- 
sonic gliders  and  transatmospheric  vehicles  are  featured  prominently  in  Hansen,  Engineer 
in  Charge,  esp.  pp.  367-373  and  377-381.  On  Becker's  winged  manned  satellite  proposal 
in  1957-1958,  see  also  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  p.  89. 

10.  Gilruth,  "Memoir,"  p.  34.    John  V.  Becker  offers  what  might  be  interpreted  as  a  mild 
dissent  from  this  consensus  in  "The  Development  of  Winged  Reentry  Vehicles,"  p.  34. 

11.  Gilruth  quoted  in  Collins,  Liftoff,  p.  61. 


458 


Notes  for  Chapter  3 

12.  Gilruth  interview,  10  July  1986.  See  also  Gilruth,  "Memoir,"  p.  41;  and  Donlan  interview, 
17  Aug.  1990,  p.  7. 

13.  Jack  C.  Heberlig  interview  with  Robert  B.  Merrifield,  Oct.  1967,  quoted  in  comment 
draft  of  Dr.  Merrifield 's  history  of  the  early  years  of  NASA's  Manned  Spacecraft  Center 
in  Houston,  Tex.,  dated  16  Feb.  1970,  chap.  3,  p.  11.  Merrifield's  unpublished  manuscript 
provides  a  thorough  and  thoughtful  account  of  the  history  of  the  STG  and  its  move  from 
Virginia  to  Texas.    A  copy  of  the  comment  draft  of  his  manuscript  is  preserved  in  the 
LHA.    I  do  not  know  why  Merrifield's  work,  which  was  sponsored  by  NASA,  was  never 
published. 

14.  Robert  R.  Gilruth  to  Associate  Director  Floyd  L.  Thompson,    "Space  Task  Group," 
3  Nov.  1958.  A  copy  of  this  document  can  be  found  in  the  Floyd  L.  Thompson  Collection 
in  the  LHA.  With  the  memorandum  is  a  cover  sheet  with  a  handwritten  note  from 
Thompson  to  Administrative  Chief  Officer  T.  Melvin  Butler  saying  "For  prompt  action. 
Important — Do  Not  Destroy."     See  also  the  earlier  document  from  Langley's  Paul  E. 
Purser  to  Gilruth,  "Initial  hiring  of  personnel  for  Space  Center,"  3  Sept.  1958,  also  in  the 
Thompson  Collection.    Both  memoranda  are  in  the  folder  labeled  "Space  Task  Group, 
Formulation." 

15.  Donlan  interview,  17  Aug.  1990,  p.  22. 

16.  Ibid.,  p.  7. 

17.  This  quote,  from  an  anonymous  Langley  engineer  who  was  not  part  of  the  STG,  is  from 
Murray  and  Cox,  Apollo,  p.  30. 

18.  On  Langley's  PARD,  see  Gilruth,  "Memoir,"  pp.  3-28;  former  Langley  engineer  Joseph 
A.  Shortal's  exhaustive  A  New  Dimension:     Wallops  Island  Flight   Test  Range,    The 
First  Fifteen  Years,  NASA  RP-1028  (Washington,  1978);  as  well  as  Hansen,  Engineer 
in  Charge,  esp.  pp.  269-270. 

On  the  air  force's  proposed  Project  MISS,  see  McDougall,  Heavens  and  the  Earth, 
p.  197;  and  Swenson,  Grimwood,  and  Alexander,  This  New  Ocean,  p.  93. 

19.  Murray  and  Cox,  Apollo,  p.  32. 

20.  Donlan  interview,  17  Aug.  1990,  pp.  29-30. 

21.  Charles  Zimmerman  interview  with  Walter  Bonney,  30  Mar.  1973,  transcript,  p.  24,  OHC, 
LHA.    See  also  Zimmerman  interview  with  author,  1  Aug.  1990,  pp.  56  and  11-13. 

22.  The  Langley  division  chief  is  quoted  in  Murray  and  Cox,  Apollo,  p.  31.  The  authors  do 
not  identify  him. 

23.  Murray  and  Cox,  Apollo,  p.  31. 

24.  Ibid. 

25.  Floyd  L.  Thompson,  "Comments  on  Draft  Chapters  I-IV  of  MSC  [Manned  Spacecraft 
Center]  History  [Robert  B.  Merrifield's  manuscript,  see  n.  13],  Dated  February  16,  1970," 
27  Mar.  1970,  p.  7.  Copies  of  this  memorandum  to  Dr.  Eugene  Emme,  NASA  HQA,  can 
be  found  in  the  Thompson  Collection,  LHA,  in  a  folder  labeled  "MSC  History  Comments" 
and  in  El-3,  LCF.   At  the  time  of  his  "Comments,"  Thompson  was  special  consultant  to 
the  NASA  administrator. 

26.  For  complete  internal  files  documenting  NASA  Langley's  support  of  the  STG  and  Project 
Mercury,  see  the  correspondence  in  A189-5,  LCF,  and  in  the  LCF's  even  more  voluminous 
collection  of  letters  filed  under  the  project  name  "Mercury."    See  also  the  material  in 


459 


Notes  for  Chapter  3 

Thompson's  personal  collection  of  "Project  Mercury"  papers,  now  preserved  as  part  of 
the  LHA's  Thompson  Collection. 

27.  Donlan  interview,  17  Aug.  1990,  pp.  32  33. 

28.  Collins,  Liftoff,  p.  150.    Liftoff  is  an  excellent  "insider"  history,  while  Collins'  Carrying 
the  Fire:  An  Astronaut's  Journey  (New  York:  Farrar,  Straus,  and  Giroux,  1974)  is  the 
best  autobiography  yet  to  come  from  anyone — astronaut  or  otherwise — involved  in  the 
space  program. 

29.  Murray  and  Cox,  Apollo,  p.  254.   On  ground  control  of  flight  tests  at  Edwards  AFB  in 
the  late  1940s  and  1950s,  see  Richard  P.  Hallion,  On  the  Frontier:   Flight  Research  at 
Dryden,  1946-1981,  NASA  SP-4303  (Washington,  1984),  pt.  1. 

30.  Murray  and  Cox,  Apollo,  p.  255. 

31.  William  J.  Boyer  interview  with  author,  Hampton,  Va.,  24  Aug.  1990.  This  tape-recorded 
interview  has  not  been  transcribed,  but  the  tape  is  available  in  the  OHC,  LHA. 

32.  NASA  Langley  press  release,  "NASA  Langley  Research  Center  Directs  Establishment  of 
$80  Million  Tracking  System  for  Project  Mercury,"  16  Aug.  1961.  For  analysis  and  more 
information  on  the  tracking  network,  see  Swenson,  Grimwood,  and  Alexander,  This  New 
Ocean,  esp.  pp.  213-221. 

33.  Murray  and  Cox,  Apollo,  p.  255. 

34.  Boyer  interview,  24  Aug.  1990.  On  Goddard's  responsibility  for  Minitrack,  see  This  New 
Ocean,  p.  146;  and  Alfred  Rosenthal,  Venture  into  Space:  Early  Years  of  Goddard  Space 
Flight  Center,  NASA  SP-4301  (Washington,  1968),  pp.  65-66. 

35.  Boyer  interview,  24  Aug.  1990. 

36.  Paul  Vavra  interview  with  Robert  B.   Merrifield,   quoted  in  Merrifield's  manuscript, 
chap.  3,  p.  39  n.  74. 

37.  See  the  Western  Electric  Company's  Final  Progress  Report  to  NASA:  Project  Mercury, 
Contract  Number  NAS  1-430  (June  1961),  copy  in  Project  Mercury  files,  LHA. 

38.  Boyer  interview,  24  Aug.  1990. 

39.  Edmond  C.  Buckley,  director,  Tracking  and  Data  Acquisition,  NASA  headquarters,  to 
Floyd  L.  Thompson,  director,  NASA  Langley,  27  Feb.   1962,  E1-2C,  LCF.     Another 
copy  of  this  letter  is  in  the  folder  labeled  "Mercury  Tracking  Range"  that  is  part  of  the 
Thompson  Collection  in  the  LHA. 

40.  Ibid. 

41.  Gilruth,  "Memoir,"  p.  47.  Reflecting  on  Thompson's  agreement  with  Gilruth,  Caldwell 
Johnson  has  jested  that  all  STG  members  to  this  day  "wonder  which  half  they  were  in," 
the  half  Gilruth  wanted  or  the  half  Thompson  wanted  to  give  away. 

42.  Thompson,  "Comments  on  Draft  Chapters,"  p.  7. 

43.  Donlan  interview,  17  Aug.  1990,  p.  17. 

44.  Thompson,  "Comments  on  Draft  Chapters,"  pp.  7-8. 

45.  Laurence  K.  Loftin,  Jr.,  interview  with  author,  Newport  News,  Va.,  26  June  1990,  OHC, 
LHA. 

46.  Donlan  interview,  17  Aug.  1990,  pp.  27-28. 

47.  Thompson,  "Comments  on  Draft  Chapters,"  p.  8. 

48.  Merrifield's  manuscript,  chap.  3,  pp.  61-64.  See  also  Swenson,  Grimwood,  and  Alexander, 
This  New  Ocean,  pp.  251-253.  The  official  moves  to  lessen  STG's  dependence  on  Langley 
can  be  followed  in  the  Langley  correspondence  files:  Code  A189-5  and  "Mercury." 

460 


Notes  for  Chapter  3 

49.  On   the   difficult   genesis   of  Goddard   Space   Flight    Center,    see   Rosenthal,    Venture 
into  Space,  pp.  27-35,  and  Robert  L.  Rosholt,  An  Administrative  History  of  NASA, 
1958-1963,  NASA  SP-4101  (Washington,  1966),  p.  79. 

50.  Donlan  interview,  17  Aug.  1990,  p.  15. 

51.  Ibid.,  pp.  16-17. 

52.  Swenson,    Grimwood,    and   Alexander,    This  New   Ocean,    pp.   390-392;    Murray   and 
Cox,  Apollo,  pp.  130-132;  McDougall,  Heavens  and  the  Earth,  pp.  373-374;  Merrifield 
manuscript,  chap.  4,  pp.  21-33. 

53.  Joe  Stein,  NASA  headquarters  news  release,  "Manned  Space  Flight  Laboratory  Location 
Study  Completed,"    19  Sept.   1961.     Constant  rumors  of  the  STG's  impending  move 
appeared  in  the  Hampton  Roads  newspapers  for  weeks;  see,  for  example,  Bill  Delany, 
"Mercury  Unit  Move  Rumors  Cited  by  NASA  as  'Unofficial  Reports,'  "  Newport  News 
(Va.)  Daily  Press,  30  May  1961. 

54.  Editorial,    "A    Terrible   Waste   of  Time   and   Money,"    Newport    News    Daily   Press, 
27  June  1961.     So  disturbed  were  members  of  the  Peninsula  Chamber  of  Commerce 
(PCC)  that  they  launched  a  campaign  of  letter  writing,  petitioning,  and  lobbying  to 
keep  the  STG  where  it  was.  Leading  the  "call  to  arms,"  was  a  special  subcommittee  of 
the  PCC's  legislative  committee  whose  members  worked  in  conjunction  with  the  efforts 
of  home  district  Congressman  Thomas  N.  Downing.    Members  of  this  "Save  the  STG" 
subcommittee  included  local  newspaper  magnate  William  R.  Van  Buren;  prominent  real 
estate  broker  John  P.  Yancey  (the  subcommittee's  chairman);  Charles  K.  Hutchens,  del- 
egate to  the  Virginia  General  Assembly;  Mayor  O.  J.  Brittingham,  Jr.,  of  Newport  News; 
Mayor  George  C.  Bentley  of  Hampton;  Mayor  G.  S.  Forrest  of  Poquoson;  Rodger  Smith 
of  the  York  County  Board  of  Supervisors;  Leslie  O'Hara,  a  PCC  director;  Frank  Floyd, 
co-manager  of  the  Peninsula  Shipbuilders  Association;  and  Irving  L.  Fuller,  the  PCC's 
executive  vice-president.    The  local  newspapers,  besides  staying  on  top  of  rumors  and 
NASA  statements  about  the  STG's  relocation,  reported  regularly  on  the  actions  of  this 
special  committee.    One  can  follow  the  outline  of  the  losing  battle  to  keep  the  STG  at 
Langley  from  the  late  spring  to  the  early  fall  of  1961  in  articles  from  the  Newport  News 
Daily  Press  from  1  June  through  21  Sept. 

55.  Caldwell  Johnson  quoted  in  Murray  and  Cox,  Apollo,  p.  131. 

56.  Donlan  interview,  17  Aug.  1990,  p.  39. 

57.  Ibid.,  pp.  39-40. 

58.  Ibid.,  p.  23. 

59.  See   the   day-to-day   correspondence   from   this   period   pertaining   to   STG    activities 
in  A189-5,   LCF.   For  the  details  of  the  Glenn  and  Carpenter  flights,   see  Swenson, 
Grimwood,  and  Alexander,    This  New  Ocean,  pp.  422-436  (Glenn's  Mercury-Atlas  6 
flight  in  Friendship  7,  20  Feb.  1962)  and  446-460  (Carpenter's  Mercury-Atlas  7  flight 
in  Aurora  7,  24  May  1962).    For  a  more  popular  and  abbreviated  history,  see  Collins, 
Liftoff,  pp.  52-57. 

60.  This  letter  was  quoted  at  length  in  "Gilruth  Thanks  Langley  Staff  for  Support  of  Project 
Mercury,"  Langley  Researcher,  5  July  1963. 

61.  My  account  is  derived  from  the  following  articles  in  the  Newport  News  Daily  Press: 
"Saturday's  Tribute  to  Astronauts  Set,"   13  Mar.  1962;   "Sen.  Byrd  Accepts  Invitation 
to  Program  Honoring  Space  'Team,'  Parade  Route  Set,"   14  Mar.  1962;   "State  to  Pay 

461 


Notes  for  Chapter  4 

Tribute  to  Project  Mercury  Space  'Team'  Today,"  17  Mar.  1962;  "Seven  Astronauts  Get 
Rousing;  Peninsula  Throng  Hails  Project  Mercury  Team,"  18  Mar.  1962;  "Respect,  Awe 
Mix  with  Pride  as  Peninsula  Pays  Its  Respects  to  America's  Astronauts,"  18  Mar.  1962. 
62.  Parke  Rouse,  Jr.,  The  Good  Old  Days  in  Hampton  and  Newport  News  (Richmond:  Dietz 
Press,  1986),  p.  69. 

Chapter  4 
Change  and  Continuity 

1.  Various  people  have  related  this  story  to  me,  notably  Edgar  M.  Cortright  in  a  July  1988 
interview,  Yorktown,  Va.,  transcript,  pp.  9-10,  OHC,  LHA,  and  Laurence  K.  Loftin, 
Jr.,  in  a  July  1990  interview,  Newport  News,  Va.    The  versions  told  by  Cortright  and 
Loftin  are  virtually  identical  in  the  details  of  the  incident;  however,  Cortright,  who  at 
the  time  of  the  incident  was  the  deputy  director  of  the  Office  of  Lunar  and  Planetary 
Programs  at  NASA  headquarters,  tells  the  story  with  an  implicit  criticism  of  Thompson's 
behavior.  Loftin,  Thompson's  assistant  director,  praises  the  boldness  and  independence 
of  his  center  director.  Cortright  witnessed  the  incident;  Loftin  heard  about  it. 

2.  Dr.  John  E.  Duberg  interview  with  author,  Hampton,  Va.,  8  Aug.  1991. 

3.  Senator  John  F.  Kennedy  quoted  in  Michael  Collins,  Liftoff:    The  Story  of  America's 
Adventure  in  Space  (New  York:  NASA/Grove  Press,  1988),  p.  63. 

4.  For  a  critique  of  Webb's  approach  to  the  management  of  NASA  and  its  application  to 
other  arenas  of  American  life,  see  Walter  A.  McDougall,  The  Heavens  and  the  Earth:  A 
Political  History  of  the  Space  Age  (New  York:  Basic  Books,  1985),  pp.  380-383. 

5.  Quoted  in  McDougall,    Heavens  and  the  Earth,   p.   383,   from  Webb's  remarks  to  a 
"Conference  on  Space,  Science,  and  Urban  Life,"  held  at  the  Dunsmuir  House,  Oakland, 
Calif.,  30  Mar.  1963. 

6.  My  presentation  of  the  general  character  of  Langley's  formal  organization  in  the  1960s 
is  based  on  my  study  of  organization  charts,  correspondence  pertaining  to  the  formal 
organization  found  in  file  E26-3C,  LCF,  and  a  staff  office  notebook  that  chronicles 
the  major  organizational  changes  at  Langley  from  World  War  II  to  the  present.    This 
notebook,  in  the  LHA,  has  been  kept  up-to-date  in  recent  years  by  Langley  historical 
program  coordinator  Richard  T.  Layman. 

7.  Langley  Announcement,  No.  7-62,  "Reorganization  of  Langley  Research  Center,"  16  Feb. 
1962,  E26-3C,  LCF. 

8.  Ira  H.  Abbott,   director  of  Advanced  Research  Programs,   to  Langley,   Attn.:    F.  L. 
Thompson,   "Establishment  of  an  Analysis  and  Computation  Division  at  the  Langley 
Research  Center,  and  authorization  to  appoint  a  division  chief  for  the  new  division," 
13  Dec.  1960,  E26-3C,  LCF. 

9.  For  the  complete  history  of  PARD,  see  Joseph  A.  Shortal's  exhaustive  A  New  Dimension: 
Wallops  Flight  Test  Range,  The  First  Fifteen  Years,  NASA  RP-1028  (Washington,  1978). 
For  a  shorter  treatment  with  special  emphasis  on  PARD's  role  as  a  training  ground  for 
space,  see  Robert  R.  Gilruth,  "Memoir:   From  Wallops  Island  to  Mercury,  1945-1958," 
unpublished    manuscript    presented    at    the    Sixth   International    History    Symposium, 
Vienna,  Austria,  13  Oct.  1972,  copy  in  LHA. 


462 


Notes  for  Chapter  4 

10.  Duberg  interview,  8  Aug.   1991.  For  some  interesting  history  related  to  the  "maverick" 
character  of  PARD,  see  Henry  S.  F.  Cooper's  essay,  "We  Don't  Have  to  Prove  Ourselves," 
in  The  New  Yorker,  2  Sept.  1991,  esp.  pp.    57-58.    Dr.  Cooper's  article  tells  the  story 
of  former  PARD  employees  Maxime  A.  Faget  and  Caldwell  Johnson;  after  working  on 
the  STG,  the  two  N  AC  A/NASA  engineers  went  on  to  operate  their  own  company,  Space 
Industries  International,  on  the  outskirts  of  Johnson  Space  Center. 

11.  Cortright  interview,  July  1988,  esp.  pp.  1-6.     On  NACA  Director  George  Lewis  and 
his  paternalistic  and  obfuscatory  style  of  management,  see  James  R.  Hansen,   Engi- 
neer in  Charge:  A  History  of  the  Langley  Aeronautical  Laboratory,  1917-1958,  NASA 
SP-4305  (Washington,  1987),  pp.  24-40,  and  "George  W.  Lewis  and  the  Management  of 
Aeronautical  Research,"  in  Aviation's  Golden  Age:  Portraits  from  the  1920s  and  1930s, 
ed.  William  M.  Leary  (Iowa  City:  University  of  Iowa  Press,  1989),  pp.  93-112. 

12.  Laurence  K.  Loftin,  Jr.,  interview  with  author,  Newport  News,  Va.,  26  June  1991. 

13.  Ibid. 

14.  Israel  Taback  quoted  in  James  Schultz,  Winds  of  Change:  Langley  Research  Center's  75 
Years  of  Accomplishment,  NP-130  (Washington,  1992),  p.  21. 

15.  On  the  dissolution  of  the  Hydrodynamics  Division,  see  H.  J.  E.  Reid,  internal  memo- 
randum for  all  concerned,  "Changes  in  Organization  of  the  Langley  Research  Center," 
23  Dec.  1959,  E26-3C,  LCF. 

16.  For  a  brief  history  of  Langley's  hydrodynamics  work,   see  the  NASA  Langley  press 
release,  "Changing  Research  Emphasis  Leads  to  Dissolution  of  NASA's  Hydrodynamics 
Division  after  29  Years,"    29  Dec.   1959,   E26-3C,   LCF.     For  design  information  on 
the  Martin  YP6M-1  Seamaster  jet-propelled  flying  boat,  see  Laurence  K.  Loftin,  Jr., 
Quest  for  Performance:    The  Evolution  of  Modern  Aircraft,  NASA  SP-468  (Washington, 
1985),  pp.  382-385.   For  basic  information  about  Convair's  Sea  Dart,  see  Ray  Wagner, 
American  Combat  Planes,  3d  enlarged  ed.  (Garden  City,  N.Y.:  Doubleday  &  Co.,  1982), 
pp.  515-517. 

17.  See   H.    J.    E.    Reid   to   NASA   headquarters,    "Proposed   Reorganization   of  Langley 
Research  Center  and  Changes  in  Excepted  Positions,"  16  Dec.  1959,  E26-3C,  LCF.    For 
Parkinson's  principal  work  on  the  hydrodynamics  of  seaplanes,  see  his  "Appreciation  and 
Determination  of  the  Hydrodynamics  Qualities  of  Seaplanes,"  NACA  TN-1290,  1947. 

18.  On  the  NACA  Tank  No.  1,  see  Starr  Truscott,  "The  NACA  Tank— A  High-Speed  Towing 
Basin  for  Testing  Models  of  Seaplane  Floats,"  NACA  TR-470,  1933.    For  an  excellent 
detailed  presentation  on  the  nature  and  contributions  of  Langley's  Hydrodynamics 
Research  Division  in  the  1930s  and  early  1940s,  see  "Air  and  Water,"   in  George  W. 
Gray,  Frontiers  of  Flight:   The  Story  of  NACA  Research  (New  York:   Alfred  A.  Knopf, 
1948),  pp.  63-81.  For  the  history  of  flying  boats,  see  Kenneth  Munson,  Flying  Boats  and 
Seaplanes  Since  1910  (New  York:  Macmillian  Co.,  1971);  and  Capt.  Richard  C.  Knott, 
The  American  Flying  Boat,  An  Illustrated  History  (Annapolis:    Naval  Institute  Press, 
1979).    On  the  floatplanes  of  World  War  II  specifically,  see  Wagner,  American  Combat 
Planes,  pp.  334-345. 

19.  On  the  U.S.  Navy's  termination  of  its  flying-boat  program  in  the  late  1950s,  see  Loftin, 
Quest  for  Performance,  p.  385,  and  Wagner,  American  Combat  Planes,  p.  481. 

20.  Reid,  "Changes  in  Organization  of  Langley  Research  Center,"  23  Dec.  1959. 

21.  Quoted  in  NASA  Langley  press  release,  "Changing  Research  Emphasis,"  p.  1. 

463 


Notes  for  Chapter  4 

22.  Laurence  K.   Loftin,   Jr.,   interview  with  author,   Newport  News,   Va.,   5  Aug.   1989, 
transcript,  p.  5,  OHC,  LHA. 

23.  Raymond  L.  Bisplinghoff,  "Twenty-Five  Years  of  NASA  Aeronautical  Research:  Reflec- 
tions and  Projections,"  in  Space  Applications  at  the  Crossroads:  21st  Goddard  Memorial 
Symposium,  eds.  John  H.  McElroy  and  E.  Larry  Heacock  (San  Diego:  Published  for  the 
American  Astronautical  Society  by  Univelt,  1983),  pp.  30-31. 

24.  John  V.  Becker,   "The  Development  of  Winged  Reentry  Vehicles,  1952-1963,"   unpub- 
lished, 23  May  1983,  p.  31,  copy  in  LHA. 

25.  Ibid. 

26.  Mark  R.  Nichols  interview  with  author,  Hampton,  Va.,  6  July  1988,  transcript,  pp.  2-3, 
OHC,  LHA. 

27.  All  of  the  statistical  information  in  this  chapter  regarding  personnel,  payroll,  budget, 
expenditures,  and  so  forth  is  from  the  NASA  Historical  Data  Book,  NASA  SP-4012 
(Washington,   1988),  vol.   1,  NASA  Resources  1958-1968,  by  Jane  Van  Nimmen  and 
Leonard  Bruno  with  help  from  Robert  L.  Rosholt.   NASA  personnel  statistics  given  in 
various  sources  may  seem  contradictory.  The  NASA  Historical  Data  Book  gives  certain 
personnel  numbers  for  a  given  year,  and  the  NASA  Historical  Pocket  Statistics  may  give 
others.    Neither  is  in  error;  the  differences  depend  on  when  the  count  was  taken.    The 
number  of  NASA  employees  changed  every  day.  The  NASA  Historical  Data  Book,  vol.  1, 
NASA  Resources  1958-1968,  which  is  my  source,  provides  personnel  statistics  for  30  June 
and  31  Dec.  of  each  year. 

28.  Earl  D.  Hilburn,   deputy  associate  administrator,   NASA  headquarters,   to  Dr.  Floyd 
L.  Thompson,   16  Aug.   1963,  copy  in  C131-1,  LCF.     Floyd  L.  Thompson  to  NASA 
headquarters,  "Manpower  Requirements,"  9  Sept.  1963,  C131-1,  LCF. 

29.  On  NASA's  early  problems  involving  the  low  ceiling  on  its  personnel  complement,  see 
Robert  L.  Rosholt,  An  Administrative  History  of  NASA,   1958-1963,  NASA  SP-4101 
(Washington,  1966),  pp.  139-141. 

30.  See  the  summary  of  the  NASA  Staff  Conference,  Monterey,  Calif.,  3-5  Mar.  1960,  p.  56, 
copy  in  NASA  HQA.  Glennan's  views  are  outlined  in  Rosholt,  Administrative  History  of 
NASA,  p.  140. 

31.  Rosholt  compared  Glennan's  1962  "Budget  Guidelines"  with  the  "Summary  Financial 
Plan  for  Fiscal  Year  1961,  NASA,"  that  NASA  submitted  to  the  Bureau  of  the  Budget 
on  19  Aug.  1960,  Administrative  History  of  NASA,  p.  140. 

32.  Roger  E.  Bilstein,  Orders  of  Magnitude:  A  History  of  the  NACA  and  NASA,  1915-1990, 
NASA  SP-4406  (Washington,  1989),  p.  71. 

33.  See  Hilburn  to  Thompson,  16  Aug.  1963. 

34.  Thompson  to  NASA  headquarters,  "Manpower  Requirements,"  9  Sept.  1963. 

35.  Axel  T.  Mattson,  research  assistant  for  Manned  Spacecraft  Projects,  Langley,  to  Manned 
Spacecraft  Center,  Attn.:  Mr.  M.  A.  Faget,  "Langley  Research  Center  Tests  of  Interest 
to  Manned  Spacecraft  Center,"  27  Aug.  1963.  Mattson  compiled  these  reports  regularly 
from  late  1962  into  1967.  Copies  of  them  are  preserved  in  the  Project  Apollo  files,  LCF. 
Their  contents  will  be  discussed  in  chap.  11  in  relation  to  Langley 's  work  in  support  of 
Apollo. 

36.  See  Table  4-19,    "Administrative   Operations   Direct   Obligations,    by  Installation   (in 
millions  of  dollars),"  in  NASA  Historical  Data  Book,  1:146. 

464 


Notes  for  Chapter  4 

37.  For  a  general  analysis  of  how  procurement  and  contracting  were  done  by  NASA  in  its  early 
years,  see  Rosholt,  Administrative  History  of  NASA,  pp.  61-65,  and  Arnold  S.  Levine, 
Managing  NASA  in  the  Apollo  Era,  NASA  SP-4102  (Washington,  1982),  pp.  65-105.  On 
the  allocation  of  resources  by  NASA  headquarters,  see  Levine,  pp.  63-64. 

38.  See  Table  5-19,    "Value  of  Awards  by  Installation   (in  millions  of  dollars),"    NASA 
Historical  Data  Book,  1:202. 

39.  See  Table  4-26,  "Amounts  Programmed  for  Research  and  Development,  by  Installation 
(in  millions  of  dollars),"  NASA  Historical  Data  Book,  1:166. 

40.  See  tables  5-20  through  5-27  of  the  NASA  Historical  Data  Book,  1:203-226,  for  detailed 
information  on  NASA  contracts  and  contractors  in  the  period  1958  to  1968. 

41.  Steven  T.  and  Sarah  W.  Corneliussen,  "NASA  Langley  Research  Center  Support  Services 
Contracting,  Historical  Summary,"  29  May  1991,  p.  1.  Much  of  the  discussion  about  the 
history  of  support  services  contracting  at  Langley  that  follows  in  this  chapter  is  based 
on  the  unpublished  Corneliussen  study  or  on  the  oral  histories  they  conducted  for  it.  A 
copy  of  this  study  and  all  the  materials  collected  for  it  are  preserved  in  the  LHA. 

42.  Duberg  interview,  8  Aug.  1991. 

43.  Corneliussen  and  Corneliussen,  "Support  Services  Contracting,"  p.  6.  For  an  analysis  of 
support  services  and  the  issue  of  the  integrity  of  the  NASA  civil  service  staff,  see  Levine, 
Managing  NASA  in  the  Apollo  Era,  pp.  86-87. 

44.  Corneliussen  and  Corneliussen,  "Support  Services  Contracting,"  pp.  9-10. 

45.  For  definitions  of  the  various  types  of  contracts  authorized  by  NASA,  see  app.  D  of  Levine, 
Managing  NASA  in  the  Apollo  Era,  pp.  313-314.  Levine's  definitions  are  based  on  House 
Committee  on  Science  and  Astronautics,  1965  NASA  Authorization,  88th  Cong.,  2d  sess. 
(1964),  pp.  1327-1328. 

46.  Levine,  Managing  NASA  in  the  Apollo  Era,  app.  D,  p.  313. 

47.  Corneliussen  and  Corneliussen,  "Support  Services  Contracting,"  p.  17. 

48.  Edwin  Kilgore  interview  with  Steven  T.  Corneliussen,  Hampton,  Va.,  Apr.  1991,  LHA. 

49.  Sherwood  Butler  interview  with  Sarah  W.  Corneliussen,  Hampton,  Va.,  Apr.  1991,  LHA. 

50.  Robert  Moore  interview  with  Sarah  W.  Corneliussen,  Hampton,  Va.,  Apr.  1991,  LHA. 

51.  On  the  American  atomic  bomb  project,  see  H.  D.  Smyth,  Atomic  Energy  (Washington: 
Government  Printing  Office,  1945);  this  is  the  original  official  report.  For  a  very  readable 
historical  analysis,  see  Stephane  Groueff,  Manhattan  Project:   The  Untold  Story  of  the 
Making  of  the  Bomb  (Boston:    Little,  Brown  &  Co.,  1967),  and  Richard  Rhodes,   The 
Making  of  the  Atomic  Bomb  (New  York:    Simon  &  Schuster,   1986).     See  also  R.  G. 
Hewlett  and  O.  E.  Anderson,  A  History  of  the  United  States  Atomic  Energy  Commission 
(Pennsylvania  State  University  Press,  1962),  vol.  1,  The  New  World. 

On  Project  Sherwood  and  the  history  of  the  program  in  controlled  thermonuclear 
reactions  carried  out  by  the  U.S.  Atomic  Energy  Commission  during  the  period  1951  to 
1958,  see  Amasa  Bishop,  Project  Sherwood:  The  U.S.  Program  in  Controlled  Fusion 
(Reading,  Mass.:  Addison- Wesley  Publishing  Co.,  1958),  and  Joan  Lisa  Bromberg, 
Fusion:  Science,  Politics,  and  the  Invention  of  a  New  Energy  Source  (Cambridge,  Mass.: 
MIT  Press,  1983.) 

52.  NASA  Management  Manual,  pt.  1,   "General  Management  Instructions,"  chap.  4,  No. 
4-1-1,  8  Mar.  1963,  p.  1. 


465 


Notes  for  Chapter  5 

53.  On  the  administrative  evolution  of  Apollo  from  a  project  to  a  program,  see  Levine, 
Managing  NASA  in  the  Apollo  Era. 

54.  National  Research  Council,  NASA  's  Aeronautics  Research  and  Technology  Base:  A  Study 
of  the  Report  of  the  Ad  Hoc  Aeronautics  Research  Committee,  Aeronautics  and  Space 
Engineering  Board  (Washington:  National  Academy  of  Sciences,  1979),  p.  6. 

55.  On  the  NACA/NASA  work  on  the  XB-70  and  XB-70A,  see  Richard  P.  Hallion,  On  the 
Frontier:   Flight  Research  at  Dryden,  1946-1981,  NASA  SP-4303  (Washington,  1984), 
pp.  180-189. 

56.  John  V.  Becker,  High-Speed  Frontier:    Case  Histories  of  Four  NACA  Programs,  1920- 
1950,  NASA  SP-445  (Washington,  1980),  pp.   86-87.   For  more  on  the  feelings  of  other 
Langley  divisions  about  PARD,  see  Hansen,  Engineer  in  Charge,  pp.  269  270. 

57.  Macon  C.  Ellis,  Jr.,  interview  with  author,  Hampton,  Va.,  5  Nov.  1991. 

58.  Loftin  telephone  interview  with  author,  30  Jan.  1992. 

59.  Charles  J.  Donlan  interview  with  author,   Hampton,   Va.,    17  Aug.   1990,   transcript, 
pp.  27-28,  OHC,  LHA. 

60.  Rosholt,  Administrative  History  of  NASA,  pp.  145-146. 

61.  Ibid. 

62.  Ibid.,  pp.  157-158. 

63.  Levine,  Managing  NASA  in  the  Apollo  Era,  pp.  43-45. 

64.  What  has  been  described  in  the  text  is  a  system  of  project  management  that  evolved 
gradually,  in  fits  and  starts,  through  several  major  reorganizations  taking  place  during 
the  first  five  years  of  NASA,  from  October  1958  to  October  1963.     For  a  far  more 
comprehensive   examination  of  the  development   of  the   NASA   organization   and   its 
management  in  this  period,  the  reader  should  consult  the  previously  cited  works  by 
Rosholt  and  Levine.  Both  books  go  into  great  detail  about  NASA's  project  management. 

65.  On  the  nature  of  the  wartime  work  at  NACA  Langley,  see  Hansen.  Engineer  in  Charge, 
pp.  194-202. 

Chapter  5 
The  "Mad  Scientists"  of  MPD 

1.  The  scientific  and  technical  literature  generated  by  the  field  of  plasma  physics  since  its 
emergence  as  an  identifiable  scientific  discipline  in  the  1950s  is  too  vast  to  cite.  For  basic 
technical  information  about  the  field  provided  in  this  chapter,  I  have  primarily  used 
a  well-known  textbook  from  the  early  1960s,  Ali  Bulent  Cambel,  Plasma  Physics  and 
Magneto  fluid- Mechanics  (New  York:  McGraw-Hill,  1963)  and  one  NASA  publication 
by  Langley  authors  Adolf  Busemann,  Robert  V.  Hess,  Paul  W.  Huber,  Clifford  H. 
Nelson,  and  Macon  C.  Ellis,  Jr.,  Plasma  Physics  and  Magnetohydrodynamics  in  Space 
Exploration,  NASA  SP-25  (Washington,  1962).  These  papers  were  originally  presented  by 
their  authors  at  a  NASA/University  Conference  on  the  Science  and  Technology  of  Space 
Exploration,  Chicago,  111.,  1-3  Nov.  1962.  Ironically,  in  the  midst  of  using  Cambel's 
book,  I  discovered  that  the  MPD  Branch  at  Langley  did  not  care  for  Cambel,  a  physics 
professor  at  Northwestern  University  in  Evanston,  111.  In  the  1960s,  the  head  of  the 
MPD  Branch,  Macon  C.  Ellis,  Jr.,  turned  down  more  than  one  of  Cambel's  ideas  for 
doing  MPD  research  under  contract  to  Langley. 

466 


Notes  for  Chapter  5 

For  a  more  basic  discussion  of  the  nature  of  gases  and  plasma  than  can  be  found  in 
the  Cambel  textbook,  consult,  as  I  did,  Paul  G.  Hewitt,  Conceptual  Physics,  6th  ed. 
(Glenview,  111.:  Scott,  Foresman  and  Co.,  1989),  pp.  244-246;  Gary  L.  Buckwalter  and 
David  M.  Riban,  College  Physics  (New  York:  McGraw-Hill,  1987),  pp.  832-835;  or  some 
other  textbook  in  physics. 

2.  See  Thomas  G.  Cowling,  Magnetohydrodynamics  (London:  Adam  Hilger,  1976),  pp.  1-2. 
However,  the  best  discussion  of  the  confounding  nomenclature  of  plasma  physics  can  be 
found  in  the  introduction  to  an  unpublished  paper  by  Langley  MPD  researcher  George  P. 
Wood.  A  copy  of  this  paper,  which  is  undated  (but  seems  to  have  been  written  in  1959), 
can  be  found  in  the  collection  of  personal  papers  donated  in  1991  to  the  LHA  by  Macon  C. 
Ellis,  Jr. 

3.  For  a  historical  analysis  of  Chapman's  work  on  solar  physics  and  solar  and  terrestrial  rela- 
tionships, see  Karl  Hufbauer,  Exploring  the  Sun:  Solar  Science  Since  Galileo  (Baltimore 
and  London:  Johns  Hopkins  University  Press,  1991),  esp.  pp.  154-158  and  217-221.  Sev- 
eral references  are  made  to  the  contributions  of  the  Chapman  group  in  Homer  Newell, 
Beyond  the  Atmosphere:   Early  Years  of  Space  Science,  NASA  SP-4211  (Washington, 
1980). 

4.  For  those  interested  in  the  many  complex  details  of  Alfven's  work,  see  Alfven,  Cosmical 
Electrodynamics  (Oxford  University  Press,  1950).  Also  see  Hufbauer,  Exploring  the  Sun, 
pp.  205-206,  216,  242,  253,  and  311. 

5.  For  details  about  the  discovery  of  the  earth's  radiation  belts  from  the  scientist  who  dis- 
covered them,  see  James  Van  Allen,  The  Origins  of  Magnetospheric  Physics  (Washington: 
Smithsonian  Institution  Press,  1983).   For  a  brief  historical  analysis  of  the  discovery  of 
the  Van  Allen  radiation  belts,  see  Newell,  Beyond  the  Atmosphere,  esp.  pp.  173-176; 
Hufbauer,  Exploring  the  Sun,  pp.  165-166;  Roger  Bilstein,  Orders  of  Magnitude:  A  His- 
tory of  the  NACA  and  NASA,  1915-1990,  NASA  SP-4406  (Washington,  1989),  pp.  46-47 
and  81-82;  and  John  E.  Naugle,  First  Among  Equals:    The  Selection  of  NASA  Space 
Science  Experiments,  NASA  SP-4215  (Washington,  1991),  pp.  15-17. 

6.  On  the  emergence  of  the  concept  of  the  magnetosphere,  see  Van  Allen,  The  Origins  of 
Magnetospheric  Physics;  Newell,  Beyond  the  Atmosphere,  pp.  172-186;  and  Hufbauer, 
Exploring  the  Sun,  pp.  228,  235,  and  240-243. 

7.  Newell,  Beyond  the  Atmosphere,  pp.  182-184.    On  Eugene  Parker's  theory  of  the  solar 
wind  and  its  impact  on  astrophysics  and  solar  science  from  1957  to  1970,  see  chap.  6  in 
Hufbauer,  Exploring  the  Sun,  pp.  213-258. 

8.  In   response   to   questions   that   the   author   had   posed   to   him   in   a   letter,    retired 
NAG  A/NASA  researcher  Macon  C.  Ellis,  Jr.,  in  early  Nov.  1991  tape-recorded  answers 
and  provided  other  oral  reminiscences  concerning  the  history  of  MPD  at  Langley.   The 
information  contained  on  the  roughly  90-minute  tape  provides  an  excellent  overview  of 
MPD  research  at  Langley.  However,  because  of  the  candid  nature  of  some  of  the  remarks 
made  about  fellow  MPD  personnel,  Ellis  chose  to  restrict  access  to  the  tape.    This  he 
did  by  giving  the  tape  to  the  author  for  his  personal  use  and  asking  that  a  copy  not  be 
placed  in  the  LHA.  The  quotation  from  Ellis  in  the  text  is  from  the  author's  handwritten 
transcript  of  this  audiotape,  p.  2. 

9.  Quoted  from  Ellis's  handwritten  notes  entitled  "Branch  Re-Cap,"  for  a  meeting  of  the 
Compressibility  Research  Division  on  18  June  1958,  p.  1.  The  branch  referred  to  was  the 

467 


Notes  for  Chapter  5 

Gas  Dynamics  Branch.  These  notes  are  in  a  folder  labeled  "Editorial  Copy,  1958,"  Ellis 
Collection,  LHA. 

10.  A  significant  amount  of  MPD-related  research  was  conducted  at  the  other  NASA  research 
centers.   For  example,  a  parallel  MPD  branch  at  Lewis  was  especially  interested  in  the 
promise  of  ion  thrusters  and  electric  propulsion.    The  best  sources  for  NASA's  early 
MPD  activities  agencywide  are  the  published  reports  of  the  intercenter  conferences  on 
plasma  physics;  see,  for  example,  Program  of  the  Third  NASA  Intercenter  Conference 
on  Magnetoplasmadynamics,   held  at  Langley,   24-25  Apr.   1962,   or   "Program  of  the 
Fourth  NASA  Intercenter  Conference  on  Plasma  Physics"  held  at  NASA  headquarters, 
2-4  Dec.   1964.     Copies  of  most  of  these  NASA  meeting  programs  are  in  the  Ellis 
Collection,  LHA.  For  a  list  of  organizations  that  conducted  MPD  and  plasma  physics 
research,  as  well  as  for  lists  of  the  dozens  of  people  involved,  see  the  relevant  documents 
in  the  folder  labeled  "Statistics  on  Visitors,  etc.,"  Ellis  Collection,  LHA. 

11.  NASA    Langley    press    release,     "NASA    1959    Inspection:    Magnetoplasmadynamics," 
Oct.  1959,  p.  1,  "1959-1964  Inspection"  drawer,  LHA. 

12.  Ellis  audiotape,  author's  transcript,  p.  9. 

13.  Ibid.,  p.  5. 

14.  "Agenda  and  Tentative  Schedule  for  Interlaboratory  Magnetohydrodynamics  Meeting, 
Sept.  22-23,  1958,"  copy  in  folder  labeled  "MHD  Meeting     9  22/23-58,"  Ellis  Collection, 
LHA.  The  folder  contains  several  documents  related  to  the  conference.    Dr.  Busemann 
served  as  the  "meeting  chairman,"  and  Macon  C.  Ellis,  Jr.,  served  as  its  "coordinator." 
Busemann  continued  to  work  on  MHD-related  problems  throughout  the  early  1960s;  see, 
for  example,  "On  the  Karman  Vortex  Street  in  Magnetofluid-Dynamics,"   80th  Annual 
Proceedings  of  the  Institute  of  Aerospace  Sciences,  1962;  "Relations  between  Aerodynam- 
ics and  Magnetohydrodynamics,"  which  was  presented  at  a  NASA-sponsored  conference 
at  the  University  of  Chicago,  1-3  Nov.  1962;  and  "Lift  Control  in  Magnetohydrodynam- 
ics," a  paper  presented  at  the  International  Symposium  of  Applications  of  the  Theory 
of  Functions  in  Continuum  Mechanics  at  Tbilisi  in  the  Soviet  Union,  18-24  Sept.  1963. 
Copies  of  all  three  papers  are  available  in  the  Langley  Technical  Library. 

15.  Macon  C.   Ellis,   Jr.,   to  Charles  J.  Donlan,   Langley  associate  director,    "Outline  of 
NASA-Postdoctoral  Resident  Research  Associate  Program  Administered  by  National  Re- 
search Council — NASA — NAE  and  recommendations  relative  to  Langley  participation," 
29  Nov.  1966,  copy  in  folder  labeled  "RRA  Program,"  Ellis  Collection,  LHA.  This  folder 
contains  a  rather  complete  file  on  MPD's  involvement  in  the  RRA  program  at  Langley. 

16.  Special  personnel  folders  for  Adamson  and  Feix  are  in  the  Ellis  Collection,  LHA. 

17.  For  insight  into  the  amount  and  type  of  MPD  research  conducted  by  outside  groups  under 
contract  to  Langley,  see  "MPD  Branch  Research  Grants  and  Contracts,"  a  seven-page 
document,  dated  Oct.  1966,  in  the  Ellis  Collection,  LHA.    MPD  monitored  37  contracts 
amounting  to  almost  $4  million. 

The  MPD  Branch  collected  complete  information  on  its  committee  memberships, 
publications,  contracts,  and  visitors  in  preparation  for  a  major  June  1968  briefing  of  new 
Langley  Director  Edgar  M.  Cortright.  Much  of  this  information  is  in  a  folder  labeled 
"Statistics  on  Visitors,  etc.,"  Ellis  Collection,  LHA. 

18.  NASA  established  a  Technical  Working  Group  for  Electric  Propulsion  (TWGEP)  on 
1  Sept.  1960.  Chaired  by  Dr.  Ernst  Stuhlinger  of  NASA  Marshall,  the  group  comprised 

468 


Notes  for  Chapter  5 

representatives  from  Marshall,  JPL,  Lewis,  Goddard,  Langley,  and  NASA's  Nuclear 
Engine  Project  Office.  MPD  Branch  Head  Mike  Ellis  was  Langley's  representative.  The 
Ellis  Collection  contains  a  complete  set  of  minutes  to  the  TWGEP  meetings.  The  first 
of  these  meetings  convened  at  Marshall  on  13  Oct.  1960. 

19.  NASA    Langley    press    release,     "NASA    1959    Inspection:    Magnetoplasmadynamics," 
Oct.  1959,  p.  1. 

20.  John  V.  Becker  interview  with  author,  Hampton,  Va.,  5  Aug.  1991. 

21.  Donald  D.   Baals  and  William  R.   Corliss,    Wind   Tunnels  of  NASA,  NASA   SP-440 
(Washington,  1981),  p.  85. 

22.  For  information  on  the  arc-jet  facilities  built  and  operated  at  Langley  in  the  1960s, 
see  Martin  A.  Weiner,  Resume  of  Research  Facilities  at  the  Langley  Research  Center 
(Washington:  NASA,  July  1968).  A  copy  of  this  catalogue  (No.  CN-123,837)  is  available 
in  the  Langley  Technical  Library. 

23.  For  more  information  on  the  hotshot  tunnel,  see  Baals  and  Corliss,   Wind  Tunnels  of 
NASA,  pp.  84-85;  and  Weiner,  Resume  of  Research  Facilities,  facilities  in  Building  1247B. 

24.  William  H.  Allen,  ed.,  Dictionary  of  Technical  Terms  for  Aerospace  Use,  NASA  SP-7 
(Washington,  1965),  p.  251.  See  also  Baals  and  Corliss,  Wind  Tunnels  of  NASA,  p.  84. 

25.  A  one-page  chronology  of  early  shock-tube  operation  in  Langley's  Gas  Dynamics  Labo- 
ratory from  1951  to  1953  is  located  in  the  Ellis  Collection,  LHA;  see  the  folder  labeled 
"Editorial  Copy,  1958."    Also  in  this  folder  is  a  chart  (circa  1957)  listing  the  major  re- 
search facilities  in  the  Gas  Dynamics  Laboratory  and  giving  basic  data  on  their  test 
section  sizes,  running  times,  dates  of  initial  operation,  and  stagnation  temperatures  and 
pressures.  Mike  Ellis  used  this  chart  and  its  associated  information  to  brief  a  meeting  of 
the  NACA's  aerodynamics  committee  held  at  Langley  on  1  May  1958. 

26.  Becker  interview,  5  Aug.  1991.  See  George  P.  Wood  and  Arlen  F.  Carter,  "Considerations 
in  the  Design  of  a  Steady  DC  Plasma  Accelerator,"  presented  at  the  Third  Biennial  Gas 
Dynamics  Symposium,  Northwestern  University,  Evanston,  111.,  24-26  Aug.  1959,  copy 
in  the  Langley  Technical  Library. 

27.  George  P.  Wood,   Arlen  F.   Carter,   Alexander  P.  Sabol,   and  Richard  H.  Weinstein, 
"Experiments  in  Steady  State  Crossed-Field  Acceleration  of  Plasma,"  in  Physics  of  Fluids 
5  (May  1961):   652;  George  P.  Wood,  Arlen  F.  Carter,  Hubert  K.  Lintz,  and  J.  Byron 
Pennington,  "A  Theoretical  Treatment  of  the  Steady-Flow,  Linear,  Crossed-Field,  Direct- 
Current  Plasma  Accelerator  for  Inviscid,  Adiabatic,  Isothermal,  Constant-Area  Flow," 
NASA  TN  D-924,  June  1961;  Arlen  F.  Carter,  George  P.  Wood,  D.  R.  McFarland,  and 
W.  R.  Weaver,   "Research  on  a  Linear  Direct-Current  Plasma  Accelerator,"   in  AIAA 
Journal  3  (June  1965):    1040-1045.     For  an  earlier  version  of  this  paper,  see  George 
P.  Wood,  Arlen  F.  Carter,  Alexander  P.  Sabol,  D.  R.  McFarland,  and  W.  R.  Weaver, 
"Research  on  Linear  Crossed-Field  Steady-Flow  DC  Plasma  Accelerators  at  Langley 
Research  Center,  NASA,"  presented  at  the  AGARD  Specialists'  Meeting  on  Arc  Heaters 
and  MHD  Accelerators  for  Aerodynamic  Purposes,  Rhode-Saint-Genese,  Belgium,  21-23 
Sept.  1964,  published  as  AGARDograph  84,  copy  in  the  Langley  Technical  Library.  For 
a  general  description  of  Langley's  early  work  on  plasma  accelerators,  see  NASA  Langley 
press  release,  "NASA  1959  Inspection:  Magnetoplasmadynamics,"  pp.  2-3. 

28.  Becker  interview,  5  Aug.  1991. 


469 


Notes  for  Chapter  5 

29.  Ibid.  A  preliminary  proposal  for  what  Stack  wanted  to  call  a  "Trans- Satellite- Velocity" 
wind  tunnel  can  be  found  in  the  folder  labeled  "Editorial  Copy,  1958,"  Ellis  Collection, 
LHA. 

30.  Arlen    F.    Carter,    Willard    R.    Weaver,    Donald    R.    McFarland,    Stephen    K.    Park, 
and   George   P.   Wood,    "Development    and   Initial   Operating   Characteristics   of  the 
20-Megawatt  Linear  Plasma  Accelerator  Facility,"   NASA  TN  D-6547,  Dec.  1971;  see 
also  "Design  of  the  20-Megawatt  Linear  Plasma  Accelerator  Facility,"  NASA  TN  D-6116, 
Jan.  1971,  by  the  same  authors. 

31.  On  Langley's  work  with  Hall-current  accelerators,  see  John  R.  Sevier,  Robert  V.  Hess, 
and  Philip  Brockman,   "Coaxial  Hall-Current  Accelerator  Operation  at  Forces  and  Ef- 
ficiencies Comparable  to  Conventional  Crossed-Field  Accelerators,"    in  ARS  Journal 
32   (Jan.   1962):     78-80;    Philip  Brockman,   Robert  V.  Hess,   and  Richard  Weinstein, 
"Measurements  and  Theoretical  Interpretation  of  Hall  Currents  for  Steady  Axial  Dis- 
charges in  Radial  Magnetic  Fields,"    presented  at  the  Fifth  Biennial  Gas  Dynamics 
Symposium,  Northwestern  University,  Evanston,  111.,  14-16  Aug.  1963;  W.  Grossmann, 
H.  A.  Hassan,  and  R.  V.  Hess,  "Experiments  with  Co-Axial  Hall-Current  Plasma  Ac- 
celerator,"   presented  at  AIAA  Fourth  Electric  Propulsion  Conference,   Philadelphia, 
31  Aug.-2  Sept.  1964;  H.  A.  Hassan,  R.  V.  Hess,  and  W.  Grossmann,  "Study  of  Coax- 
ial Hall  Current  Accelerators  at  Moderate  Pressures,"   NASA  TN  D-3286,  Oct.  1966. 
Copies  of  all  these  papers  are  available  in  the  Langley  Technical  Library.    The  author 
wishes  to  thank  Mr.  Hess  for  his  generous  assistance  in  explaining,  via  letters,  telephone 
conversations,  and  personal  interviews,  the  development  of  the  MPD  accelerators  at 
Langley. 

32.  Robert  V.  Hess  and  Macon  C.  Ellis,  Jr.,  to  Floyd  L.  Thompson,  associate  director, 
"Visit  to  the  Research  Institute  of  Temple  University,  Philadelphia,  Pa.,  to  discuss  high 
temperature  flames,"  21  June  1959,  copy  in  folder  labeled  "CN  (Early),"  Ellis  Collection, 
LHA.  Three  folders  containing  material  on  Langley's  development  of  a  cyanogen  flame 
apparatus  are  in  the  Ellis  Collection,  LHA.    Ellis  also  comments  on  the  cyanogen  flame 
apparatus  in  his  audiotape,  author's  transcript,  p.  4. 

33.  NASA    Langley    press    release,     "NASA    1959    Inspection:    Magnetoplasmadynamics," 
pp.  1-2.  For  the  early  experimental  findings  involving  the  Langley  cyanogen  flame  appa- 
ratus, see  Paul  W.  Huber  and  Paul  B.  Gooderum,  "Experiments  with  Plasmas  Produced 
by  Potassium-Seeded  Cyanogen  Oxygen  Flames  for  Study  of  Radio  Transmission  at  Sim- 
ulated Reentry  Vehicle  Plasma  Conditions,"  NASA  TN  D-627,  Oct.  1961.  See  also  Huber, 
"Experiments  with  Plasmas  Produced  by  Alkali-Metal-Seeded  Cyanogen-Oxygen  Flames 
for  Study  of  Electromagnetic  Wave  Propagation  at  the  Langley  Research  Center,"  paper 
presented  at  the  Symposium  on  the  Plasma  Sheath,  Cambridge,  Mass.,  7-9  Dec.  1959, 
copy  in  the  Langley  Technical  Library. 

Calvin  T.  Swift's  major  MPD  research  papers  from  the  period  include  "Generalized 
Treatment  of  Plane  Electromagnetic  Waves  Passing  through  an  Isotropic  Inhomogeneous 
Plasma  Slab  at  Arbitrary  Angles  of  Incidence,"  NASA  TR  R-172,  Dec.  1963,  coauthored 
by  J.  S.  Evans;  "Radiation  from  Slotted-Cylinder  Antennas  in  a  Reentry  Plasma  Envi- 
ronment," NASA  TN  D-2187,  Feb.  1964;  "The  Aperture  Admittance  of  a  Ground  Plane- 
Mounted  Waveguide  Illuminating  a  Perfectly  Conducting  Sheet,"  NASA  TN  D-4366, 
Mar.  1968,  coauthored  by  J.  Earl  Jones;  "Experimental  Investigation  of  a  Plasma 

470 


Notes  for  Chapter  5 

Covered,  Axially  Slotted  Cylinder,"  in  IEEE  Transactions  on  Antennas  and  Propaga- 
tion 17  (Sept.  1969):  598  605,  coauthored  by  P.  B.  Gooderum  and  S.  L.  Castellow,  Jr.; 
and  "A  Theoretical  Investigation  of  Electromagnetic  Waves  Obliquely  Incident  Upon  a 
Plasma  Slab,"  NASA  TR  R-339,  Apr.  1970. 

34.  Very  little  can  be  found  in  any  publications  of  the  NASA  History  Series  about  Project 
RAM.  The  NASA  Historical  Data  Book,  vol.  2,   Programs  and  Projects,    1958-1968, 
pp.   464-466,   provides  only  the  most  basic  information.      For  technical  information 
on  the  radio  transmission  problem  caused  by  the  plasma  sheath  and  early  thoughts 
about  the  potential  for  attenuating  it,  see  Macon  C.  Ellis,  Jr.,  and  Paul  W.  Huber, 
"Real  Gas  Flow  Conditions  about  Hypersonic  Vehicles,"  in  Reentry  Dynamics,  Part  II: 
Proceedings  of  the  Conference  on  Physics  of  the  Solar  System  and  Reentry  Dynamics, 
July  31  to  Aug.   11,   1961  (Blacksburg,  Va.:    Virginia  Polytechnic  Institute  and  State 
University,  1962),  pp.  120-154;  and  John  S.  Evans  and  Paul  W.  Huber,   "Calculated 
Radio  Attenuation  due  to  Plasma  Sheath  on  Hypersonic  Blunt-Nosed  Cone,"  NASA  TN 
D-2043,  Dec.  1963. 

35.  Handwritten  notes  for  briefing  of  the  Langley  senior  staff,  15  Feb.  1962,  in  folder  labeled 
"Text  and  Vu-graphs  for  2/15/62  Briefing,"  Ellis  Collection,  LHA. 

36.  Quoted  from   "Langley  Feasibility  Study  of  Release  of  Barium  Cloud  within  Earth's 
Magnetosphere,"   16  Aug.  1965,  p.  1,  copy  in  folder  labeled  "Ba-Cloud  Project,"  Ellis 
Collection,  LHA. 

37.  For  a  summary  of  the  scientific  objectives  of  the  barium  cloud  experiment,  see  "Scientific 
Merits  of  Experiment  Involving  a  Barium  Release  at  High  Altitude   (Several  Earth 
Radii),"  an  unpublished  Langley  paper  from  early  1968,  Ba-Cloud  file,  Ellis  Collection, 
LHA. 

38.  "Langley  Feasibility  Study  of  Release  of  Barium  Cloud,"  p.  2. 

39.  L.  Biermann,  R.  Lust,  Rh.  Lust,  and  H.  U.  Schmidt,  "Zur  Untersuchung  des  interplan- 
etaren  Mediums  mit  hilfe  kunstliche  eingebrachtee  lonenwolken"  ["Study  of  the  Interplan- 
etary Media  by  Means  of  Artificially  Ionized  Clouds"],  in  Zeitschrift  fur  Astrophysik  53 
(1961):  226  236.  On  Biermann's  important  contributions  to  solar  physics,  see  Hufbauer, 
Exploring  the  Sun,  pp.  215-221. 

40.  R.  V.  Hess,  head,   Plasma  Physics  Section,  to  Charles  J.  Donlan,  Langley  associate 
director,  "Request  for  transmittal  of  letter  to  the  Secretary  of  the  Space  Sciences  Steering 
Committee  concerning  recent  discussion  of  R.  V.  Hess  with  Prof.  Biermann  on  use  of 
plasma  cloud  as  space  probe,"   15  July  1964,  Ba-Cloud  file,  Ellis  Collection,  LHA.  On 
the  front  of  this  memorandum,  MPD  Branch  Head  Mike  Ellis  has  penned  the  words, 
"Hess'fs]  claim  for  plasma  cloud  idea!" 

41.  NASA,     "MPI/NASA    Magnetospheric    Ion    Cloud    Experiment    Project    Proposal/ 
Development  Plan,"  15  Dec.  1967,  copy  in  Ba-Cloud  file,  Ellis  Collection,  LHA.    Ellis 
audiotape,  author's  transcript,  pp.  3  and  19. 

42.  For  draft  copies  of  the  relevant  memoranda  of  understanding  between  NASA  and  West 
Germany's  Federal  Ministry  for  Scientific  Research,  see  Ba-Cloud  file,  Ellis  Collection, 
LHA. 

43.  NASA,  "MPI/NASA  Magnetospheric  Ion  Cloud." 

44.  On  the  accident  and  the  findings  of  the  NASA  investigative  board,  see  the  story  in  the 
New  York  Times,  9  July  1967.    Max  Planck  Institut  researchers  also  experienced  an 

471 


Notes  for  Chapter  5 

accident  involving  the  mishandling  of  barium;  for  an  account  see  Linda  Shiner,  "Come 
to  Aruba  When  the  Barium  Blooms,"  Air  &  Space  6  (Feb./Mar.  1992):  58. 

45.  Macon  C.  Ellis,  Jr.,  letter  to  author,  28  Jan.  1991,  p.  3.   Ellis  communicated  Langley's 
"scooping"  of  the  West  Germans  after  talking  to  Leo  D.  Staton,  one  of  the  Langley 
men  involved  in  NASA's  analysis  and  reporting  of  the  findings  from  the  barium  cloud 
experiments.    For  a  summary  of  the  results  of  the  barium  cloud  experiments,  see  Hal 
T.  Baber,  Jr.,  Kenneth  H.  Crumbly,  and  David  Adamson,  Barium  Releases  at  Altitudes 
Between  200  and  1000  Kilometers — A  Joint  Max  Planck  Institut — NASA  Experiment, 
NASA  SP-264  (Washington,  1971).    For  further  analysis  from  NASA  researchers,  see 
David  Adamson,  Clifford  L.  Fricke,  Sheila  Ann  T.  Long,  W.  F.  Landon,  and  D.  L.  Ridge, 
"Preliminary  Analysis  of  NASA  Optical  Data  Obtained  in  Barium  Cloud  Experiment  of 
Sept.  21,  1971,"  in  Journal  of  Geophysical  Research  78  (1  Sept.  1973):  5769-5784;  David 
Adamson  and  C.  L.  Fricke,  "Barium  Cloud  Evolution  and  Striation  Formation  in  the 
Magnetospheric  Release  on  Sept.  21,  1971,"  NASA  TN  D-7722,  Sept.  1974;  and  David 
Adamson,  C.  L.  Fricke,  and  Sheila  Ann  T.  Long,  "Results  of  Magnetospheric  Barium  Ion 
Cloud  Experiment  of  1971,"  NASA  TR  R-437,  Mar.  1975. 

46.  Ellis  to  author,  28  Jan.  1991,  p.  3. 

47.  The  best  source  for  the  history  of  Project  Sherwood  remains  Amasa  Bishop,  Project 
Sherwood:    The  U.S.  Program  in  Controlled  Fusion  (Reading,  Mass.:    Addison- Wesley 
Publishing  Co.,  1958).  From  1953  to  1956,  Dr.  Bishop  served  as  chief  of  research  division 
in  the  Controlled  Thermonuclear  Branch  of  the  Atomic  Energy  Commission.    Several 
aspects  of  Project  Sherwood  are  also  covered  in  Joan  Lisa  Bromberg,  Fusion:  Science, 
Politics,  and  the  Invention  of  a  New  Energy  Source  (Cambridge  and  London:  MIT  Press, 
1983). 

48.  Most  of  my  insight  into  the  nature  of  Langley's  one-megajoule  theta-pinch  apparatus 
comes  from  the  Ellis  audiotape,  author's  transcript,  p.  7.  For  a  survey  of  the  historical 
development  of  the  various  sorts  of  "pinch"  devices,  see  chap.  3  (pp.  22-32)  and  chap.  10 
(pp.  90-105),  in  Bishop,  Project  Sherwood.    Also  see  Bromberg,  Fusion,  esp.  pp.  6-7, 
19-20,   25-26,   70-71,   and  134-135;   on  the  theta-pinch  specifically,   see  pp.   143-144 
and  226-227.    For  those  interested  in  the  technical  details  of  Langley's  one-megajoule 
energy  storage  system,  a  copy  of  the  successful  Nov.  1961  proposal  for  the  system  is 
preserved  in  the  Ellis  Collection,  LHA. 

49.  For  a  general  summary  of  the  intent  of  the  Magnetic  Compression  Experiment,  see  the 
"Rough  Text  on  Laboratory  Astrophysics,"  prepared  by  Karlheinz  Thorn  as  the  proposed 
text  for  an  MPD  presentation  at  the  1964  NASA  Inspection,  copy  in  Ellis  Collection, 
LHA.    Mike  Ellis  discusses  this  experiment  on  his  audiotape,  author's  transcript,  p.  15, 
and  in  his  28  Jan.  1991  letter  to  the  author,  p.  1. 

50.  Langley's  earliest  published  papers  on  the  results  of  experiments  in  the  plasma-focus 
facility  include  the  following,  all  of  which  appeared  in  The  Bulletin  of  the  American 
Physical  Society:   J.  H.  Lee  and  C.  E.  Roos,  "An  Investigation  of  Hard  X-rays  from  a 
Plasma  Focus,"  15  (1970):  642;  J.  H.  Lee  and  L.  P.  Shomo,  "Angular  Variation  of  Neutron 
Yield  from  a  Plasma  Focus  by  Neutron-Induced  Gamma  Spectrometry,"  15  (1970):  814; 
J.  H.  Lee  and  Nelson  W.  Jalufka,  "Investigation  of  X-ray  Emission  from  a  Plasma  Focus," 
15  (1970):  815;  J.  H.  Lee  and  L.  P.  Shomo,  "Dependence  of  Neutron  Fluence  Anisotropy 
on  the  Neutron  Yield  of  a  Plasma  Focus,"  15  (1970):   1463;  and  Nelson  W.  Jalufka  and 

472 


Notes  for  Chapter  6 

J.  H.  Lee,  "Investigation  of  Current  Sheet  Collapse  in  a  Plasma  Focus  Apparatus,"  15 
(1970):  1462.  For  a  complete  list  of  publications  and  presented  papers  based  on  research 
in  Langley's  plasma-focus  facility  from  1968  to  1985,  see  the  special  bibliography  on  the 
plasma-focus  facility  that  I  have  added  to  the  Ellis  Collection,  LHA. 

51.  Macon  C.  Ellis,  Jr.,   "15  Minute  Briefing  for  Mr.  Cortright,"  handwritten  notes,  June 
1958,  copy  in  folder  labeled  "Cortright  June  24,  1968  Briefing,"  p.  4,  Ellis  Collection, 
LHA. 

52.  Ellis  audiotape,  author's  transcript,  p.  14. 

53.  Becker  interview,  5  Aug.  1991. 

54.  Ellis  audiotape,  author's  transcript,  p.  17.  For  historical  analysis  of  the  slow  and  tortuous 
progress  of  fusion  work  through  the  doldrums  in  the  1960s,  see  Bromberg,  Fusion,  esp. 
pp.  110-174. 

55.  On  the  Nixon  administration's  rejection  of  the  plan  for  a  manned  Mars  mission  in 
the  1970s  and  1980s  as  the  follow-on  to  the  Apollo  program  and  as  the  next  great 
space  challenge  accepted  by  the  United  States,  see  Walter  A.  McDougall,  The  Heavens 
and  the  Earth:    A  Political  History  of  the  Space  Age  (New  York:    Basic  Books,  1985), 
pp.  420-423;  Michael  Collins,  Liftoff:  The  Story  of  America's  Adventure  in  Space  (New 
York:  NASA/Grove  Press,  1988),  p.  201;  and  Edward  C.  Ezell  and  Linda  Neuman  Ezell, 
On  Mars:  Exploration  of  the  Red  Planet,  1958-1978,  NASA  SP-4212  (Washington,  1984), 
p.  186.   Mike  Ellis  also  comments  on  the  ramifications  of  President  Nixon's  decision  for 
the  study  of  electric  propulsion  in  his  audiotape:     "Nixon  said  that  a  manned  Mars 
mission  was  a  subject  we  could  not  talk  about.  'It  shall  be  relevant  or  it  shall  not  be  at 
all.'   Electric  propulsion  was  no  longer  relevant  to  any  practical  application;  so,  bingo, 
politically,  we  couldn't  mention  it,"  Ellis  audiotape,  author's  transcript,  p.  17. 

If  a  manned  mission  to  Mars  is  to  be  attempted  in  the  twenty-first  century,  as  many 
advocates  of  the  American  space  program  have  strongly  suggested,  it  will  require  a 
radically  new  nuclear  or  electric  propulsion  system  like  those  studied  by  NASA's  MPD 
experts  in  the  1960s.  If  that  is  the  case,  and  the  United  States  does  endeavor  to 
land  astronauts  on  Mars  sometime  during  the  next  century,  the  decision  of  the  Nixon 
administration  to  stop  basic  research  in  that  area  will  be  seen  as  a  significant  temporary 
setback. 

56.  Robert  V.  Hess  interview  with  author,  12  Nov.  1991;  Ellis  audiotape,  author's  transcript, 
p.  6. 

57.  Becker  interview,  5  Aug.  1991. 

58.  Ellis  to  author,  28  Jan.  1991,  p.  3. 

Chapter  6 
The  Odyssey  of  Project  Echo 

1.  Norman  L.  Crabill  interview  with  author,  Newport  News,  Va.,  11  Nov.  1991,  OHC,  LHA. 

2.  Ibid. 

3.  Ibid. 

4.  Howard  Gibbons,   "Earthlings  Stirred  by  NASA  Balloon,  Awesome  Sight  in  the  Sky," 
Newport  News  (Va.)  Daily  Press,  29  Oct.  1959. 


473 


Notes  for  Chapter  6 

5.  On  the  Upper  Atmosphere  Rocket  Research  Panel  and  its  earlier  incarnation  as  the 
V-2  Panel,  see  Homer  Newell,  Beyond  the  Atmosphere:   Early  Years  of  Space  Science, 
NASA  SP-4211  (Washington,  1980),  pp.  33-49,  and  John  E.  Naugle,  First  Among  Equals: 
The  Selection  of  NASA  Space  Science  Experiments,  NASA  SP-4215  (Washington,  1991), 
pp.  1-4.  Both  Newell  and  Naugle  participated  as  research  scientists  in  the  history  they 
examine  in  these  books.  A  mathematician-turned-physicist,  Newell  worked  with  the  V-2 
missiles  at  White  Sands,  New  Mexico,  from  1947  to  1955,  when  he  became  the  NRL's 
science  program  coordinator  for  Project  Vanguard.    Newell  then  joined  NASA  in  1958 
and  led  its  space  science  program  through  much  of  the  1960s.  In  1960  physicist  Naugle 
transferred  from  the  Space  Science  Division  at  NASA  Goddard,  where  he  had  been 
studying  cosmic  rays  and  protons  in  the  magnetosphere  using  nuclear  emulsions  exposed 
during  the  flights  of  upper-atmosphere  sounding  rockets,  to  NASA  headquarters  where  he 
supervised  the  space  agency's  research  into  fields  and  particles.  On  the  history  of  the  V-2, 
both  in  Germany  and  White  Sands,  see  Walter  Dornberger,  V-2 — Shot  into  the  Universe: 
The  History  of  a  Great  Invention,  trans.  James  Cleugh,  with  intro.  by  Willy  Ley  (New 
York:  Viking  Press,  1958).  Dornberger  was  the  German  army  officer  responsible  for  the 
V-2  program  at  Peenemiinde.   He  came  to  work  for  the  American  rocket  program  after 
World  War  II  as  part  of  Operation  Paperclip. 

6.  On  the  central  role  of  Project  Vanguard  in  the  U.S.  celebration  of  the  IGY,  see  Constance 
Green  and  Milton  Lomask,  Vanguard:  A  History  (Washington:  Smithsonian  Institution 
Press,   1971).     For  final  edited  versions  of  the  presentations  made  at  the  Ann  Arbor 
meeting,  see  James  Van  Allen,  ed.,   Scientific   Uses  of  Earth  Satellites  (Ann  Arbor: 
University  of  Michigan  Press,   1956).     For  a  readable  history  of  the  IGY,  see  Walter 
Sullivan,  Assault  on  the   Unknown:    The  International  Geophysical   Year,  2d  ed.  (New 
York:  McGraw-Hill,  1971). 

7.  C.  E.  Brown,  W.  J.  O'Sullivan,  Jr.,  and  C.  H.  Zimmerman,  "A  Study  of  the  Problems 
Relating  to  High  Speed,   High  Altitude  Flight,"    25  June  1953,  copy  in  the  Langley 
Technical  Library.    For  several  years,  this  report  was  classified  as  "Secret."    For  more 
on  the  history  of  the  Brown-Zimmerman-O'Sullivan  study  group,  see  James  R.  Hansen, 
Engineer  in  Charge:  The  History  of  Langley  Aeronautical  Laboratory,  1917-1958,  NASA 
SP-4305  (Washington,  1987),  pp.  351-353. 

8.  William  J.  O'Sullivan  to  Don  Murray,  Glen  Ridge,  N.J.,  29  Sept.  1960,  p.  3,  copy  in 
Milton  Ames  Collection,  box  6,  LHA.  O'Sullivan  wrote  this  letter  to  answer  questions 
that  Murray,  a  free-lance  writer,  had  asked  in  preparation  for  a  magazine  article  on 
O'Sullivan  and  the  Echo  balloon.  The  article  was  published  as  "O'Sullivan's  Wonderful 
Lead  Balloon"  in  the  Feb.  1961  issue  of  Popular  Science  Monthly.    Also,  an  interview 
with  O'Sullivan  about  the  genesis  of  inflatable  satellites,  conducted  by  Edward  Morse 
and  dated  28  Aug.  1964,  is  in  a  biography  file  concerning  O'Sullivan  in  NASA  HQA. 
Morse  eventually  drafted  a  "Preliminary  History  of  the  Origin  of  Project  Syncom,"  NASA 
HNN-40,  1  Sept.  1964,  copy  in  NASA  HQA. 

9.  O'Sullivan  to  Murray,  29  Sept.  1960,  pp.  3-4. 

10.  Ibid.,  p.  4.    See  also  O'Sullivan  to  Associate  Director  [Floyd  L.  Thompson],   "Report 
on  Project  Vanguard  and  the  IAS,  UAARP,  and  TPR  meetings  of  Jan.  25-27,  1956," 
7  Feb.  1956,  in  Project  Vanguard  file,  LCF. 

11.  O'Sullivan  to  Murray,  29  Sept.  1960,  pp.  4-5. 

474 


Notes  for  Chapter  6 

12.  Ibid.,  p.  6. 

13.  Ibid. 

14.  Ibid.,  pp.  6-7. 

15.  Ibid.,  p.  7. 

16.  For  Arthur  C.  Clarke's  thoughts  on  his  1945  article  on   "Extraterrestrial  Relays"   in 
Wireless  World,  see  his  Profiles  of  the  Future:  An  Inquiry  Into  the  Limits  of  the  Possible 
(New  York:  Popular  Library,  1977),  pp.  205-207.  Profiles  of  the  Future  first  appeared  in 
print  in  1958. 

17.  Clarke,  Profiles  of  the  Future,  pp.  205-207. 

18.  John  R.  Pierce,  "Orbital  Radio  Relays,"  Jet  Propulsion  25  (Apr.  1955):    153-157.    For 
a  later  presentation  by  Pierce,  see  "Satellite  Systems  for  Commercial  Communications," 
a  paper  presented  at  the  28th  Annual  Meeting  of  the  Institute  of  Aeronautical  Sciences 
(IAS),  New  York  City,  25-27  Jan.  1960,  copy  in  Langley  Technical  Library.  On  Pierce's 
early  ideas  for  communications  satellites,  see  Delbert  D.  Smith,   Communications  via 
Satellite:  A  Vision  in  Retrospect  (Boston:  Little,  Brown  &  Co.,  1976),  pp.  18-19. 

19.  For  contemporary  technical  distinctions  between  active  and  passive  satellites,  see  NASA, 
Space  Communications  and  Navigation,  1958-1964,  NASA  SP-93  (Washington,  1966), 
pp.  v  and  1;  see  also  Linda  Neuman  Ezell,  NASA  Historical  Data  Book,  vol.  2,  Programs 
and  Projects,  1958-1968,  NASA  SP-4012  (Washington,  1988),  p.  364.  For  insights  into 
the  nature  of  early  U.S.  satellite  proposals,  see  R.  Cargill  Hall,   "Early  U.S.  Satellite 
Proposals,"  Technology  and  Culture  4  (Fall  1963):  410-434. 

20.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  7. 

21.  Ibid. 

22.  Ibid.,   pp.   7-8.   On  the  IGY  satellite  program,   see  Newell,   Beyond  the  Atmosphere, 
pp.  46-49. 

23.  See  the  introduction  to  James  A.  Van  Allen,  ed.,  Scientific  Uses  of  Earth  Satellites  (Ann 
Arbor:  University  of  Michigan  Press,  1956);  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  8. 

24.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  8. 

25.  William  J.  O'Sullivan,  Jr.,  to  Director  [Henry  J.  E.  Reid],  "Proposal  for  the  NACA  to 
undertake  research  on  development  of  earth  satellites,"  29  June  1956;  Robert  R.  Gilruth 
to  Associate  Director  [Floyd  L.  Thompson],  "Comments  on  suggestion  for  research  on 
drag  of  earth  satellite,"  13  July  1956.  Both  letters  are  in  the  Project  Vanguard  file,  LCF. 
A  long  section  of  O'Sullivan's  1956  proposal  to  NACA  Langley  for  satellite  research  is 
quoted  in  Joseph  A.  Shortal,  A  New  Dimension:   Wallops  Flight  Test  Range,  The  First 
Fifteen  Years,  NASA  RP-1028  (Washington,  1978),  p.  600.  In  the  late  1950s  and  early 
1960s  during  the  genesis  of  inflatable  spheres  for  space  research,  Shortal  served  as  chief 
of  PARD. 

26.  O'Sullivan    to   Murray,    29    Sept.    1960,    pp.    8-9;    O'Sullivan    to   Associate   Director 
[Thompson] ,  "Report  on  Action  Taken  by  USNC/IGY  Technical  Panel  on  Earth  Satellites 
on  Sept.  5,  1956,  regarding  NACA  Proposal,"  19  Sept.  1956,  in  Project  Vanguard  file, 
LCF.   See  Newell,  Beyond  the  Atmosphere,  pp.  50  57  and  app.  E,  "Original  Membership 
of  the  U.S.  National  Committee  for  the  International  Geophysical  Year,"  p.  433. 

27.  Joseph  A.  Shortal,  memorandum  for  all  concerned,  26  Dec.  1956,  Project  Vanguard  file, 
LCF;  Walter  E.  Bressette  to  author,  11  Mar.  1992,  p.  2,  author's  files;  Shortal,  A  New 
Dimension,  p.  600. 

475 


Notes  for  Chapter  6 

28.  Murray,  "O'Sullivan's  Wonderful  Lead  Balloon,"  p.  47. 

29.  Ibid.  Murray  visited  Langley  while  writing  his  story  and  witnessed  the  balloon-folding 
procedure. 

30.  Bressette  to  author,  11  Mar.  1992,  p.  3. 

31.  "Sub-Satellite  Originated  at  Langley  Lab,"  Air  Scoop,  5  Jan.  1957,  p.  1.  For  details  about 
the  Sub-Satellite,  see  O'Sullivan,  "The  USNC/IGY-NACA  Earth  Sub-Satellite  Experi- 
ment," a  paper  presented  to  the  Comite  Speciale  de  1'Annee  Geophysique  Internationale 
(CSAGI)  Rocket  and  Satellite  Conference,  Washington,  D.C.,  30  Sept.-5  Oct.  1957,  copy 
in  Langley  Technical  Library. 

32.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.   10.     For  a  chronology  of  the   "Highlights  in 
the  Inflatable  Satellite  Program,"    including  a  reference  to  the  reconfirrnation  of  an 
allotment  of  space  for  O'Sullivan's  Sub-Satellite  on  Vanguard  by  the  Technical  Panel 
on  the  Earth  Satellite  Project,  see  "Recommendation  for  Distinguished  Service  Medal, 
Mr.  William  J.  O'Sullivan,  Jr.,  Aeronautical  Research  Engineer,  Assistant  to  Division 
Chief,  Applied  Materials  and  Physics  Division,  and  Head,  Space  Vehicles  Group,  AMPD, 
Langley  Research  Center,"  [ca.  1961],  copy  in  Milton  Ames  Collection,  box  6,  LHA. 

33.  O'Sullivan  to  Murray,  29  Sept.  1960,  pp.  9-10;  Shortal,  A  New  Dimension,  pp.  600- 
604.  For  a  chronology  of  the  development  and  operations  of  Project  Vanguard,  see  Ezell, 
NASA  Historical  Data  Book,  2:87. 

34.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  10.    A  copy  of  O'Sullivan's  San  Diego  paper, 
"The  USNC/IGY-NACA  Earth  Sub-Satellite  Experiment,"  can  be  found  in  the  Langley 
Technical  Library.  O'Sullivan  presented  an  earlier  version  of  this  paper  at  a  rocket  and 
satellite  conference  held  at  the  National  Academy  of  Sciences  in  Washington  the  previous 
month. 

35.  Bressette  to  author,  11  Mar.  1992,  p.  3;  O'Sullivan  to  Associate  Director  [Thompson], 
"NACA  12-foot  spherical  satellite,"  5  Dec.  1957,  Project  Vanguard  file,  LCF. 

36.  Walter  A.  McDougall,  The  Heavens  and  the  Earth:  A  Political  History  of  the  Space  Age 
(New  York:  Basic  Books,  1985),  p.  100. 

37.  "Moonwatchers,"  Newport  News  Daily  Press,  6  Oct.  1957,  front  page. 

38.  Quoted  in  McDougall,  Heavens  and  the  Earth,  p.  141.  A  transcript  of  an  interview  with 
Gerald  Siegel  about  Sputnik  and  other  issues  related  to  politics  and  space  is  in  NASA 
HQA. 

39.  Robert  R.  Gilruth,  "Memoir:  From  Wallops  Island  to  Mercury,  1945^1958,"  unpublished 
manuscript  presented  at  the  Sixth  International  History  Symposium,  Vienna,  Austria, 
13  Oct.   1972,   pp.   33-34,   copy  in  LHA;   Charles  J.   Donlan  interview  with  author, 
17  Aug.  1990,  transcript,  p.  6,  OHC,  LHA. 

40.  O'Sullivan  to  Associate  Director  [Thompson],  "Decisions  made  on  NACA  12-foot  satellite 
at  Apr.  3,  1958  meeting  of  TPESP,"  9  Apr.  1958,  Project  Vanguard  file,  LCF.  See  also 
O'Sullivan  to  Murray,  29  Sept.  1960,  p.  10.  For  a  summary  description  of  the  Jupiter  C's 
development,  see  Wernher  von  Braun,  Frederick  I.  Ordway  III,  and  Dave  Dooling,  Space 
Travel,  A  History:   An  Update  of  History  of  Rocketry  and  Space  Travel,  4th  ed.  (New 
York:  Harper  &  Row,  1985),  pp.  128-129. 

41.  At  the  same  time  O'Sullivan  was  directing  the  development  of  the  NACA's  12-foot 
inflatable,  he  was  also  working  on  a  pneumatically  erectable  satellite  intended  specifically 
for  navigation  purposes.    Because  a  sphere  would  reflect  radar  signals  in  all  directions, 

476 


Notes  for  Chapter  6 

O'Sullivan  realized  that  this  shape  would  not  work  as  a  navigation  satellite.  What  was 
required  instead  was  a  "corner  reflector"  that  would  reflect  signals  back  precisely  from 
whence  they  came.  With  this  basic  knowledge  of  radar  signals  and  optics  in  mind, 
O'Sullivan  designed  a  little  6.4-inch-diameter  passive  reflector  satellite  equipped  with 
tiny  solar  cells  for  flight  into  orbit  aboard  a  Vanguard.  Ground  tests  demonstrated, 
however,  that  an  active  satellite  with  a  more  powerful  transmitter  would  make  a  much 
better  navigation  satellite,  as  it  could  be  tracked  by  a  comparatively  small  tracking 
receiver  requiring  little  space  and  power  and  would  not  require  a  large  and  powerful 
radar.  O'Sullivan,  therefore,  laid  the  corner  reflector  satellite  aside  in  the  spring  of  1958 
and  concentrated  on  the  Beacon  satellite.  See  O'Sullivan  to  Murray,  29  Sept.  1960, 
pp.  10-11. 

42.  Shortal,  A  New  Dimension,  p.  607;  Eugene  M.  Emme,  Aeronautics  and  Astronautics:  An 
American  Chronology  of  Science  and  Technology  in  the  Exploration  of  Space,  1915-1960 
(Washington:    NASA,   1961),   p.   110.  Ezell,   NASA  Historical  Data  Book  (2:370-371) 
provides  a  chronology  of  the  development  and  operations  of  Project  Echo,  including 
information  on  the  30-inch  Sub-Satellite  and  12- foot  inflatable  sphere. 

43.  O'Sullivan  to  Associate  Director  [Thompson],  "Launch  vehicle  for  NACA  12-foot  satel- 
lite," 11  July  1958,  Project  Vanguard  file,  LCF.    See  also  Shortal,  A  New  Dimension, 
p.  610;   "Highlights  in  the  Inflatable  Satellite  Program,"  in  "Recommendation  for  Dis- 
tinguished Service  Medal,  Mr.  William  J.  O'Sullivan,  Jr.,  [ca.  1961],"  Ames  Collection, 
box  6,  LHA;  Bressette  to  author,  11  Mar.  1992,  p.  4.    Bressette  has  pointed  out  that 
O'Sullivan's  dogged  insistence  on  a  circular  orbit  for  the  inflatable  satellites   "nearly 
negated  the  Scout  missions  with  the  12-foot  sphere,  because  for  the  early  Scout  satellite 
launchings  the  Scout  management  required  a  6-percent  over-velocity  at  satellite  insertion. 
This  resulted  in  a  very  elliptical  orbit."  Scout  was  a  low-budget,  solid-propellant  booster 
built  by  LTV  of  Dallas,  Tex.,  which  specialized  in  putting  small  payloads  in  orbit.   As 
shall  be  explained  in  the  next  chapter  of  this  book,  NASA  Langley  served  as  project 
manager  for  Scout. 

For  details  concerning  the  deployment  and  inflation  techniques  used  for  the  12- foot 
satellite,  see  Alan  B.  Kehlet  and  Herbert  G.  Patterson,  "Free-Flight  Test  of  a  Technique 
for  Inflating  an  NASA  12-Foot-Diameter  Sphere  at  High  Altitude,"  NASA  Memo  2-5-59L, 
1959,  copy  in  the  Langley  Technical  Library.  Details  about  the  design  of  the  30-inch  and 
12- foot  inflatable  spheres  were  presented  in  a  special  NASA  report  by  Claude  W.  Coffee, 
Jr.,  Walter  E.  Bressette,  and  Gerald  M.  Keating,  Design  of  the  NASA  Lightweight 
Inflatable  Satellites  for  the  Determination  of  Atmospheric  Density  at  Extreme  Altitudes, 
NASA  TN  D-1243,  Apr.  1962. 

44.  O'Sullivan  to  Murray,  29  Sept.  1960,  pp.  14-15;  Shortal,  A  New  Dimension,  p.  610; 
von  Braun,  Ordway,  and  Dooling,  Space  Travel,  pp.  156  and  172-173;  Roger  Bilstein, 
Orders  of  Magnitude:  A  History  of  the  NACA  and  NASA,  1915-1990,  NASA  SP-4406 
(Washington,  1980),  p.  47. 

45.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  15. 

46.  Walter  E.  Bressette,  letter  to  author,  16  May  1992,  author's  files. 

47.  [Langley  Aeronautical  Laboratory],    "Preliminary  Langley  Staff  Study,   NASA  Space 
Flight  Program,"   15  May  1958,  copy  in  folder  "Space  Material,"  Floyd  L.  Thompson 
Collection,  LHA.  On  the  early  work  by  U.S.  industry  on  communications  satellites,  see 

477 


Notes  for  Chapter  6 

Smith,   Communications  via  Satellite,  pp.  58-59;  McDougall,  Heavens  and  the  Earth, 
p.  352. 

48.  O' Sullivan  to  Murray,  29  Sept.  1960,  p.  11. 

49.  Shortal,  A  New  Dimension,  p.  686;  Bressette  to  author,  11  Mar.  1992,  p.  4. 

50.  Research  Authorization  (RA)  A12L145,  dated  8  May  1958,  in  Research  Authorization 
Files,  LHA.  See  also  Shortal,  A  New  Dimension,  p.  686.    For  an  outline  of  the  early 
organization  at  Langley  for  Echo,  see  Jesse  L.  Mitchell  to  Associate  Director  [Thompson], 
"NACA  100-Foot  Diameter  Communications  Satellite  Experiment,"  copy  in  Project  Echo 
file,  LCF.    See  also  Shortal,  A  New  Dimension,  pp.  686-687. 

51.  William  J.  O'Sullivan  to  Associate  Director  [Floyd  L.  Thompson],  "Meeting  on  estab- 
lishment of  a  task  group  to  carry  work  load  on  30-inch,  12-foot,  and  100-foot  satellite, 
held  on  February  18,  1959,"  19  Feb.  1959,  Project  Echo  file,  LCF.    For  NASA's  public 
justification  for  Project  Echo,  see  "NASA  Urges  Project  Echo  Participation,"  Aviation 

Week,  71  (14  Dec.  1959):  65-67. 

52.  Leonard  Jaffe,   chief,   Communications  Satellite  Program,    "Project  Echo:    Results  of 
Second  Planning  Meeting,"  13  Oct.  1959,  Project  Echo  file,  LCF.    The  first  Echo  plan- 
ning meeting  had  taken  place  on  10  June  1959;  see  Leonard  Jaffe,  "Results  of  Project 
Echo  Planning  Meeting  on  June  10,   1959,"    16  June  1959,  in  Project  Echo  file.     On 
JPL's  part  in  the  early  stages  of  Project  Echo,  see  Walter  K.  Victor  and  Robertson 
Stevens,    "The  Role  of  the  Jet  Propulsion  Laboratory  in  Project  Echo,"    JPL  Tech. 
Release  34-141,    14  Oct.   1960;   on  the  Goldstone  receiving  station,   see  Stevens  and 
Victor,  "The  Goldstone  Station  Communications  and  Tracking  System  for  Project  Echo," 
JPL  TR  32-59,   1  Dec.   1960.     Copies  of  both  are  in  the  Langley  Technical  Library. 
On  Bell  Telephone's  role,  see  William  C.  Jakes,  Jr.,  Bell  Labs,   "Participation  of  Bell 
Telephone  Laboratories  in  Project  Echo  and  Experimental  Results,"  NASA  TN  D-1127, 
Dec.   1961.     The  July  1961  issue  of  the  Bell  System  Technical  Journal  is  devoted  to 
extensive  discussion  of  Project  Echo  and  Bell's  participation  in  it. 

53.  Robert  J.  Mackey,  Jr.,  head,  Communications  Branch,  NASA  Goddard,   "Echo  A- 12 
Working  Agreement:  GSFC-LRC,"  4  Jan.  1961,  Project  Echo  file,  LCF;  Shortal,  A  New 
Dimension,  p.  689.     At  the  second  Project  Echo  planning  meeting  on  13  Oct.  1959, 
Leonard  Jaffe  had  named  Goddard's  Robert  Mackey  to  be  the  "action  officer"  for  Project 
Echo. 

54.  On  Project  Shotput,  see  Shortal,  A  New  Dimension,  pp.  685  695.  The  author  would  like 
to  thank  Norman  L.  Crabill,  one  of  the  heads  of  the  Shotput  program  at  Langley,  for 
sharing  his  personal  papers  relevant  to  Shotput,  for  example,  his  unpublished  talk  given 
at  Langley  on  24  May  1961  to  the  NASA/DOD  Technical  Committee  on  Communications 
Satellites,  "The  Echo  A-12  Vertical  Test  Program."    The  complete  documentary  record 
of  Shotput  is  preserved  in  the  Project  Shotput  file,  LCF. 

55.  O'Sullivan  to  Murray,  29  Sept.  1960,  pp.  12  13. 

56.  Ibid.,  pp.  13-14;  Shortal,  A  New  Dimension,  pp.  690-691. 

57.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  14. 

58.  Norman  L.  Crabill  interview,  11  Nov.  1991. 

59.  Bressette  to  author,    11   Mar.    1992,    pp.   4-5.      On   19   May   1961,    the  president   of 
Schjeldahl,  G.  T.  Schjeldahl,  testified  about  his  company's  involvement  with  "Erectable 
and  Inflatable  Structures  in  Space"  at  a  Hearing  before  the  Committee  on  Science  and 

478 


Notes  for  Chapter  6 

Astronautics,  U.S.  House  of  Representatives,  87th  Congress,  First  Session,  May  19, 
1961,  no.  12  (Washington:  Government  Printing  Office,  1961),  pp.  2-5.  At  this  same 
hearing,  Laurence  K.  Loftin,  Jr.,  technical  assistant  to  Langley  Director  Floyd  Thompson, 
and  William  J.  O'Sullivan,  Jr.,  head  of  the  Space  Vehicle  Group  at  Langley,  also  gave 
testimony  concerning  inflatable  structures  in  space,  including  structures  that  could  be 
used  as  space  stations.  Anyone  interested  in  the  manufacturing  of  the  balloons  should 
see  G.  T.  Schjeldahl  Co.,  "Design  and  Inflation  of  Inflatable  and  Rigidizable  Passive 
Communication  Satellites  (Echo  I  and  Echo  II),"  in  Aerospace  Expandable  Structures 
(Mar.  1964):  576-604. 

60.  Crabill  interview,  11  Nov.  1991;  Shortal,  A  New  Dimension,  p.  690.    By  the  early  1960s, 
Norman  Crabill  had  become  an  expert  on  the  ascent  problems  of  sounding  rockets  and 
other  launch  vehicles;  see  his  paper,  "Ascent  Problems  of  Sounding  Rockets,"  presented 
at  an  AGARD  meeting  on  the  Use  of  Rocket  Vehicles  in  Flight  Research,  Scheveningen, 
the  Netherlands,  18-21  July  1961,  copy  in  Langley  Technical  Library. 

61.  O'Sullivan  to  Associate  Director  [Thompson],  "Results  of  Shotput  test  no.  1,"  16  Nov. 
1959,  in  Project  Echo  file,  LCF.    For  technical  analysis  of  the  launch  problems  involved 
with  Shotput  1,  see  Robert  L.  James,  Jr.,  and  Ronald  J.  Harris,  Calculation  of  Wind  Com- 
pensation for  Launching  of  Unguided  Rockets,  NASA  TN  D-645,  Apr.  1961;  Robert  L. 
James,  Jr.,  A  Three- Dimensional  Trajectory  Simulation  Using  Six  Degrees  of  Freedom 
with  Arbitrary  Wind,  NASA  TN  D-641,  Mar.  1961;  and  Sherman  A.  Clevenson,  memoran- 
dum for  files,  "Spin-balancing  and  alignment  of  the  second  stage  of  the  first  SHOT  PUT 
at  Wallops  Island,  Virginia,"  4  Nov.  1959,  Project  Shotput  file,  LCF.  See  also  Norman  L. 
Crabill,  Ascent  Problems  of  Sounding  Rockets,  AGARD  Report  391,  July  1961. 

62.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  12. 

63.  Ibid.;  Shortal,  A  New  Dimension,  p.  691. 

64.  O'Sullivan  to  Murray,  29  Sept.  1960,  p.  13;  O'Sullivan  to  Associate  Director  [Thompson], 
"Echo  satellite  collapsing  and  inflation  pressures,"  19  Jan.  1960,  Project  Echo  file,  LCF. 

65.  For  details  about  the  launch-vehicle  problem  in  Shotput  2,  see  Norman  L.  Crabill  to 
Clarence  L.  Gillis,  "Preliminary  data  and  analysis  for  the  SHOTPUT  No.  2  vehicle,  fired 
Jan.  16,  1960,  at  5:30:00.108  p.m.  EST,"  29  Jan.  1960.  On  the  balloon  tear  in  Shotput  3, 
see  C.  D.  Bailey,  memorandum  for  files,  "Shotput  3,"  29  Feb.  1960,  as  well  as  "NASA 
Trajectory  Data  for  Shot  Put  3,  Flown  Feb.  27,  1960,  at  6:20:54.3  p.m.  EST,"  9  May  1960. 
These  documents  are  in  the  Project  Shotput  file,  LCF. 

66.  Crabill  interview,  11  Nov.  1991;  Shortal,  A  New  Dimension,  pp.  690-692.    For  general 
information  and  the  public  reaction  to  Shotputs  2  and  3,  see  "Wallops  Launches  Inflatable 
Satellite,   Visible  from  Afar,"    Newport  News  Daily  Press,   17  Jan.   1960;    "Recorded 
Message  Bounced  Off  NASA's  Space  Sphere,"  Daily  Press,  28  Feb.  1960.  Many  technical 
documents  relating  to  the  trajectory  and  tracking  of  the  Shotputs  are  in  the  Project 
Shotput  file,  LCF. 

67.  Edwin  P.  Kilgore  to  author,  17  Feb.  1992,  author's  files. 

68.  Ibid.;  Shortal,  A  New  Dimension,  p.  693. 

69.  Crabill  interview,  11  Nov.  1991. 

70.  Crabill  interview,  11  Nov.  1991;  Shortal,  A  New  Dimension,  pp.  689-690.  Bressette  and 
Coffee  conducted  an  analysis  in  the  spring  of  1959  to  determine  the  gas  leakage  to  be 
expected  from  meteoroid  puncture,  concluding  that  the  balloon  should  retain  sufficient 

479 


Notes  for  Chapter  6 

pressure  to  remain  spherical  for  at  least  seven  days.  For  their  optimistic  findings,  see 
their  memorandum  for  Space  Vehicle  Branch  files,  "Gas  leakage  from  the  100-foot  sphere 
caused  by  meteoroid  puncture,"  22  May  1959,  in  Project  Echo  file,  LCF. 

71.  President  Eisenhower's  message  is  quoted  in  full  in  "Highlights  in  the  Inflatable  Satellite 
Program,"  p.  4,  box  5,  Milton  Ames  Collection,  LHA. 

72.  For  these  articles,  see  the  front  pages  of  the  Newport  News  Daily  Press,  12-16  Aug.  1960. 
On  14  Aug.,  the  Chesapeake  and  Potomac  Telephone  Company  of  Virginia  ran  a  full-page 
ad  in  the  Daily  Press  in  celebration  of  Echo  and  the  "First  Phone  Call  Via  a  Man-Made 
Satellite." 

73.  Crabill  interview,  11  Nov.  1991.  For  a  complete  technical  overview  of  the  Echo  1  mission 
from  the  point  of  view  of  the  launch- vehicle  contractor,  see  Douglas  Aircraft  Co.,  Inc., 
"Detailed  Test  Objectives  for  Launch  Vehicle  Mission  No.   1 — Echo  1  Payload  Delta 
Program,"  Feb.  1960,  copy  in  Langley  Technical  Library. 

74.  Leonard  Jaffe,  "Project  Echo  Results,"  Astronautics  6  (May  1961):  32-33,  80.  See  also 
Shortal,   A  New  Dimension,  pp.  694-695,  and  Alfred  Rosenthal,    Venture  into  Space, 
NASA  SP-4301  (Washington,  1968),  p.  89. 

75.  See  George  R.  Thompson,   "NASA's  Role  in  the  Development  of  Commercial  Satellite 
Communications,"  NASA  HHM-8,  31  Jan.  1966;  Bilstein,  Orders  of  Magnitude,  pp.  56- 
57  and  65-65.  For  basic  information  about  the  NASA  satellite  communications  programs 
of  the  1960s,  see  Ezell,  NASA  Historical  Data  Book,  2:364-394.  For  the  contemporary 
thinking  on  satellite  programs,  see  "The  NASA  Communications  Satellite  Program," 
9  Feb.  1961,  copy  in  Langley  Technical  Library,  code  N-96842;  this  report  reviews  NASA 
objectives  in  passive  and  active  communications  satellite  programs  as  of  early  1961.  See 
also  Leonard  Jaffe,  "NASA  Communications  Satellite  Program,"  in  Proceedings  of  the 
Second  National  Conference  on  the  Peaceful  Uses  of  Space,  Seattle,  May  8-10,   1962 
(Nov.  1962):  103-114,  copy  in  Langley  Technical  Library. 

76.  Bressette  to  author,  16  May  1992. 

77.  See  Bressette,  Study  of  a  Passive  Communication,  Gravity-Gradient  Stabilized,  Lenticular 
Satellite,  Jan.   1965,  copy  in  Langley  Technical  Library,  GER  11893.     For  Langley 's 
commitment  to  passive  reflectors  in  the  early  1960s,  see  Warren  Gillespie,  "An  Earth- 
Oriented  Communication  Satellite  of  the  Passive  Type,"  in  Advances  in  Astronautical 
Sciences  6  (1961):  51-63. 

78.  Newell,  Beyond  the  Atmosphere,  pp.  144-153.  For  information  on  the  Explorer  missions, 
see  also  Ezell,  NASA  Historical  Data  Book,  2:232-253;  Helen  T.  Wells,  Susan  H.  Whitely, 
and  Carrie  E.  Karegeannes,  "Explorer,"    in  Origins  of  NASA  Names,  NASA  SP-4402 
(Washington,    1976),   pp.  49-52.  For  technical  details  of  the  Explorer  9  mission,   see 
O'Sullivan,  Claude  W.  Coffee,  Jr.,  and  Gerald  M.  Keating,  "Air  Density  Measurements 
from  the  Explorer  9   Satellite,"    paper  presented  at  a  Meeting  of  the  International 
Committee  on  Space  Research,  Washington,  D.C.,  30  Apr. -11  May  1962,  copy  in  Langley 
Technical  Library. 

79.  Ezell,  NASA  Historical  Data  Book,  2:298,  400.  See  also  Coffee,  Bressette,  and  Keating, 
Design  of  the  NASA  Lightweight  Inflatable  Satellites,  NASA  TN  D-1243,  Apr.  1962. 

80.  Von  Braun,  Ordway,  and  Dooling,  Space  Travel,  pp.  174-175. 

81.  In  keeping  with  a  historic  and  unique  national  pattern  of  competitive  private  ownership 
of  the  means  of  communications,  it  was  NASA's  job  to  help  private  enterprise  set  up 

480 


Notes  for  Chapter  7 

an  effective  commercial  comsat  system  through  its  state-funded  R&D,  not  to  create  the 
entire  system  itself.  A  rather  flexible  policy  directive  from  the  Eisenhower  administration 
supporting  the  early  establishment  of  a  commercial  comsat  system  led  to  NASA's  joint 
project  with  AT&T  for  Telstar  but  still  left  room  for  NASA's  own  development  of  Relay. 
In  May  1961,  however,  the  political  framework  for  NASA's  participation  in  the  comsat 
business  changed  dramatically  with  President  Kennedy's  announcement  of  his  ambitious 
plan  for  a  single  global  comsat  system.  This  announcement  was  made  in  his  lunar  landing 
speech.  For  analysis  of  these  developments  involving  NASA  and  commercial  comsats,  see 
McDougall,  Heavens  and  the  Earth,  pp.  352-359. 

82.  Von  Braun,  Ordway,  and  Dooling,  Space  Travel,  p.  174. 

83.  Bilstein,  Orders  of  Magnitude,  pp.  65-66. 

84.  Arthur  C.  Clarke,  "A  Short  Pre-History  of  Comsats,  Or:  How  I  Lost  a  Billion  Dollars  in 
My  Spare  Time,"  in  Voices  from  the  Sky:  Previews  of  the  Coming  Space  Age  (New  York, 
1965),  pp.  119-128. 

85.  Crabill  interview,  11  Nov.  1991. 

Chapter  7 

Learning  Through  Failure:  The  Early  Rush 
of  the  Scout  Rocket  Program 

1.  Abraham  Leiss  interview  with  author,  Hampton,  Va.,  19  July  1990;  see  also  Roland  D. 
English  interview  with  David  Ferraro,  Hampton,  Va.,  14  Dec.  1990.    Transcripts  of  all 
Ferraro's  interviews,  which  were  conducted  in  preparation  for  a  1991  NASA  video  on  the 
history  of  Scout,  are  located  in  the  OHC,  LHA.  The  author  wishes  to  thank  Mr.  Ferraro, 
a  talented  producer  of  video  documentaries,  for  generously  sharing  his  many  interviews 
concerning  Scout.  The  essential  documents  regarding  the  history  of  Scout  can  be  found 
in  the  Scout  Project  files,  LHA,  which  are  organized  chronologically. 

Fortunately,  personnel  in  the  Scout  Project  Office  (notably  Abraham  Leiss),  because 
of  their  understandable  pride  in  the  many  accomplishments  of  the  Scout  rocket,  kept 
detailed  scrapbooks  with  clippings,  documents,  photos,  and  other  miscellaneous  materials 
central  to  the  history  of  Scout.  This  chapter  offers  only  a  brief  look  into  that  history 
during  its  troubled  early  period.  A  comprehensive  scholarly  history  of  the  Scout  program 
has  not  yet  been  written. 

2.  For  a  detailed  history  of  PARD's  work  at  Wallops  Island,  see  Joseph  A.  Shortal,  A 
New  Dimension:   Wallops  Island  Flight  Test  Range,  the  First  Fifteen  Years,  NASA  RP- 
1028  (Washington,  Dec.  1978).   Shortal  was  the  second  head  of  PARD  operations  (and 
subsequently  of  the  Applied  Materials  and  Physics  Division)  at  Wallops  Island  from 
1951  to  his  retirement  in  the  late  1960s.  His  history  is  especially  valuable  as  an  insider's 
account  of  what  went  on  at  Wallops  Island  just  before  and  after  the  start  of  the  spaceflight 
revolution. 

3.  On  Vanguard,   nothing  has  yet  surpassed  Constance  M.  Green  and  Milton  Lomask, 
Vanguard:    A  History,  NASA  SP-4202  (Washington,   1970);   on  the  IGY,  see  Sydney 
Chapman,  IGY:  Year  of  Discovery  (Ann  Arbor:    University  of  Michigan  Press,  1959); 
Walter  Sullivan,  Assault  on  the  Unknown:    The  IGY  Discovery  (New  York:    McGraw- 
Hill,  1961);  and  J.  Tuzo  Wilson,  IGY:    The  Year  of  the  New  Moons  (New  York:  Knopf, 

481 


Notes  for  Chapter  7 

1961).  For  a  concise  and  insightful  analysis  of  the  IGY  history,  see  Walter  A.  McDougall, 
The  Heavens  and  the  Earth:  A  Political  History  of  the  Space  Age  (New  York:  Basic 
Books,  1985),  pp.  112-140. 

4.  English  interview  with  Ferraro,   14  Dec.  1990;  James  R.  Hall  interview  with  Ferraro, 
Hampton,  Va.,  4  Dec.  1990. 

5.  Shortal,  A  New  Dimension,  pp.  706-707. 

6.  Ibid. 

7.  Remarks  made  by  William  E.  Stoney  to  Ferraro  and  used  in  the  1991  NASA  video, 
"Scout:  The  Unsung  Hero  of  Space." 

8.  Shortal,  A  New  Dimension,  p.  707. 

9.  NASA  Space  Flight  Program,  Preliminary  Langley  Staff  Study,  15  May  1958,  pp.  1  and 
28,  Scout  Project  files,  LHA. 

10.  This  research  authorization  (RA)  is  not  in  the  RA  files  in  the  LHA,  which  for  some 
unknown  reason  only  encompass  the  early  1950s.  For  a  discussion  of  this  RA,  see  Shortal, 
A  New  Dimension,  p.  707. 

11.  This  air  force  effort  to  develop  an  advanced  solid- fuel  rocket  launch  vehicle  was  known 
to  the  DOD  as  "Project  609 A"  (Hyper-Environmental  Test  System  609A).    For  details 
of  the  agreement  between  the  air  force  and  the  NACA  (and  later  NASA),  see  Clotaire 
Wood  to  Langley  Director  [Reid],  11  July  1958,  in  the  Scout  Project  files,  LHA.  Enclosed 
in  this  letter  is  a  document  entitled  "Specifications  for  Solid-Fuel  Rocket  Test  Vehicle," 
dated  a  day  earlier.    See  also  "Aeroneutronic  Wins  New  Space  Contract,  AF  Project 
Involves  High-Altitude  Tests,"  Aeroneutronic  News,  Newport  Beach,  Calif.,  9  May  1960. 
For  an  example  of  Scout's  early  role  in  the  air  force's  nuclear  weapons  testing  program, 
see  the  10  Oct.  1960  issue  of  Aviation  Week  and  Space  Technology;  its  lengthy  cover 
story  concerns  the  air  force's  first  launch  of  a  "Blue  Scout"  rocket.    See  also  "Nuclear 
Detector  Sent  Up,  Blue  Scout's  1st  Flight,"  Norfolk-Portsmouth  (Va.)   Virginian- Pilot, 
22  Sept.  1960. 

12.  Shortal,  A  New  Dimension,  p.  707. 

13.  Ibid.,  p.  708;  see  also  Ferraro  letter  to  author,  9  Sept.  1991,  copy  in  LHA. 

14.  Warren  J.  North,  chief,  Manned  Satellite  Program,  to  NASA  Administrator  (T.  Keith 
Glennan),    "Status  Report  No.   1 — Scout,"   3  Mar.   1959,  copy  in  Scout  Project  files, 
LHA.  See  also  "NASA  Awards  Scout  Contracts,"   Aviation  Week,  2  Mar.  1959,  p.  2; 
and  "First  Scout  Vehicle  Delivery  Due  Aug.  15,"  Aviation  Week,  16  Mar.  1959,  p.  26. 
For  a  concise  and  well-illustrated  introduction  to  the  details  of  the  Scout  rocket,  see  LTV 
publication,   The  Scout:    Solid  Propellant  Launch  Vehicles  (Dallas,  1962),  copy  in  the 
Floyd  L.  Thompson  Collection,  LHA,  folder  labeled  "Scout."    See  also  Shortal,  A  New 
Dimension,  pp.  708-710. 

For  an  excellent  technical  summary  of  the  evolution  of  the  Scout  program  to  1968  by 
someone  intimately  involved  with  it,  see  Milt  Green,  "Project  Review — Scout  Launch 
Vehicle,"  in  Spacecraft  Systems  Education,  2  vols.  (Belgrade,  [formerly]  Yugoslavia: 
Pergamon  Press,  1968),  2:3-15.  Green  was  the  program  director  for  the  Scout  Missiles 
and  Space  Division  of  LTV  in  Dallas. 

15.  On  the  X-l  (or  more  technically,  XS-1)  program  at  Langley,  see  chap.  10  of  James  R. 
Hansen,    Engineer  in    Charge:     A    History   of  the   Langley   Aeronautical   Laboratory, 


482 


Notes  for  Chapter  7 

1917-1958,  NASA  SP-4305  (Washington,  1987),  pp.  271-310;  for  more  general  histo- 
ries of  the  X-l  flight  test  program,  see  Richard  P.  Hallion,  Supersonic  Flight:  The  Story 
of  the  Bell  X-l  and  Douglas  D-558  (New  York  &  London:  Macmillan  Co.,  1972)  and 
On  the  Frontier:  Flight  Research  at  Dryden,  1946-1981,  NASA  SP-4303  (Washington, 
1984),  esp.  chap.  1,  pp.  3-22. 

16.  James  R.  Hall  quoted  in  "Scout  Moves  to  Goddard:    Leaves  Legacy  of  Camaraderie," 
Langley  Researcher  News,  25  Jan.  1991,  p.  7. 

17.  H.  J.  E.  Reid,  memorandum  for  all  concerned,  "Establishment  of  Scout  Project  Group," 
29  Feb.  1960,  in  Scout  Project  files,  LHA.  On  Scout's  organization,  see  also  Eugene  C. 
Draley,  memorandum  for  files,  "Scout  Organization,"  4  Mar.  1960,  Scout  Project  files, 
LHA.  This  group  was  already  located  on  the  second  floor  of  the  impact  basin  building 
on  Langley's  west  side.    The  individuals  assigned  to  Scout  from  the  various  Langley 
research  divisions  were:    Carl  A.  Sandahl  and  Andrew  G.  Swanson  from  the  Applied 
Materials  and  Physics  Division;  William  J.  Nelson  from  the  Aero-Physics  Division;  Max 
C.  Kurbjun  from  the  Aero-Space  Mechanics  Division;  Harry  L.  Runyan,  Jr.,  from  the 
Dynamic  Loads  Division;  Axel  T.  Mattson  from  the  Full-Scale  Research  Division;  James 
E.  Stitt  from  IRD;  Leonard  Sternfield  from  the  Theoretical  Mechanics  Division;  Isidore 
G.  Recant  from  the  Unitary  Plan  Wind  Tunnel  Division;  and  Edwin  C.  Kilgore  from  the 
Engineering  Service  Division.  One  individual  each  was  also  assigned  from  the  Mechanical 
Service  Division,  the  Electrical  Service  Division,  the  Procurement  Division,  as  well  as 
from  Maintenance.    The  project  office  gave  regular  written  status  reports  to  Langley 
management,  copies  of  which  are  preserved  in  the  Scout  Project  files,  LHA. 

18.  Hall  interview  with  Ferraro,   4  Dec.   1990.     In   1959  two  Langley  researchers  in  the 
Theoretical  Mechanics  Division,   Robert  H.  Tolson  and  Wilbur  C.  Mayo,   conducted 
"Studies  of  Use  of  the  Scout  as  a  Booster  Vehicle  for  Lunar  Exploration."    For  more 
about  the  lunar  exploration  studies  based  in  this  Langley  research  division,  see  the  next 
chapter. 

19.  Leiss  interview,  19  July  1990. 

20.  Hall  quoted  in  "Scout  Moves  to  Goddard,"  Langley  Researcher  News,  25  Jan.  1991,  p.  7. 
For  basic  information  on  the  Scout  Project  in  its  very  early  period,  see  the  unofficial 
"Scout  Manual,"  prepared  by  the  Scout  Project  Office  at  Langley  on  25  Apr.  1960  and 
distributed  to  all  Langley  Scout  team  members.     The  cover  of  this  manual  shows  a 
triangle- faced  Boy  Scout  saluting  in  front  of  a  Scout  rocket.  A  copy  of  this  manual  and 
all  subsequent  Scout  manuals  are  in  the  notebook  "Scout  Project,  1959-1961,"   LHA. 
Such  notebooks,  unofficial  scrapbooks,  were  kept  by  the  Scout  Project  Office  for  the 
entire  period  of  Scout's  operation. 

21.  Shortal,  A  New  Dimension,  pp.  709-710. 

22.  Hall  interview  with  Ferraro,  4  Dec.  1990;  Ken  Jacobs  interview  with  Ferraro,  Hampton, 
Va.,  29  Nov.  1990;  Milt  Green  interview  with  Ferraro,  Hampton,  Va.,  28  Nov.  1990. 

23.  Larry  Tant  interview  with  Ferraro,  Hampton,  Va.,  14  Dec.  1990;  Jon  Van  Cleve  interview 
with  Ferraro,  Hampton,  Va.,  16  Nov.  1990. 

24.  William  E.  Stoney,  Jr.,  and  Edwin  C.  Kilgore,  memorandum  for  the  Langley  Associate 
Director  (Floyd  L.  Thompson),  "Outline  of  steps  required  to  insure  an  earliest  possible 
firing  of  the  Scout  and  those  required  to  make  possible  subsequent  firings  at  intervals 
of  not  over  two  months,"   28  Jan.   1960;   H.  J.  E.  Reid,  Langley  director,  to  NASA, 

483 


Notes  for  Chapter  7 

Code  R  (Research),  "Acceleration  of  Scout  Program  to  insure  earliest  possible  firing 
and  to  make  possible  subsequent  firings  at  intervals  of  not  over  2  months,"  29  Jan.  1960. 
Both  documents  are  in  the  Scout  Project  files,  LHA.  See  also  Shortal,  A  New  Dimension, 
p.  717. 

25.  Hall  quoted  in  "Scout  Moves  to  Goddard,"  p.  7. 

26.  Shortal,  A  New  Dimension,  pp.  717-718. 

27.  Ibid.,  p.  718;  "NASA  Launches  'Scout'  Vehicle  in  Wallops  Test,"  Newport  News  (Va.) 
Daily  Press,  19  Apr.  1960;   "Ignition  Failure  in  Third  Stage  Aborts  Initial  Scout  Test 
Launch,"  Daily  Press,  25  Apr.  1960. 

28.  Hall  quoted  in  "Scout  Moves  to  Goddard,"  p.  7. 

29.  Ibid. 

30.  Ibid.;   "Scout  Veers,  Is  Destroyed  After  Firing  From  Wallops,"   Baltimore  Sun,  2  July 
1960.  See  also  Robert  J.  Mayhue,  Scout  ST-1  Flight-Test  Results  and  Analyses,  Launch 
Operations,  and  Test  Vehicle  Description,  NASA  TN-1240  (Washington,  June  1962),  and 
Shortal,  A  New  Dimension,  p.  720. 

31.  Hall  quoted  in    "Scout   Moves  to   Goddard,"    p.   7.      For  details  of  the  launch,    see 
"Preliminary  Flight  Report  on  Scout  ST-1,"  Dec.  1960,  Scout  Project  files,  LHA. 

32.  "Obtain  Valuable  Data  in  First  Scout  Launching,"   NASA  Langley  Air  Scoop,  8  July 
1960;   "NASA  Fires  Scout,"   Newport  News  (Va.)   Times  Herald,  2  July  1960.    Copies 
of  the  press  releases  given  out  by  NASA  after  the  launch  are  preserved  in  the  "Scout 
Project,  1959-1961"  notebook,  LHA.    Also  see  Shortal,  A  New  Dimension,  p.  720. 

33.  "Preliminary  Flight  Test  Results  of  Scout  ST-2,"  Dec.  1960,  Scout  Project  files,  LHA. 
On  the  military  launches,  see  "USAF  Launches  Second  Blue  Scout,"   Aviation  Week, 
21  Nov.  1960,  p.  30.  Copies  of  all  newspaper  articles  mentioned  in  the  text  can  be  found 
in  the  notebook,  "Scout  Project,  1959-1961,"  LHA. 

34.  Vought  Corp.  brochure,   "100  Scout  Launches,   1960-1979,"   published  for  program  in 
celebration  of  100th  Scout,  held  in  Williamsburg,  Va.,  27  July  1979. 

35.  English  interview  with  Ferraro,  14  Dec.  1990. 

36.  Hall  quoted  in  "Scout  Moves  to  Goddard,"  p.  7. 

37.  "Scout  Shot  Veers  Off,  Destroyed,"  Newport  News  Daily  Press,  21  July  1963;  English 
interview  with  Ferraro,  14  Dec.  1990.   For  details,  see  LTV,  "NASA  Scout  S-110,  Final 
Flight  Report,"  7  Aug.  1963,  report  code  3-30000/3R-233,  copy  in  Scout  Project  files, 
LHA. 

38.  Tom  Perry  interview  with  Ferraro,  Hampton,  Va.,  14  Dec.  1990. 

39.  Eventually  two  Scout  review  committees  were  formed.    The  first,  created  on  22  July 
1963,  was  chaired  by  Richard  R.  Heldenfels,  the  hard-nosed  director  of  the  Structures 
Research  Division  at  Langley.    The  other  six  members  of  this  review  committee  were 
Langley  engineers  Edward  M.  Gregory,  Edwin  C.  Kilgore,  Clifford  H.  Nelson,  Eugene  D. 
Schult,  Joseph  A.  Shortal,  and  Joseph  G.  Thibodaux,  Jr.   This  committee  was  to  look 
specifically  into  the  most  recent  Scout  110  failure.  The  second  committee,  an  intercenter 
group  chaired  by  Richard  B.  Morrison  of  NASA  headquarters,  was  to  look  more  generally 
into  the  problems  of  the  Scout  program.  The  other  members  of  this  panel  included  Eugene 
C.  Draley  of  Langley;  Hermann  W.  Kroeger  of  Marshall  Space  Flight  Center;  Daniel  G. 
Mazur  of  Goddard  Space  Flight  Center;  and  William  A.  Fleming,  Milton  W.  Rosen,  and 
Warren  A.  Guild,  all  of  NASA  headquarters.   For  details  on  the  second  committee,  see 

484 


Notes  for  Chapter  7 

Homer  E.  Newell,  director,  Office  of  Space  Sciences,  NASA  headquarters,  to  director, 
Langley  Research  Center,  "Scout  Review  Committee,"  9  Oct.  1963,  in  Scout  Project 
files,  LHA.  All  the  documents  collected  by  the  Langley  committee,  all  the  minutes  of 
meetings,  plus  the  final  reports  and  recommendations  of  both  committees  can  be  found 
in  the  Scout  Project  files,  LHA. 

40.  Hall  interview  with  Ferraro,  4  Dec.  1990. 

41.  Eugene  D.  Schult  interview  with  Ferraro,  Hampton,  Va.,  6  Dec.  1990. 

42.  Jacobs  interview  with  Ferraro,  29  Nov.  1990;  English  interview  with  Ferraro,  14  Dec.  1990. 

43.  Jacobs  interview  with  Ferraro,  29  Nov.  1990.  The  ups  and  downs  of  the  Scout  program 
can  best  be  followed  by  surveying  the  "Scout  Vehicle  Status  Reports."    These  reports 
were  reviewed  after  every  one  or  two  flights.  Copies  of  all  these  reports  are  in  the  Scout 
Project  files,  LHA. 

44.  Hall  interview  with  Ferraro,  4  Dec.  1990. 

45.  See  Langley  Research  Center,  Project  Development  Plan:    Scout  Program,  Jan.  1964. 
This  100-page  document  expresses  the  essentials  of  the  recertification  process.    Other 
documents  from  the  recertification  process  include  "Scout  Engineering  Program  Plan," 
revised  4  Mar.  1964,  and  "NASA/DOD  Scout  Launch  Complex  Maintenance  Manual," 

2  vols.,  15  Feb.  1964.  All  these  documents  are  in  the  Scout  Project  files,  LHA. 

46.  Green  interview  with  Ferraro,  28  Nov.  1990. 

47.  Schult  interview  with  Ferraro,  6  Dec.  1990. 

48.  English  interview  with  Ferraro,  14  Dec.  1990. 

49.  LTV  Missiles  and  Electronics  Group  brochure,  "Scout:  30  Years  of  Service,  1960-1990," 
(Dallas,  Feb.  1991).  For  more  basic  historical  information  on  Scout,  see  Linda  Neuman 
Ezell,   NASA  Historical  Data  Book,  vol.  2,   Programs  and  Project  1958-1968,  NASA 
SP-4012  (Washington,  1988),  pp.  19-21,  61-67,  244-247,  444  448,  and  465-477.    For 
a  comprehensive  listing  of  the  technical  reports  published  by  NASA  and  its  contractors 
relevant  to  Scout,  see  Abraham  Leiss,   Scout  Launch   Vehicle  Program,   Final  Report, 

3  vols.  (NASA  Langley:  NASA  CR-165950,  May  1982),  3:1122-1206. 

50.  The  preceding  summary  of  Scout's  accomplishments  has  been  composed  largely  from 
a  survey  of  the  newspaper  clippings,  magazine  stories,  documents,  and  miscellaneous 
materials  in  the  Scout  Project  Office  scrapbooks.    For  more  exhaustive  details  on  the 
Scout  launch  vehicle  program,  including  the  San  Marco  operation,  see  Abraham  Leiss, 
Scout  Launch  Vehicle  Program.    Mr.  Leiss,  as  the  chief  administrative  assistant  of  the 
Scout  Project  Office  from  its  inception  in  1960,  has  collected  in  these  three  volumes 
everything  any  researcher  could  possibly  want  to  know  about  Scout.    For  an  excellent 
short  history  of  the  Scout  program  to  1979,  see  Andrew  Wilson,  "Scout:  NASA's  Small 
Satellite  Launcher,"  Spaceflight  21  (Nov.  1979):  446-459. 

51.  See  "Scout  Financial  Data"  in  Leiss,  Scout  Launch  Vehicle  Program,  1:268-270. 

52.  Perry  interview  with  Ferraro,  14  Dec.  1990.    David  Ferraro  used  Perry's  phrase  "the 
unsung  hero  of  space"  as  the  subtitle  for  his  1991  NASA  video  on  Scout. 

53.  Van  Cleve  interview  with  Ferraro,  16  Nov.  1990. 

54.  Hall  quoted  in  "Scout  Moves  to  Goddard,"  p.  7. 


485 


Notes  for  Chapter  8 

Chapter  8 
Enchanted  Rendezvous:  The  Lunar-Orbit  Rendezvous  Concept 

1.  President  John  F.  Kennedy,  quoted  in  John  M.  Logsdon,   The  Decision  to  Go  to  the 
Moon:  Project  Apollo  and  the  National  Interest  (Chicago:  University  of  Chicago  Press, 
1976),  p.  128. 

2.  Robert  R.  Gilruth,   "Experts  Were  Stunned  by  Scope  of  Mission,"   New  York  Times, 
Moon  Special  Supplement,  17  July  1969.   The  description  of  Gilruth's  reaction  is  from 
my  interview  with  him  in  Kilmarnock,  Va.,  10  July  1986,  and  from  Charles  Murray  and 
Catherine  Ely  Cox,  Apollo:  The  Race  to  the  Moon  (New  York:  Touchstone  Books,  1989), 
pp.  16-17. 

3.  Robert  R.  Gilruth  to  Francis  W.  Kemmett,  director  of  the  staff,  Inventions  and  Con- 
tributions Board,  NASA  headquarters,  28  Aug.  1973.    The  main  purpose  of  Gilruth's 
letter,  solicited  by  a  NASA  awards  board,  was  to  evaluate  Dr.  John  C.  Houbolt's  role  in 
NASA's  July  1962  decision  in  favor  of  LOR  for  Project  Apollo  and  to  determine  whether 
Houbolt's  contribution  was  worthy  of  the  maximum  prize  that  NASA  had  been  autho- 
rized to  give  ($100,000)  for  an  outstanding  national  contribution.   To  do  that,  however, 
Gilruth  had  to  review  the  STG's  position  on  LOR  and  the  entire  Apollo  mission-mode 
controversy.  I  do  not  believe  any  historian  besides  myself  has  seen  this  letter,  a  copy  of 
which  is  in  my  personal  LOR  file. 

4.  Clinton  E.  Brown  interview  with  the  author,  Hampton,  Va.,   17  July  1989.    Brown's 
remarks  are  from  a  panel  discussion  (involving  Brown,  William  H.  Michael,  Jr.,  and 
Arthur  W.  Vogeley),  which  I  organized  and  led  as  part  of  NASA  Langley's  celebration 
of  the  20th  anniversary  of  Apollo  11,  the  first  manned  lunar  landing.  A  videotape  of  the 
evening  program  featuring  this  panel  discussion  is  preserved  in  the  LHA. 

5.  Interview  with   William   H.   Michael,    Jr.,    Hampton,    Va.,    17  July   1989,    videotape, 
LHA.    F.  R.  Moulton's  book  on  celestial  mechanics  was  available  by  1958  in  a  second 
edition  (London:    Macmillan  Co.,  1956),  but  the  Langley  Technical  Library  appears  not 
to  have  had  it.  The  library  has  obtained  one  since. 

6.  On  the  history  of  pioneering  thought   about   and  proposals  for  space  stations,    see 
Frederick  I.  Ordway  III,   "The  History,  Evolution,  and  Benefits  of  the  Space  Station 
Concept,"  presented  to  the  13th  International  Congress  of  History  of  Science,  Aug.  1971; 
and  Barton  C.  Hacker,  "And  Rest  as  on  a  Natural  Station:  From  Space  Station  to  Orbital 
Operations  in  Space- Travel  Thought,  1885-1951."    Both  of  the  preceding  unpublished 
papers  are  available  in  the  HQA.    For  published  information,  see  Wernher  von  Braun 
and  Frederick  I.  Ordway  III,  Space  Travel:  A  History:  An  Update  of  'History  of  Rocketry 
&  Space  Travel'   (New  York:  Harper  &  Row,  1985),  pp.  18-20;  and  Howard  E.  McCurdy, 
The  Space  Station  Decision:  Incremental  Politics  and  Technological  Choice  (Baltimore 
and  London:     Johns  Hopkins  University  Press,    1991),   pp.   5-8  and  237  n.   7.      On 
NASA's  belief  that  a  space  station  was  the  logical  follow-on  to  Project  Mercury,  see 
Barton  C.  Hacker  and  James  M.  Grimwood,   On  the  Shoulders  of  Titans:    A  History 
of  Project  Gemini,  NASA  SP-4203  (Washington,  1977),  pp.  5r6;  and  McCurdy,  Space 
Station  Decision,  pp.  7-9. 

7.  Brown  interview,  17  July  1989. 


486 


Notes  for  Chapter  8 

8.  Clinton  E.  Brown  to  Eugene  C.  Draley,   "Formation  of  a  working  group  to  study  the 
problems  of  Lunar  exploration,"  24  Mar.  1959,  A200-1B,  LCF. 

9.  William  H.  Michael,  Jr.,  to  Eugene  C.  Draley,  "Attendance  at  meeting  of  Working  Group 
on  Lunar  and  Planetary  Surfaces  Exploration  at  NASA  Headquarters  on  February  14, 
1959,"   A200-1B,  LCF;  Michael,   "Attendance  at  meeting  for  Discussion  of  Advanced 
Phases  of  Lunar  Exploration  at  NASA  Offices,  Silver  Spring,  Md.,  Saturday,  May  2, 
1959,"  A200-1B,  LCF.  On  the  Jastrow  Committee,  see  R.  Cargill  Hall,  Lunar  Impact: 
A   History  of  Project  Ranger,   NASA   SP-4210   (Washington,    1977),   pp.    15-16;    and 
William  David  Compton,  Where  No  Man  Has  Gone  Before:  A  History  of  Apollo  Lunar 
Exploration  Missions,  NASA  SP-4214  (Washington,  1989),  pp.  13-14. 

10.  Harry  J.  Goett  to  Ira  H.  Abbott,  "Interim  Report  on  Operation  of  Research  Steering 
Committee  on  Manned  Space  Flight,"    17  July  1959,   A200-1B,  LCF.   On  the  Goett 
Committee,  see  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  pp.  9-10;  Logsdon, 
Decision  to  Go  to  the  Moon,  pp.  56-57;  and  Murray  and  Cox,  Apollo,  pp.  43-45. 

11.  Laurence  K.   Loftin,    Jr.,    interview   with  author,    Newport   News,    Va.,    5  Aug.  1989, 
transcript,  p.  100,  OHC,  LHA;  Murray  and  Cox,  Apollo,  pp.  43-45. 

12.  Michael  interview,  17  July  1989. 

13.  W.  H.  Michael,  Jr.,  "Weight  Advantages  of  Use  of  Parking  Orbit  for  Lunar  Soft  Landing 
Mission,"   in  Lunar  Trajectory  Group's  [Theoretical  Mechanics  Division]  unpublished 
"Studies  Related  to  Lunar  and  Planetary  Missions,"  26  May  1960,  A200-1B,  LCF. 

14.  On  13  Nov.  1948,  H.  E.  Ross  presented  the  essential  elements  of  LOR  to  a  meeting  of  the 
British  Interplanetary  Society  in  London.  His  conclusion  was  that  LOR — compared  with 
a  direct  flight  to  the  lunar  surface  from  the  earth — would  reduce  the  earth-launch  weight 
by  a  factor  of  2.6.    In  his  paper  Ross  credited  Hermann  Oberth,  Guido  von  Pirquet, 
Hermann  Noordung,  Walter  Hohmann  (of  "Hohmann  transfer"  fame),  Tsiolkovskii,  and 
F.  A.  Tsander  for  their  early  ideas  concerning  LOR.   See  H.  E.  Ross,  "Orbital  Bases,"  in 
Journal  of  the  British  Interplanetary  Society  26  (Jan.  1949):  1-18.  For  the  history  of  the 
first  pioneering  inklings  about  the  value  of  rendezvous  in  orbit,  lunar  and  otherwise,  see 
Barton  C.  Hacker,  "The  Idea  of  Rendezvous:  From  Space  Station  to  Orbital  Operations 
in  Space- Travel  Thought,  1895-1951,"   Technology  and  Culture  15  (July  1974):  373-388; 
as  well  as  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  pp.  5-6. 

15.  For  the  Vought  concept  that  came  out  of  the  MALLAR  study,  see  Vought  Astronautics 
brochure,  Manned,  Modular,  Multi-Purpose  Space  Vehicle,  Jan.  1960. 

16.  The  quote  from  Clint  Brown  about  Michael's  reaction  to  the  Vought  briefing  is  from 
Murray  and  Cox,  Apollo,  pp.  114-115. 

17.  Michael  interview  with  author,  Hampton,  Va.,  7  Apr.  1989,  transcript,  p.  6,  OHC,  LHA. 
The  story  about  Tom  Dolan  discussing  the  essentials  of  LOR  with  unknown  engineers  in 
PARD  is  taken  from  John  D.  Bird,  "A  Short  History  of  the  Development  of  the  Lunar- 
Orbit-Rendezvous  Plan  at  the  Langley  Research  Center,"  6  Sept.  1963  (supplemented 
5  Feb.  1965  and  17  Feb.  1966),  box  6,  Milton  Ames  Collection,  LHA. 

18.  Michael,  "Weight  Advantages  of  Use  of  Parking  Orbit  for  Lunar  Soft  Landing  Mission," 
p.  2. 

19.  The  "Jaybird"  story  is  taken  from  Murray  and  Cox,  Apollo,  p.  115.    For  the  technical 
reports  that  resulted  from  the  early  lunar  studies  in  the  Theoretical  Mechanics  Division, 
see  among  others:  J.  P.  Gapcynski,  "A  Consideration  of  Some  of  the  Factors  Involved  in 

487 


Notes  for  Chapter  8 

the  Departure  of  a  Vehicle  from  a  Circular  Orbit  About  the  Earth" ;  and  W.  L.  Mayo, 
"Energy  and  Mass  Requirements  for  Lunar  and  Martian  Missions."  Both  articles  are 
in  the  Lunar  Trajectory  Group's  "Studies  Related  to  Lunar  and  Planetary  Missions." 
On  Bird's  lunar  bug  ideas,  see  Michael  interview,  7  Apr.  1989,  pp.  14-15.  See  also 
William  H.  Michael,  Jr.,  and  Robert  H.  Tolson,  "Effect  of  Eccentricity  of  the  Lunar 
Orbit,  Oblateness  of  the  Earth,  and  Solar  Gravitational  Field  on  Lunar  Trajectories," 
June  1960,  copy  in  the  Langley  Technical  Library. 

20.  John  C.  Houbolt,  "A  Study  of  Several  Aerothermoelastic  Problems  of  Aircraft  Structures 
in  High-Speed  Flight,"  in  Eidgenossische  Technische  Hochshule  Mitteilung  5  (1958):  108. 
Throughout  his  career  at  NACA  and  NASA  Langley,  Houbolt  was  not  a  terribly  prolific 
author.  A  complete  bibliography  of  his  papers  is  available  in  the  LHA. 

21.  John  C.  Houbolt  interview  with  author,  Williamsburg,  Va.,  24  Aug.  1989,  transcript, 
p.  3,  OHC,  LHA. 

22.  Ibid.,  pp.  7-8. 

23.  Ibid.,  p.  9.  On  Rand  and  the  early  space  program,  see  Walter  A.  McDougall,  The  Heavens 
and  the  Earth:  A  Political  History  of  the  Space  Age  (New  York:  Basic  Books,  1985),  esp. 
pp.  89,  102,  106-110,  and  121-123. 

24.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  p.  13;  Loftin  interview,  5  Aug.  1989, 
pp.  100-101. 

25.  NASA   Langley,    "Minutes   of  Meeting   of  LRC   Manned   Space   Laboratory   Group," 
18  Sept.  1959,  A200-4,  LCF.   See  paragraph  5  for  Houbolt's  statement  on  the  rendezvous 
problem. 

26.  Houbolt  interview,  24  Aug.  1989,  pp.  9  10. 

27.  Bernard  Maggin  to  Milton  B.  Ames,  Jr.,  "Inter-center  discussions  of  space  rendezvous," 

23  May  1960;  John  C.  Houbolt,  "Considerations  of  the  Rendezvous  Problems  for  Space 
Vehicles,"  presented  at  the  National  Aeronautical  Meeting  of  the  Society  of  Automotive 
Engineers,  New  York  City,  5-8  Apr.  1960.   Both  documents  are  in  A200-1A,  LCF.  The 
point  about  Marshall's  limited  interest  in  the  rendezvous  problem  is  made  in  Hacker  and 
Grimwood,  On  the  Shoulders  of  Titans,  pp.  14-15. 

28.  Lowell    E.    Hasel,     "Minutes    of   Meeting   of   LRC    Lunar    Mission    Steering    Group," 

24  May  1960,  A200-1B,  LCF. 

29.  STG,  "Guidelines  for  Advanced  Manned  Space  Vehicle  Program,"  June  1960.    For  the 
summary  details  of  these  guidelines,  see  Ivan  D.  Ertel  and  Mary  Louise  Morse,    The 
Apollo  Spacecraft:  A  Chronology  (Washington:  NASA,  1969),  1:38-41.  In  brief,  the  STG 
identified  a  manned  circumlunar  mission  as  the  "logical  intermediate  step"  toward  future 
goals  of  lunar  and  planetary  landing.  Essential  to  the  guidelines  were  plans  for  advanced 
earth-orbital  missions  and  an  earth-orbiting  space  station. 

30.  STG,  "Apollo  Technical  Liaison  Plan,"  16  Nov.  1960,  A200-1B,  LCF;  Langley  to  STG, 
"Langley  appointments  to  Apollo  technical  liaison  groups,"  7  Dec.  1960,  in  Project  Apollo 
files,  LCF.    Interestingly,  John  Houbolt  was  not  appointed  to  any  of  the  liaison  groups. 

On  the  first  NASA/Industry  Apollo  Technical  Conference,  see  the  Wall  Street  Journal, 
18  July  1961.  Among  the  many  valuable  papers  given  by  Langley 's  John  Becker  to  the 
Archives  of  Aerospace  Exploration  at  Virginia  Polytechnic  Institute  and  State  University 
is  his  file  on  the  Lunar  Mission  Steering  Group.  On  the  cover  of  this  file,  Becker  provides 
a  brief  written  introduction  to  its  contents.  Inside  is  a  copy  of  the  Apollo  Technical 

488 


Notes  for  Chapter  8 

Liaison  Plan,  with  handwritten  notes  by  Becker.  For  information  about  this  file  and 
others  in  the  Archives  of  Aerospace  Exploration,  contact  the  special  collections  archivist 
at  the  university  library  in  Blacksburg,  Va. 

31.  Houbolt,  "Considerations  of  the  Rendezvous  Problems  for  Space  Vehicles,"  p.  1. 

32.  NASA   Langley,    "Minutes   of  Meeting  of  LRC   Manned   Space   Laboratory   Group," 
5  Feb.  1960,  A200-4,  LCF.    See  paragraphs  3  and  10. 

33.  William  A.  Mrazek,  Marshall  Space  Flight  Center,  to  John  C.  Houbolt,  16  May  1960, 
B10-6,  LCF;  Houbolt  interview,  24  Aug.  1989,  p.  11. 

34.  John    C.    Houbolt,    "Lunar    Rendezvous,"    International   Science   and    Technology   14 
(Feb.  1963):    63.    Several  attempts  have  been  made  to  clarify  the  detailed  history  of 
the  genesis  of  LOR  at  Langley  and  Houbolt's  role  in  it.    The  best  efforts  to  date  are: 
Hacker  and  Grimwood,    On  the  Shoulders  of  Titans,  pp.   14-16  and  60-68;   John  M. 
Logsdon,  "The  Choice  of  the  Lunar  Orbital  Rendezvous  Mode,"  in  Aerospace  Historian 
(June  1971):    63-70;  and  Murray  and  Cox,  Apollo,  chaps.  8  and  9.    However,  none  of 
these  are  complete,  and  none  fully  satisfy  the  Langley  participants  involved:   Houbolt, 
Michael,  Brown,  et  al.  In  a  footnote  to  their  excellent  book  on  Project  Apollo,  Murray 
and  Cox  remark  on  the  gaps  in  the  overall  LOR  story,  suggesting  that  "there  is  a  fas- 
cinating doctoral  dissertation  yet  to  be  written  on  this  episode."   This  chapter  may  not 
close  off  the  possibility  of  a  doctoral  dissertation,  but  my  goal  in  writing  it  was  to  fill  in 
many  of  the  gaps  and  stress  Langley's  role  in  LOR. 

35.  Houbolt  interview,  24  Aug.  1989,  p.  15;  Hacker  and  Grimwood,   On  the  Shoulders  of 
Titans,  pp.  15-16. 

36.  "RCA  Will  Do  Saint  Payload,"   Aviation   Week  and  Space  Technology  5  (Dec.   1960): 
27.  For  other  sources  on  Project  Saint,  see  Hacker  and  Grimwood,  On  the  Shoulders  of 
Titans,  p.  416  n.  64. 

37.  Houbolt  interview,  24  Aug.  1989,  pp.  16  and  21-25.     On  Houbolt's  final  chart  were 
three  conclusions:    (1)   "RENDEZVOUS  OPENS  POSSIBILITY  FOR  EARLIER  AC- 
COMPLISHMENT OF  CERTAIN  SPACE  MEASUREMENT  WITH  EXISTING  VE- 
HICLES"; (2)  "THERE  IS  A  NEED  FOR  RAPID  DEVELOPMENT  OF  MANNED 
RENDEZVOUS  TECHNIQUES— SHOULD  MAKE  USE  OF  MERCURY  AND  'SAINT' 
TECHNOLOGY";  and  (3)  "NASA  SHOULD  HAVE  MANNED  RENDEZVOUS  PRO- 
GRAM IN  LONG-RANGE  PLANS  WITH  OBJECTIVES  OF  EXPEDITING  SOFT 
LUNAR  LANDINGS  AND  FLEXIBLE  ORBITAL  OPERATIONS." 

38.  Murray  and  Cox,  Apollo,  p.  116. 

39.  Houbolt  interview,  24  Aug.  1989,  pp.  17-18. 

40.  See  the  reference  to  Brown's  presentation  in  Charles  J.  Donlan  to  NASA  headquarters, 
director  of  the  staff,  Inventions  and  Contributions  Board,  1  Mar.  1974,  author's  LOR 
file.  John  D.  Bird  also  refers  to  it  in  "A  Short  History  of  the  Development  of  the  Lunar- 
Orbit-Rendezvous  Plan,"  p.  2. 

41.  The  preceding  two  paragraphs  are  derived  from  my  unsigned  feature  story,  "The  Ren- 
dezvous That  Was  Almost  Missed:  Lunar-Orbit  Rendezvous  and  the  Apollo  Program," 
NASA  News  Release  No.  89-98,  7  July  1989,  pp.  5-6. 

42.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  pp.  28-29. 

43.  Ibid.,  and  Ertel  and  Morse,  The  Apollo  Spacecraft:  A  Chronology,  1:66. 

44.  Houbolt  interview,  24  Aug.  1989,  pp.  21-26. 

489 


Notes  for  Chapter  8 

45.  John  I.  Cumberland,  executive  secretary,  SEPC,  to  members  and  speakers,  "Agenda  for 
SEPC  Meeting,  Jan.  5-6,  1961,"  A200-1B,  LCF. 

46.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  p.  29. 

47.  Robert  L.  O'Neal  to  Charles  J.  Donlan,  associate  director,  "Discussion  with  Dr.  Houbolt, 
LRC,  concerning  the  possible  incorporation  of  a  lunar  orbital  rendezvous  phase  as  a 
prelude  to  manned  lunar  landing,"  30  Jan.  1961,  A200-1B,  LCF. 

48.  Ibid.,   p.  60;    Owen  Maynard  to  Frederick  J.   Lees,   chairman,   NASA  Inventions  and 
Contributions  Board,   13  Nov.   1982,  copy  in  author's  LOR  file.     In  truth,  Houbolt's 
numbers  were  overly  optimistic  in  estimating  the  required  weights  for  the  LEM,  because  in 
some  critical  areas  detailed  information  about  the  necessary  subsystems  was  not  available. 
Subsequent  analysis  by  NASA  and  its  industrial  contractors  provided  much  more  realistic 
weight  numbers.  Still,  the  later  values  for  these  weights  did  not  turn  out  to  be  so  radically 
different  than  Houbolt's  projections:  they  were  within  the  single-launch  capability  of  the 
Saturn  V  vehicle  and  therefore  validated  the  advertised  feasibility  of  the  LOR  mode  for 
the  manned  lunar  landing  mission. 

49.  Von  Braun  quoted  in  Murray  and  Cox,  Apollo,  pp.  116-117.  Quote  expressing  skepticism 
about   LOR  from   Dr.   Harvey   Hall  to  director   of  the  staff,    NASA   Inventions  and 
Contributions  Board,  28  Mar.  1973,  author's  LOR  file. 

50.  Murray  and  Cox,  Apollo,  p.  117. 

51.  Houbolt  interview,  24  Aug.  1989,  p.  22. 

52.  Gilruth  to  Kemmett,  28  Aug.  1973,  p.  1. 

53.  Donlan  to  director  of  the  staff,  NASA  Inventions  and  Contributions  Board,  1  Mar.  1974, 
p.  2. 

54.  Gilruth  to  Kemmett,  28  Aug.  1973,  p.  1. 

55.  Gilruth  to  Nicholas  E.  Golovin,  12  Sept.  1961,  quoted  in  Hacker  and  Grimwood,  On  the 
Shoulders  of  Titans,  p.  61. 

56.  The  quoted  phrases  are  from  Houbolt's  letters  to  Francis  W.  Kemmett,  NASA  Inventions 
and  Contributions  Board,  23  May  1978  and  2  Sept.  1981,  copies  in  author's  LOR  file. 

57.  George  Low,    "A  Plan  for  Manned  Lunar  Landing,"    24  Jan.   1961,   A200-1B,   LCF. 
On  the  cover  sheet  of  this  report,  Low  identifies  the  members  of  his  Lunar  Landing 
Working  Group;  Houbolt's  name  is  not  included.  George  M.  Low,  president,  Rensselaer 
Polytechnic  Institute,   Troy,   N.Y.,   to  Mr.   Frederick  J.   Lees,   NASA  Inventions  and 
Contributions  Board,  21  Oct.  1982,  copy  in  author's  LOR  file.   See  also  E.  O.  Pearson, 
"Notes  on  Key  Problems  of  Manned  Lunar  Missions,"  13  Jan.  1961,  A200-1B,  LCF.  The 
phrase  "extremely  far-out  thing  to  do"  is  George  Low's,  quoted  in  Murray  and  Cox, 
Apollo,  p.  117. 

58.  Low  to  Lees,  21  Oct.  1982,  p.  2. 

59.  See  M.  J.  Queijo  to  associate  director,  "Techniques  and  problems  associated  with  manned 
lunar  orbits  and  landings,"  21  Feb.  1961,  A200-1B,  LCF. 

60.  Bernard  Maggin  to  John  Houbolt,  1  Mar.  1961,  A200-1B,  LCF. 

61.  Brown  interview,  17  July  1989.  The  politics  also  involved  at  least  one  major  industrial 
firm,  North  American  Aviation  in  Los  Angeles,  which  already  had  a  contract  for  a 
command-and-service   module  based  on  the  direct-ascent   mode.      If  NASA  selected 
LOR,  North  American  most  probably  would  have  to  "share  the  pie"  with  some  other 
contractor  who  would  be  responsible  for  the  separate  lunar  lander.  The  other  contractor 

490 


turned  out  to  be  Grumman.  For  more  on  the  politics  of  the  mission-mode  decision, 
see  Henry  S.  F.  Cooper,  "We  Don't  Have  to  Prove  Ourselves"  in  The  New  Yorker 
(2  Sept.  1991):  64. 

62.  NASA  Langley,  "Work  at  LRC  in  Support  of  Project  Apollo,"  3  May  1961,  Project  Apollo 
file,  LCF.    See  also  Rufus  O.  House  to  the  Langley  director,  "Number  of  professionals  in 
support  of  Project  Apollo,"  19  May  1961,  Project  Apollo  file,  LCF.   This  memo  advised 
center  management  that  326.5  professionals  were  currently  involved  in  research  projects 
supporting  Apollo;  91  of  these  professionals  were  involved  in  the  study  of  reentry  heating 
problems. 

63.  Loftin  interview,  5  Aug.  1989,  p.  93. 

64.  NASA  Langley,  "Manned  Lunar  Landing  Via  Rendezvous,"  19  Apr.  1961,  A200-1B,  LCF. 

65.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  p.  61. 

66.  Houbolt  interview,  24  Aug.  1989,  p.  30. 

67.  Ibid.,  p.  28. 

68.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  p.  61. 

69.  Ibid. 

70.  See  McDougall,  Heavens  and  the  Earth,  pp.  8,  318,  and  328,  and  Murray  and  Cox,  Apollo, 
pp.  79-80. 

71.  For  an  analysis  of  Vice-President  Lyndon  Johnson's  enthusiasm  for  the  lunar  mission, 
see  McDougall,  Heavens  and  the  Earth,  pp.  319-320,  and  Logsdon,  Decision  to  Go  to  the 
Moon,  pp.  119-121. 

72.  For  a  more  complete  analysis  of  the  political  thinking  behind  Kennedy's  lunar  commit- 
ment, see  chap.  15  "Destination  Moon"  (pp.  307-324)  of  McDougall,  Heavens  and  the 
Earth. 

73.  John   C.   Houbolt   to   Dr.    Robert   C.   Seamans,    Jr.,    NASA   associate   administrator, 
9  May  1961,  copy  in  box  6,  Milton  Ames  Collection,  LHA. 

74.  Hacker  and  Grimwood,   On  the  Shoulders  of  Titans,  p.  36;  Murray  and  Cox,  Apollo, 
pp.  81-82  and  110.  The  Fleming  Committee  had  23  members,  including  18  from  NASA 
headquarters;    Langley  had  no  representative.     The  committee  members  from  head- 
quarters were  Fleming,   Addison  M.  Rothrock,   Albert  J.  Kelley,  Berg  Paraghamian, 
Walter    W.    Haase,    John    Disher,    Merle    G.    Waugh,    Eldon    Hall,    Melvyn    Savage, 
William  L.  Lovejoy,  Norman  Rafel,  Alfred  Nelson,  Samuel  Snyder,  Robert  D.  Briskman, 
Secrest  L.  Barry,  James  P.  Nolan,  Jr.,  Earnest  O.  Pearson,  and  Robert  Fellows.    The 
other  members  were  Heinz  H.  Koelle  (Marshall),  Kenneth  S.  Kleinknecht  and  Alan  Kehlet 
(STG),  A.  H.  Schichtenberg  (The  Lovelace  Foundation),  and  William  S.  Shipley  (JPL). 
Not  surprisingly,  most  of  these  men  were  big-rocket  specialists. 

75.  Seamans   to   director,    Launch   Vehicle   Programs   (Don   R.   Ostrander)    and   director, 
Advanced  Research  Programs   (Ira  H.   Abbott),    "Broad  Study  of  Feasible  Ways  for 
Accomplishing  Manned  Lunar  Landing  Mission,"  25  May  1961,  A200-1B,  LCF. 

76.  Seamans,  "Broad  Study."  Seamans,  associate  administrator,  to  John  C.  Houbolt,  NASA 
Langley,  2  June  1961,  copy  in  box  6,  Milton  Ames  Collection,  LHA. 

77.  Murray  and  Cox  state  that  Houbolt  was  a  member  of  the  Lundin  Committee  (Apollo, 
p.  118).    Loftin  interview,  5  Aug.  1989,  pp.  91-97;  Houbolt  interview,  24  Aug.  1989, 
pp.  32-34.    In  reading  a  preliminary  draft  of  this  chapter,  Loftin  challenged  Houbolt 's 
statement.  Loftin  insists  that  he  spent  weeks  in  Washington  on  this  committee  and  only 

491 


Notes  for  Chapter  8 

sent  Houbolt  in  his  place  the  few  times  he  could  not  attend  a  meeting.  Other  members 
of  the  Lundin  Committee,  besides  Lundin  and  Loftin,  were  Walter  J.  Downhower  (JPL), 
Alfred  E.  Eggers  (Ames),  Harry  O.  Ruppe  (Marshall),  and  Lt.  Col.  George  W.  S. 
Johnson  (U.S.  Air  Force).  Unlike  the  Fleming  Committee,  this  task  force — by  design — 
had  no  members  from  NASA  headquarters  and  was  conceived  to  represent  the  technical 
judgments  of  the  NASA  centers. 

78.  Laurence   K.   Loftin,    Jr.,    told  this  story  from  the   audience   during  the  videotaped 
celebration  program  for  the  20th  anniversary  of  the  Apollo  11  lunar  landing.    See  also 
Loftin  interview,  5  Aug.  1989,  p.  93. 

79.  Houbolt  interview,  24  Aug.  1989,  pp.  33-34.    NASA  (Lundin  Committee),   "A  Survey 
of  Various  Vehicle  Systems  for  the  Manned  Lunar  Landing  Mission,"    10  June  1961, 
A200-1B,  LCF. 

80.  Ibid.,  p.  16;  Houbolt  interview,  24  Aug.  1989,  p.  34. 

81.  Houbolt  interview,  24  Aug.  1989,  p.  33;  see  also  Murray  and  Cox,  Apollo,  p.  118. 

82.  Loftin  interview,  5  Aug.  1989,  pp.  93-94. 

83.  Seamans  to  directors  for  Launch  Vehicle  Programs,  Advanced  Research  Programs,  and 
to  acting  director  for  Life  Sciences  Program,  "Establishment  of  Ad  Hoc  Task  Group  or 
Manned  Lunar  Landing  by  Rendezvous  Techniques,"  20  June  1961,  A200-1B,  LCF.    See 
also  Hacker  and  Grimwood,    On  the  Shoulders  of  Titans,  pp.  37-38.     Serving  on  the 
Heaton  Committee  were  10  officials  from  NASA  headquarters,  5  from  Marshall,  1  from 
the  Flight  Research  Center,  and  2  from  Langley,  Houbolt  and  W.  Hewitt  Phillips.  One 
representative  from  the  U.S.  Air  Force  was  also  a  member. 

84.  Houbolt's  paper  "Problems  and  Potentialities  of  Space  Rendezvous,"  first  presented  at 
the  International  Academy  of  Astronautics'  International  Symposium  on  Space  Flight 
and  Reentry  Trajectories,  was  published  under  the  same  title  in  Astronautica  Acta  7 
(1961):  406-429. 

85.  Houbolt  interview,  29  Aug.  1989,  p.  39. 

86.  For  the  conclusions  of  the  Heaton  Committee,  see  Ad  Hoc  Task  Group  for  Study  of 
Manned  Lunar  Landing  by  Rendezvous  Techniques,  "Earth  Orbital  Rendezvous  for  an 
Early  Manned  Lunar  Landing,"  pt.  1,  "Summary  Report  of  Ad  Hoc  Task  Group  Study," 
Aug.  1961.  Houbolt  interview,  29  Aug.  1989,  p.  39. 

87.  Murray  and  Cox,  Apollo,  p.  119.  See  also  Houbolt  interview,  24  Aug.  1989,  p.  46. 

88.  Houbolt  interview,  24  Aug.  1989,  pp.  47-48. 

89.  Ibid. 

90.  Ibid. 

91.  "Report  of  DOD-NASA  Large  Launch  Vehicle  Planning  Group,  Vol.  1,  1961,"  A200-1B, 
LCF.  See  Hacker  and  Grimwood,   On  the  Shoulders  of  Titans,  pp.  67-68.    The  mem- 
bers of  this  committee,  besides  those  mentioned  in  the  text,  were:    Kurt  R.  Stehling 
and  William  A.  Wolman  (NASA  headquarters);    Warren  H.  Amster  and  Edward  J. 
Barlow  (Aerospace  Corp.);  Seymour  C.  Himmel  (Lewis);  Wilson  Schramm  and  Francis  L. 
Williams  (Marshall);  Col.  Matthew  R.- Collins  (army);  Rear  Adm.  Levering  Smith  and 
Capt.  Lewis  J.  Stecher,  Jr.  (navy);  and  Col.  Otto  J.  Glaser,  Lt.  Col.  David  L.  Carter, 
and  Heinrich  J.  Weigand  (air  force).  No  Langley  representatives  were  on  the  committee. 

92.  Bird,  "A  Short  History  of  the  Development  of  the  Lunar-Orbit-Rendezvous  Plan,"  final 
supplement,  17  Feb.  1966,  p.  3. 

492 


Notes  for  Chapter  8 

93.  Harvey  Hall,  NASA  coordinator,  NASA/DOD  Large  Launch  Vehicle  Planning  Group,  to 
Langley,  23  Aug.  1961,  A200-1B,  LCF. 

94.  Houbolt  interview,  24  Aug.  1989,  p.  49. 

95.  Hacker  and  Grimwood,  On  the  Shoulders  of  Titans,  pp.  55-60. 

96.  On  the  Shoulders  of  Titans  is  an  outstanding,  detailed  history  of  the  Gemini  Program. 
For  a  more  brief  and  somewhat  more  colorful  insight  into  this  important  preparatory 
program  for  the  Apollo  mission,  see  Michael  Collins,  Liftoff:    The  Story  of  America's 
Adventure  in  Space  (New  York:  NASA/Grove  Press,  1988),  pp.  63-113.  See  also  Collins' 
memoir  Carrying  the  Fire:  An  Astronaut's  Journeys  (New  York:   Macmillan  Co.,  1977). 
Collins,  the  CM  pilot  for  the  Apollo  11  mission  to  the  moon,  was  also  an  astronaut  in  the 
Gemini  program  (Gemini-Titan  X,  18-21  July  1966).  His  memoir  is  one  of  the  best  books 
about  the  manned  space  program  of  the  1960s.   On  Gemini's  relationship  to  MORAD, 
see  Houbolt  interview,  24  Aug.  1989,  pp.  49-50. 

97.  NASA  Langley,  Manned  Lunar- Landing  through  Use  of  Lunar- Orbit  Rendezvous,  2  vols., 
31  Oct.  1961,  copy  in  box  6,  Milton  Ames  Collection,  LHA.  Other  Langley  researchers 
who  made  contributions  to  this  two-volume  report  were  Jack  Dodgen,  William  Mace, 
Ralph  W.  Stone,  Jack  Queijo,  Bill  Michael,  Max  Kurbjun,  and  Ralph  Brissenden.    In 
essence,  Houbolt  and  associates  prepared  this  report  as  a  working  paper  that  could 
provide,  as  NASA  Deputy  Administrator  Hugh  L.  Dryden  would  later  explain,  "a  quick 
summary  of  the  information  then  available  on  LOR  as  a  mode  of  accomplishing  manned 
lunar  landing  and  return."  See  Dryden  to  the  Honorable  Clinton  P.  Anderson,  chairman, 
Committee  on  Aeronautical  and  Space  Sciences,  U.S.  Senate,  11  Apr.  1963,  A200-1B, 
LCF. 

98.  John  C.  Houbolt  to  Robert  C.  Seamans,  associate  administrator,  NASA,  15  Nov.  1961, 
p.  1,  copy  in  box  7,  Milton  Ames  Collection,  LHA. 

99.  Ibid.,  p.  1. 

100.  Robert  C.  Seamans,  Jr.,  to  Dr.  John  C.  Houbolt,  4  Dec.  1961,  box  6,  Milton  Ames 
Collection,  LHA. 

101.  George  M.  Low,  president,  Rensselaer  Polytechnic  Institute,  Troy,  N.Y.,  to  Mr.  Frederick 
J.  Lees,  chairman,  Inventions  and  Contributions  Board,  NASA,  21  Oct.  1982,  copy  in 
author's  LOR  file.  See  also  Murray  and  Cox,  Apollo,  p.  120. 

102.  For  an  excellent  capsule  portrait  of  Dr.  Joseph  F.  Shea,  see  Murray  and  Cox,  Apollo, 
pp.  120-125. 

103.  Ibid.,  p.  124. 

104.  Ibid. 

105.  Ibid.,  p.  125. 

106.  On  Glenn's  historic  flight,  see  Loyd  S.  Swenson,  James  H.  Grimwood,  and  Charles  C. 
Alexander,  This  New  Ocean:  A  History  of  Project  Mercury,  NASA  SP-4201  (Washington, 
1966),  pp.  420-436.  For  more  on  the  Chance  Vought  contract,  see  Ertel  and  Morse,  Apollo 
Spacecraft:   A  Chronology,  1:141.    On  the  Chance  Vought  team  briefing,  see  Bird,   "A 
Short  History  of  the  Development  of  the  Lunar-Orbit-Rendezvous  Plan,"  supplemented 
17  Feb.  1966,  p.  4.  Regarding  the  headquarters  meeting,  see  NASA,  "Minutes  of  Lunar 
Orbit  Rendezvous  Meeting,  April  2-3,  1962,"  copy  in  A200-1B,  LCF.   See  also  Ertel  and 
Morse,  Apollo  Spacecraft:  A  Chronology,  1:147-152. 

107.  Arthur  W.  Vogeley  interview  with  author,  Hampton,  Va.,  17  July  1989,  videotape,  LHA. 

493 


Notes  for  Chapter  8 

108.  Faget  quoted  in  Henry  S.  F.  Cooper,  "We  Don't  Have  to  Prove  Ourselves,"  p.  64. 

109.  Ibid.  For  details  of  the  Manned  Spacecraft  Center's  final  evaluation  in  favor  of  LOR,  see 
Charles  W.  Matthews,  chief,  Spacecraft  Research  Division,  MSC,  to  Robert  R.  Gilruth, 
MSC  director,  "Summary  of  MSC  Evaluation  of  Methods  for  Accomplishing  the  Manned 
Lunar  Landing  Mission,"  2  July  1962;  and  Robert  R.  Gilruth  to  NASA  headquarters 
("Attn:  Mr.  D.  Brainerd  Holmes"),  "Summary  of  Manned  Spacecraft  Center  evaluation 
methods  for  accomplishing  the  manned  lunar  landing  mission,"  5  July  1962.  Both  mem- 
oranda are  included  in  app.  A  of  the  NASA  Office  of  Manned  Space  Flight's  confidential 
169-page  report,  Manned  Lunar  Landing  Program  Mode  Comparison,  30  July  1962. 

110.  Gilruth  to  Kemmett,  28  Aug.  1973. 

111.  See  Murray  and  Cox,  Apollo,  p.  125. 

112.  Ibid. 

113.  Low  to  Lees,  21  Oct.  1982;  Low  to  John  C.  Houbolt,  Senior  Vice-President  and  Senior 
Consultant,   Aeronautical  Research  Associates  of  Princeton,    7  Aug.   1969,   copies  in 
author's  LOR  file. 

114.  Axel  T.  Mattson,  research  assistant  for  Manned  Spacecraft  Projects,  to  Charles  J.  Donlan, 
"Report  on  activities  (April  16-April  19,  1962)  regarding  Manned  Spacecraft  Projects," 
A189-5,   LCF.  Mattson's  memo  lists  all  personnel  that  Houbolt  saw  at  the  Manned 
Spacecraft  Center.    One  of  the  Houston  engineers  with  whom  Houbolt  and  Bird  met, 
Chuck  Matthews,  had  just  returned  from  a  meeting  at  NASA  Marshall.  There  Matthews 
had  reviewed  Houston's  thinking  on  the  LOR  concept.   According  to  Mattson's  memo, 
that  presentation  was  "apparently  well  received  by  von  Braun,  since  he  made  favorable 
comments."     See  also  Mattson  interview  with  author,  Hampton,  Va.,   14  Aug.   1989, 
transcript,  LHA. 

115.  Statement  by  Axel  T.   Mattson  at  the   17  July   1989  evening  program  on  the  20th 
anniversary  of  Apollo  11,  videotape,  LHA. 

116.  "Minutes  of  the  MSF  Management  Council,"  6  Feb.  1962,  p.  1,  A200-1B,  LCF;  Houbolt 
interview,  24  Aug.  1989,  p.  64. 

117.  Brown  interview,  17  July  1989;  Houbolt  interview,  24  Aug.  1989,  pp.  68-72. 

118.  "Concluding  Remarks  by  Dr.  Wernher  von  Braun  about  Mode  Selection  for  the  Lunar 
Landing  Program  Given  to  Dr.  Joseph  F.  Shea,  Deputy  Director  (Systems)  Office  of 
Manned  Space  Flight,"  7  June  1962,  copy  in  box  6,  Milton  Ames  Collection,  LHA.  For 
more  on  von  Braun's  surprise  announcement  in  favor  of  LOR  and  the  reaction  of  the 
Marshall  audience  to  it,  see  Murray  and  Cox,  Apollo,  p.  139. 

119.  For  an  introduction  to  the  concept  of  "closure"  in  science  and  technology,  see  The  Social 
Construction  of  Technological  Systems:  New  Directions  in  the  Sociology  and  History  of 
Technology,  eds.,  Wiebe  E.  Bijker,  Thomas  P.  Hughes,  and  Trevor  Pinch  (Cambridge  and 
London:  MIT  Press,  1990),  pp.  12-13. 

120.  John  C.  Houbolt  to  Dr.  Wernher  von  Braun,  director,  NASA  Marshall,  9  Apr.  1962, 
A189-7,  LCF;  von  Braun  to  Houbolt,  20  June  1962,  copy  in  author's  LOR  file.    Von 
Braun  argued  later  that  he  really  had  not  changed  his  mind  from  EOR  to  LOR;  he  had 
not  been  that  strong  a  supporter  of  EOR  in  the  first  place.  His. people  at  Marshall  had 
investigated  EOR,  and  Gilruth's  people  in  Houston  had  investigated  LOR,  but  that  was 
part  of  a  NASA  management  strategy  to  cover  all  the  options  thoroughly.  He  claims  he 


494 


Notes  for  Chapter  8 

personally  did  not  take  sides  until  he  had  all  the  facts.  When  he  did,  he  supported  LOR. 
See  Murray  and  Cox,  Apollo,  p.  139. 

121.  NASA,  "Lunar  Orbit  Rendezvous:   News  Conference  on  Apollo  Plans  at  NASA  Head- 
quarters on  July  11,  1962,"  copy,  box  6,  Milton  Ames  Collection,  LHA.  In  the  press 
conference,  Dr.  Robert  Seamans  credited  John  Houbolt  specifically  for  his  contribution 
to  the  LOR  concept:  "I  would  first  like  to  say  that  when  I  joined  NASA  almost  two  years 
ago  one  of  the  first  places  that  I  went  to  was  Langley  Field,  and  there  reviewed  work  going 
on  on  a  research  basis  under  Dr.  John  Houbolt.   This  work  related  both  to  rendezvous 
and  what  a  man  could  do  at  the  controls,  of  course  under  simulated  conditions,  as  well 
as  the  possibility  of  lunar  orbit  rendezvous"  (p.  8).  On  Wiesner's  opposition  to  LOR,  see 
McDougall,  Heavens  and  the  Earth,  p.  378,  and  Murray  and  Cox,  Apollo,  pp.  140-143. 

Even  after  its  July  1962  announcement  in  favor  of  LOR,  NASA  continued  to  evaluate 
the  other  major  options  for  the  Apollo  mission  mode.  See,  for  example,  the  Office  of 
Manned  Space  Flight's  confidential  Manned  Lunar  Landing  Program  Mode  Comparison, 
30  July  1962,  and  the  follow-up,  Manned  Lunar  Landing  Mode  Comparison,  24  Oct.  1962. 
Copies  of  both  documents  are  in  A200-1B,  LCF.  Both  reports  concluded  that  on  the 
basis  of  "technical  simplicity,  scheduling,  and  cost  considerations,"  LOR  was  the  "most 
suitable"  and  the  "preferred  mode."  Some  forms  of  EOR  were  also  feasible  and  would 
have  adequate  weight  margins. 

122.  John  C.  Houbolt,  Roy  Steiner,  and  Kermit  G.  Pratt,  "Flight  Data  and  Considerations  of 
the  Dynamic  Response  of  Airplanes  to  Atmospheric  Turbulence,"  July  1962. 

123.  Houbolt  interview,  24  Aug.  1989,  p.  73. 

124.  I.  E.  Garrick,   Distinguished  Research  Associate,   NASA  Langley,  to  Mr.  Francis  W. 
Kemmett,   Inventions  and  Contributions  Board,   NASA  headquarters,    18  Nov.   1974. 
Garrick  wrote  two  other  letters  to  Kemmett,  dated  14  Nov.  1975  and  12  Sept.  1978. 
Copies  of  all  three  are  in  author's  LOR  file. 

125.  Low  to  Lees,  21  Oct.  1982,  pp.  2-3. 

126.  On  7  Aug.  1969,  two  weeks  after  the  successful  completion  of  the  Apollo  11  mission, 
von  Braun  wrote  Houbolt  a  personal  letter  in  which  he  referred  to  Houbolt 's  "singular 
contribution  to  the  Apollo  program."     "We  know  that  it  must  be  highly  gratifying  to 
you  because  of  the  rousing  and  complete  success  of  your  Eagle.    The  LM  concept  that 
you  developed  and  defended  so  effectively — even,  on  occasion,  before  unsympathetic 
tribunals — was  indeed  a  prime  factor  in  the  success  of  man's  first  lunar  landing  mission." 
Wernher  von  Braun,  director,  NASA  Marshall,  to  John  C.  Houbolt,  senior  vice-president 
and  senior  consultant,  Aeronautical  Research  Associates  of  Princeton,  Princeton,  N.J., 
7  Aug.  1969,  copy  in  author's  LOR  file. 

Throughout  this  chapter,  I  have  often  called  Houbolt  "a  crusader,"  knowing  all  too 
well  that  this  association  has  plagued  Houbolt  for  nearly  30  years.  This  characterization 
of  Houbolt  is  one  of  the  major  factors  that  killed  his  chances  for  receiving  a  $100,000  cash 
award  from  a  NASA  inventions  and  contributions  board  in  the  late  1970s  and  early  1980s. 
This  board  decided  after  a  lengthy  inquiry  that  it  did  not  give  awards  to  individuals  who 
simply  advocated  or  "crusaded"  for  causes — however  righteous. 

The  NASA  awards  board  perhaps  significantly  underestimated  the  sometimes  vital  role 
of  a  crusader  in  the  ultimate  success  of  a  major  technological  endeavor.  Most  certainly 
in  this  case  the  awards  board  worked  from  far  too  literal  a  definition  of  "crusader," 

495 


Notes  for  Chapter  9 

because  Houbolt  was  not  just  arguing  for  something  for  which  other  people  were  more 
responsible.  He  made  LOR  a  personal  cause,  only  after  his  own  extensive  work  on  the 
relevant  problems.  "Not  until  I  showed  them  all  my  analysis  and  so  forth  did  the  awards 
committee  even  realize  that  I  had  gone  into  so  much  depth  in  terms  of  working  through 
all  the  various  parts  of  the  problem,"  says  Houbolt. 

Chapter  9 
Skipping  "The  Next  Logical  Step" 

Much  of  the  information  expanded  upon  in  this  chapter  is  based  on  an  earlier 
unpublished  essay,  "Visions  of  Man  in  Space:  NASA  Langley  Research  Center  and  the 
United  States  Space  Station  Program,"  by  Beverly  C.  McMillan  and  James  R.  Hansen, 
final  revision,  Sept.  1992.  I  would  like  to  thank  Ms.  McMillan  for  her  valuable  assistance 
in  clarifying  the  early  history  of  space  station  research  at  Langley. 

1.  Von  Puttkamer  quoted  in  Charles  R.  Pelligrino  and  Joshua  Stoff,   Chariots  for  Apollo 
(N.Y.:  Atheneum,  1986),  pp.  24-25. 

2.  For  an  analysis  of  the  long  and  twisting  political  history  of  NASA's  attempt  to  de- 
velop and  operate  an  orbiting  space  station,  see  Howard  E.  McCurdy,  The  Space  Station 
Decision:  Incremental  Politics  and  Technological  Choice  (Baltimore  &  London:   Johns 
Hopkins  University  Press,  1990),  and  Sylvia  D.  Fries,  "2001  to  1994:  Political  Environ- 
ment and  the  Design  of  NASA's  Space  Station  System,"  Technology  and  Culture  29  (July 
1988):  568-593.   On  the  operational  history  and  demise  of  Skylab,  see  Michael  Collins, 
Liftoff:    The  Story  of  America's  Adventure  in  Space  (New  York:    NASA/Grove  Press, 
1988),  pp.  163-200,  as  well  as  James  E.  Oberg's  insightful  essay,   "Skylab's  Untimely 
End,"  in  Air-&  Space /Smithsonian  6  (Feb./Mar.  1992):  73-79.  My  assertion  that  space 
station  designers  deliberately  designed  Skylab  for  only  a  short  stay  in  space  so  that  they 
could  replace  it  with  a  more  elaborate  station  is  based  on  personal  interviews  with  key 
NASA  officials  involved  in  space  station  work;  they  wish  to  remain  anonymous. 

3.  K.  E.  Tsiolkovskii,  Investigation  of  Universal  Space  by  Means  of  Reactive  Devices  (1911 
and  1926),  NASA  translations  in  Tsiolkovskii,  Works  on  Rocket  Technology,  Nov.  1965, 
HQA;  Hermann  Oberth,  Die  Rakete  zu  den  Planetenraumen  (1923),  NASA  translation 
in  HQA;  and  Guido  von  Pirquet,  "Fahrtrouten,"  in  Die  Rakete  (May  1923  to  Apr.  1929). 
For  a  summary  of  the  pioneering  ideas  on  space  stations,  see  McCurdy,  Space  Station 
Decision,  pp.  5-8.  On  Tsiolkovskii  specifically,  see  I.  A.  Kol'chenko  and  I.  V.  Strazheva, 
"The  Ideas  of  K.  E.  Tsiolkovskii  on  Orbital  Space  Stations,"  in  History  of  Rocketry  and 
Astronautics,  ed.  R.  Cargill  Hall  (San  Diego:    American  Astronautical  Society,  1986), 
pp.  170-176;  on  Pirquet,  see  Fritz  Sykora,   "Guido  von  Pirquet:    Austrian  Pioneer  of 
Astronautics,"  in  History  of  Rocketry  and  Astronautics,  pp.  140-156. 

For  a  summary  and  analysis  of  early  space  station  proposals,  see  the  following 
essays:  Frederick  I.  Ordway  III,  "The  History,  Evolution,  and  Benefits  of  the  Space 
Station  Concept,"  presented  to  the  13th  International  Congress  of  the  History  of 
Science,  Aug.  1971,  copy  in  HQA;  Barton  C.  Hacker,  "And  Rest  as  in  a  Natural 
Station:  From  Space  Station  to  Orbital  Operations  in  Space- Travel  Thought,  1885-1951," 
unpublished  manuscript,  [undated],  HQA;  Alex  Roland,  "The  Evolution  of  Civil  Space 


496 


Notes  for  Chapter  9 

Station  Concepts  in  the  United  States,"  unpublished  manuscript,  May  1983,  HQA; 
and  John  M.  Logsdon,  "The  Evolution  of  Civilian  In-Space  Infrastructure,  i.e.,  'Space 
Station'  Concepts  in  the  United  States,"  in  Civilian  Space  Stations  and  the  U.S.  Future 
in  Space  (Washington:  U.S.  Congress,  Office  of  Technology  Assessment,  OTA-STI-241, 
Nov.  1984). 

4.  Wernher  von  Braun,  "Crossing  the  Last  Frontier,"   Collier's  129  (22  Mar.  1952):  24-29 
and  72  74.  For  the  other  articles  in  this  special  issue  of  Collier's  devoted  to  the  theme 
of  space  exploration,  see  "Man  Will  Conquer  Space  Soon." 

5.  Von  Braun,  "Crossing  the  Last  Frontier,"  pp.  24-25. 

6.  Ibid.    For  additional  information  and  thoughts  on  von  Braun's  space  station  concept, 
see  McCurdy,   Space  Station  Decision,  pp.  5-7,  and  Christopher  Lampton,    Wernher 
von  Braun  (New  York:  Franklin  Watts,  1988),  p.  106. 

7.  [Laurence  K.  Loftin,  Jr.],   "Manned  Space  Flight  Proposal"   [Summer  1959];  Harry  J. 
Goett  to  Research  Steering  Committee  on  Manned  Space  Flight,  "Plans  for  May  25-26 
Meeting,"    6  May  1959.     Both  documents  can  be  found  in  the  Floyd  L.  Thompson 
Collection,  LHA.  See  also  the  Minutes  of  the  Research  Steering  Committee  on  Manned 
Space  Flight,  25-26  May  1959,  copy  in  HQA. 

8.  Laurence  K.  Loftin,  "Manned  Space  Flight  Proposal,"  pp.  2-4.   For  a  brief  analysis  of 
the  X-20  Dyna-Soar  program,  see  Walter  A.  McDougall,   The  Heavens  and  the  Earth: 
A  Political  History  of  the  Space  Age  (New  York:   Basic  Books,  1985),  pp.  190-191  and 
339-341. 

9.  Loftin  interview  with  author,  Newport  News,  Va.,  5  Aug.  1989. 

10.  [Laurence  K.  Loftin,  Jr.],  untitled  talk  at  the  1959  NASA  inspection  [Fall  1959],  AMIS 
Committee  file,  Thompson  Collection,  LHA. 

11.  Ibid.;  Loftin  interview,  5  Aug.  1989.    In  his  inspection  presentation,  Loftin  expressed 
many  of  the  same  ideas  that  he  had  offered  at  the  second  meeting  of  the  Research  Steering 
Committee  on  Manned  Space  Flight,  held  at  Ames  Research  Center,  on  25-26  May  1959. 
A  copy  of  the  minutes  of  this  meeting  is  preserved  in  the  HQA. 

12.  Mark  R.  Nichols  interview  with  author,  Hampton,  Va.,  6  July  1988,  transcript,  p.  3, 
OHC,  LHA. 

13.  [Laurence  K.  Loftin,  Jr.],  "Manned  Space  Laboratory  Research  Group,"  [Fall  1959],  an 
organization  chart  with  handwritten  notes  on  attached  page  by  Floyd  Thompson,  AMIS 
Committee  file,  Thompson  Collection,  LHA. 

14.  Ibid. 

15.  Loftin  interview,  5  Aug.  1989.  See  also  McCurdy,  Space  Station  Decision,  pp.  8-9. 

Industry's  early  interest  in  the  space  station  can  be  seen  in  several  documents  in  file 
A200-1A,  LCF.  For  example,  see  J.  O.  Charshafian,  general  manager,  Santa  Barbara 
Division,  Curtiss- Wright  Corporation,  to  E.  C.  Draley,  assistant  director,  NASA  Langley, 
"NASA  Assistance  on  Aerial  Platform,"  19  May  1959;  Beverly  Z.  Henry,  Jr.,  to  Associate 
Director  [Thompson],  "Visit  of  Goodyear  Aircraft  Corporation  representatives  to  the 
Full-Scale  Research  Division,"  19  Oct.  1959;  Ralph  W.  Stone,  Jr.,  to  Associate  Director 
[Thompson],  "Visits  of  Messrs.  A.  B.  Thompson  and  Paul  Petty  of  Chance  Vought 
Astronautics  to  Langley,  November  6,  1959,"  12  Nov.  1959;  Beverly  Z.  Henry,  Jr., 
to  Associate  Director  [Thompson],  "Visit  of  the  Martin  Company  representatives  on 
January  12,  1960,"  26  Jan.  1960. 

497 


Notes  for  Chapter  9 

16.  [Loftin],  untitled  talk  at  the  1959  NASA  inspection,  p.  1;  Loftin  interview,  5  Aug.  1989. 

17.  Emanuel  Schnitzer,  "Erectable  Torus  Manned  Space  Laboratory,"  16  May  1960,  A200-4, 
LCF.  See  also  Schnitzer  to  Associate  Director  [Thompson],   "Space  Station  Project — 
Results  of  discussions  at  NASA  Headquarters  on  May  10-11,  1960,"    23  May   1960, 
A200-4,  LCF. 

18.  Schnitzer  to  Associate  Director  [Thompson],  "Transmittal  of  updated  article  on  Erectable 
Torus  Manned  Space  Laboratory  to  the  American  Rocket  Society,"  24  Oct.  1960,  A200-4, 
LCF.  Several  documents  from  1960  and  1961  pertaining  to  the  work  on  the  inflatable 
torus  are  in  A200-4,  LCF.    See  also  W.  Ray  Hook,  "Space  Stations — Historical  Review 
and  Current  Plans,"  unpublished  paper  presented  at  the  winter  annual  meeting  of  the 
American  Society  of  Mechanical  Engineers,  Phoenix,  14-19  Nov.  1982,  Langley  Technical 
Library.  Hook  is  a  Langley  engineer  who  has  been  involved  with  space  station  work  since 
the  1960s. 

19.  On  the  wobble  problem,  see  Peter  R.  Kurzhals,  James  J.  Adams,  and  Ward  F.  Hodge, 
"Space  Station  Dynamics  and  Control,"  in  NASA  Langley,  A  Report  on  the  Research  and 
Technological  Problems  of  a  Manned  Rotating  Spacecraft,  NASA  TN  D-1504  (Washington, 
Aug.  1962),  pp.  71-72. 

20.  Floyd  L.  Thompson  to  NASA,   "100-foot  diameter  automatic  erecting  modular  space 
station — Request  for  approval  of  study  contract,"  9  June  1961,  A200-4,  LCF.   For  design 
information,  see  Rene  Berglund,   "Space  Station  Research  Configurations,"   in  NASA 
TN  D-1504,  pp.  9-21.  For  an  analytical  summary  of  Langley's  designs  of  the  inflatable 
torus  and  rotating  hexagon,  see  Fries,  "2001  to  1994,"  pp.  572-578.  The  opening  paper 
of  TN  D-1504,   "Space  Station  Objectives  and  Research  Guidelines"   by  Paul  R.  Hill 
and  Emanuel  Schnitzer  (pp.  1-9),  offers  a  general  statement  of  early  Langley  goals  and 
ambitions  regarding  its  space  station  work  with  both  the  inflatable  torus  and  rotating 
hexagon  concepts.  For  North  American's  presentation  of  the  hexagon  concept,  see  North 
American  Aviation,  "Self-Deploying  Space  Station,  Final  Report,"  SID  62-1302,  31  Oct. 
1962,  Langley  Technical  Library. 

21.  Emanuel  Schnitzer,  memorandum  for  files,   "Results  of  Space  Station  Conference  be- 
tween NASA  Headquarters  and  Goodyear  Aircraft  Corporation,"  21  Apr.  1961,  A200-4, 
LCF.     For  insights  into  the  turbulent  history  of  the  Centaur  development,  see  Roger 
Bilstein,  Stages  to  Saturn:  A  Technological  History  of  the  Apollo/Saturn  Launch  Vehicles, 
NASA  SP-4206  (Washington,  1980),  esp.  pp.  36-38,  43-46,  and  131-137;  and  Virginia  P. 
Dawson,  Engines  and  Innovation:  Lewis  Laboratory  and  American  Propulsion  Technol- 
ogy, NASA  SP-4306  (Washington,  1991),  pp.  166-169,  179,  and  188-193.  On  the  matter 
of  maneuvering  a  rendezvous  with  a  rotating  space  station,  see  Roy  F.  Brissenden,  "Some 
Considerations  of  the  Operational  Requirements  for  a  Manned  Space  Station,"  in  NASA 
TN  D-1504,  p.  97. 

22.  Low  cited  in  Schnitzer,  "Results  of  Space  Station  Conference,"  p.  1.  Details  concerning 
the  design  of  the  rotating  hexagon  can  be  found  in  Berglund,  "Space  Station  Research 
Configurations,"  pp.  9-21. 

23.  Loftin  quoted  in  U.S.  House  of  Representatives,  Hearing  Before  the  Committee  on  Science 
and  Astronautics,  87th  Cong.,   1st  sess.,   19  May  1961,  pp.  5-9  and  12-18.     For  this 
committee's  analysis  of  the  situation,  see  U.S.  House  of  Representatives,   Expandable 
Space  Structures:   Report  of  the  Committee  on  Science  and  Astronautics,  87th  Cong., 

498 


Notes  for  Chapter  9 

1st  sess.  (Washington:  Government  Printing  Office,   1961),  Serial  m.     Copies  of  both 
publications  are  in  the  Thompson  Collection,  LHA. 

24.  The  quotes  are  from  the  Hearing  Before  the  Committee  on  Science  and  Astronautics, 
pp.  12  13. 

25.  See  Kurzhals,  Adams,  and  Hodge,  "Space  Station  Dynamics  and  Control,"  pp.  76-77. 
See  also  Hook,  "Space  Stations,"  p.  4. 

26.  For  insights  into  Langley's  early  thinking  about  the  power  requirements  of  the  orbiting 
space  station,  see  the  records  of  the  Nichols  committee  in  A200-4,  LCF,  esp.  the  reports  of 
the  subcommittee  on  auxiliary  power  chaired  by  William  J.  O'Sullivan.  The  issues  of  the 
power  supply  for  the  inflatable  torus  and  rotating  hexagon  are  clarified  in  John  R.  Dawson 
and  Atwood  R.   Heath,   Jr.,    "Space  Station  Power  Systems,"    in  NASA  TN  D-1504, 
pp.  59-71.    For  an  account  of  the  debate  at  Langley  between  those  favoring  solar  and 
nuclear  power  for  a  space  station,  see  Hook,  "Space  Stations,"  p.  5,  and  listen  to  Beverly 
McMillan's  tape-recorded  interview  with  Langley  space  station  researcher  William  N. 
Gardner,  Newport  News,  Va.,  26  Sept.  1991,  OHC,  LHA.    See  also  Gardner,  "A  Review 
of  Langley  Space  Station  Studies,"   in  Compilation  of  Papers  Presented  at  the  Space 
Technology  Symposium,  NASA  Langley,  Feb.  1969,  Langley  Technical  Library. 

27.  H.  A.   Storms,   president,   Space  and  Information  Systems  Division,   North  American 
Aviation,   to  Laurence  K.  Loftin,  Jr.,   technical  advisor  to  director,   NASA  Langley, 
"Transmittal  of  Proposal  for  Engineering  Analysis  of  NASA  Self-Erecting  Space  Station," 
14  July  1961,  A200-4,  LCF;  Gardner  interview  with  McMillan,  23  Sept.  1991.    On  the 
development  of  nuclear  reactors  for  the  space  program,  see  William  R.  Corliss,  SNAP 
Nuclear  Space  Reactors  (Washington:    U.S.  Atomic  Energy  Commission,  Sept.  1966). 
The  Atomic  Energy  Commission  initiated  SNAP,  which  stood  for  Systems  for  Nuclear 
Auxiliary  Power,  in  1955  specifically  to  develop  a  nuclear  power  system  for  a  space  vehicle. 
The  commission's  effort  involved  both  nuclear  reactors  and  radioisotope  systems.  NASA 
Langley  researchers  learned  quite  a  bit  about  SNAP  findings  during  the  early  1960s,  as 
they  struggled  to  choose  the  right  power  plant  for  their  space  station  designs.  The  first 
SNAP  system  actually  to  go  into  space  was  SNAP  10- A,  a  500- watt  system  launched 
from  Vandenberg  AFB  on  13  Apr.  1965.  See  Corliss,  SNAP,  pp.  59-66  and  108-118. 

28.  See  the  reports  of  the  Subcommittee  on  Radioisotope  Power  to  the  Langley  MORL 
Technology  Steering  Committee  from  1963  and  1964,  A200-4,  LCF. 

29.  Dan  C.  Popma,  Charles  H.  Wilson,  and  Franklin  W.  Booth,  "Life  Support  Research  for 
Manned  Space  Stations,"  in  NASA  TN  D-1504,  pp.  107-120. 

30.  Ibid.,  pp.  108-112. 

31.  W.  Ray  Hook  interview  with  McMillan,  Hampton,  Va.,  16  Sept.  1991,  in  OHC,  LHA; 
Gardner  to  McMillan,  23  Sept.  1991. 

32.  Hook  interview  with  McMillan,  16  Sept.  1991. 

33.  Joseph  S.  Algranti,  Donald  L.  Mallick,  and  Howard  G.  Hatch,  Jr.,  "Crew  Performance 
on  a  Lunar-Mission  Simulation,"  in  NASA  TN  D-1504,  pp.  135-146. 

34.  George  W.  Zender  and  John  R.  Davidson,  "Structural  Requirements  of  Large  Manned 
Space  Stations,"  in  NASA  TN  D-1504,  pp.  33-44. 

35.  Ibid.;  Robert  S.  Osborne  and  George  P.  Goodman,  "Material  and  Fabrication  Techniques 
for  Manned  Space  Stations,"  in  NASA  TN  D-1504,  pp.  45-53. 

36.  See  Berglund,  "Space  Station  Research  Configurations,"  pp.  12-15. 

499 


Notes  for  Chapter  9 

37.  Ibid.,  pp.  13-14. 

38.  Emanuel  Schnitzer,  memorandum  for  files,   "Orbital  Flight  Test  of  10-Foot  Inflatable 
Space  Station  Model  on  Scout— Details,"  10  May  1961,  A200-4,  LCF. 

39.  Hill  and  Schnitzer,  "Space  Station  Objectives  and  Research  Guidelines,"  p.  2.  For  similar 
statements  about  the  ultimate  use  of  a  space  station,  as  conceived  at  Langley  in  the 
early  1960s,  see  Loftin's  testimony  in  the  Hearing  Before  the  Committee  on  Science  and 
Astronautics,  pp.  6,  9,  and  13-14. 

40.  Loftin's  testimony,  pp.  13-14. 

41.  Hook,  "Space  Stations,"  p.  4;  see  also  the  description  of  the  MORL  concept  in  the  NASA 
Langley  press  release,  "Industry  Asked  to  Propose  Manned  Orbiting  Research  Laboratory 
Plans,"  24  Apr.  1963,  A200-4,  LCF. 

42.  NASA  Langley  press  release,   "Industry  Asked  to  Propose  Manned  Orbiting  Research 
Laboratory  Plans."    Hook  interview  with  McMillan,  16  Sept.  1991;  Gardner  interview 
with  McMillan,  23  Sept.  1991. 

43.  Floyd  L.  Thompson,  Langley  Research  Center  Announcement  No.  32-63,  "Reorganiza- 
tion of  Langley  Research  Center  and  reassignment  of  personnel  effective  June  10,  1963," 
6  June  1963;  Thompson,  Langley  Research  Center  Announcement  No.  33-63,  "Changes 
in  Organization  of  the  Applied  Materials  and  Physics  Division,"  6  June  1963.  Copies  of 
both  announcements  are  preserved  in  the  Thompson  Collection,  LHA.    For  the  minutes 
of  the  meetings  of  the  Langley  MORL  Technology  Steering  Committee,  as  well  as  the 
weekly  and  (later)  biweekly  reports  of  the  MORL  Studies  Office,  see  A200-4,  LCF. 

44.  Gardner  interview  with  McMillan,  23  Sept.  1963. 

45.  William  N.  Gardner,  head,  MORL  Studies  Office,  to  Associate  Director  [Donlan],  "MORL 
Studies  Office  Weekly  Report—Week  ending  June  22,  1963,"  24  June  1963,  A200-4,  LCF. 

46.  W.  Nissim,  Douglas  Aircraft  Co.,  "London  Daily  Mail  Astronomical  Space  Observatory," 
Douglas  Report  No.  SM-36173,  Nov.  1959,  copy  in  Langley  Technical  Library. 

47.  C.  C.  Walkey,  Contracts  Division,  Douglas  Aircraft  Company,  to  Laurence  K.  Loftin,  Jr., 
"Unsolicited  Proposal  for  Advanced  Technology  Studies  Related  to  Orbiting  Laborato- 
ries," 13  July  1964,  A200-4,  LCF. 

48.  Charles  J.  Donlan,  chairman,  Source  Evaluation  Board,  Langley  Research  Center  An- 
nouncement No.  63-63,   "Establishment  of  Technical  Assessment  Team  for  Evaluation 
of  Manned  Orbital  Research  Laboratory  (MORL)   Phase  I  Study,   Douglas  Contract 
NAS1-2974  and  Boeing  Contract  NAS1-2975,  for  Comparative  Studies  and  Preliminary 
Design  of  an  MORL  Concept,"  24  Sept.  1963,  Thompson  Collection,  LHA. 

49.  NASA  Langley  press  release,  "NASA  Selects  Douglas  to  Continue  Space  Station  Study," 
2  Dec.  1963,  A200-4,  LCF. 

50.  For  some  contemporary  thoughts  on  manned  space  stations  from  a  military  vantage  point, 
see  John  W.  Massey,  "Historical  Resume  of  Manned  Space  Stations,"  ABMA,  Redstone 
Arsenal,  Ala.,  Report  No.  DSP-TM-9-60,   15  June  1960,  HQA.  On  the  history  of  the 
U.S.  military's  interest  in  a  space  station,  see  McCurdy,  Space  Station  Decision,  esp. 
pp.  101-105,  131-134,  149-150,  166-168,  and  198-199.    See  also  Emanuel  Schnitzer's 
11-page  report    (once  classified    "Secret"),    "Military   Test   Space   Station — Details   of 
Midterm  Briefing  at  WADD  [Wright  Air  Development  Division — Wright  Patterson  AFB, 
Ohio]— Preliminary  Station  for  Pre-1965,   Feb.  7-9,   1961,"    23  Feb.   1961,   as  well  as 
Alfred  C.  Loedding,  ARDC  [Air  Research  and  Development  Command]  Liaison  Office, 

500 


Notes  for  Chapter  9 

Langley  Research  Center,  memorandum  for  files,  "Man  in  Space  in  1963  for  Military 
Planning,"  6  July  1960;  both  documents  are  in  A200-4,  LCF.  As  yet,  no  complete 
(unclassified)  history  of  the  DOD's  involvement  in  space  station  planning  has  been 
written. 

51.  Douglas  Aircraft   Company,    Missile  and  Space  Systems  Division,    "MORL:   Manned 
Orbital  Research  Laboratory,"  Douglas  Report  SM-47966,  Aug.  1964,  Langley  Technical 
Library.  For  Douglas's  follow-up  work,  see  "Report  on  the  Optimization  of  the  Manned 
Orbital  Research  Laboratory  (MORL)  System  Concept,"   Douglas  Report  SM-46071, 
Sept.  1964,  and  "Report  on  the  Development  of  the  Manned  Orbital  Research  Laboratory 
(MORL)  System  Utilization  Potential,"  Douglas  Report  SM-48922,  Jan.  1966.  Copies  of 
both  are  also  available  in  the  Langley  Technical  Library. 

52.  Douglas,   "MORL,"  Aug.  1964,  p.  3.    For  a  historical  summary  of  early  space  station 
work  at  Johnson  Space  Center  in  Houston,  see  Henry  C.  Dethloff,  Suddenly,  Tomorrow 
Came  .  .  .:  A  History  of  the  Johnson  Space  Center,  NASA  SP-4307  (Houston,  1993), 
pp.  188-196  and  221  226. 

53.  On  the  air  force's  MOL  concept,   see  McCurdy,   Space  Station  Decision,  pp.  70  and 
132  133,  and  McDougall,  Heavens  and  the  Earth,  pp.  340-341.    The  DOD  announced 
its  plans  for  MOL  in  Dec.  1963,  one  day  after  terminating  the  X-20  Dyna-Soar  program. 

54.  John  R.  Dawson,  secretary,  "Minutes  of  Meeting  of  Langley  MORL  Technology  Steering 
Committee,  October  28,  1963,"  28  Oct.  1963,  A200-4,  LCF;  NASA  Langley  press  release, 
"NASA  Selects  Douglas,"  p.  1. 

55.  Gardner  interview  with  McMillan,  23  Sept.  1991. 

56.  Robert  S.  Osborne  interview  with  McMillan,  Newport  News,  Va.,  21  Nov.  1991. 

57.  Numerous  references  to  the  General  Dynamics  contract  for  the  life-support  system  can 
be  found  in  the  minutes  to  the  meetings  of  the  Langley  MORL  Technology  Steering 
Committee  and  in  the  weekly  and  biweekly  reports  of  the  MORL  Studies  Office  from 
mid-1963  to  1964. 

58.  Hook  interview  with  McMillan,  16  Sept.  1991;  Gardner  interview  with  McMillan,  23  Sept. 
1991. 

59.  Osborne  interview  with  McMillan,  21  Nov.  1991.   See  also  the  minutes  of  the  meetings 
of  the  MORL  Technology  Steering  Committee  and  the  reports  of  the  MORL  Studies 
Office  from  late  1964  and  1965.  On  the  Apollo  Extension  System  Program,  see  Wernher 
von  Braun  and  Frederick  I.  Ordway  III,  Space  Travel,  A  History,  4th  ed.  (New  York: 
Harper  &  Row,  1985),  p.  225. 

60.  Charles  W.  Matthews,  director  of  Apollo  applications,  NASA  Manned  Spacecraft  Center, 
to  Charles  J.  Donlan,  acting  director,  NASA  Langley,  25  Apr.  1968,  A200-4,  LCF;  NASA 
Langley,  MORL  Studies  Office,  "Intermediate  Orbital  Workshop  System  Study,  In-House 
Study  Summary,"  28  June  1968,  Langley  Technical  Library.  On  the  Apollo  Applications 
Program,  see  Collins,  Liftoff,  pp.  164-165;  von  Braun  and  Ordway,  Space  Travel:    A 
History,  p.  225;  Linda  Neuman  Ezell,  NASA  Historical  Data  Book,  vol.  3,  Programs  and 
Projects,  1969-1978,  NASA  SP-4012  (Washington,  1988),  pp.  56-57,  63-64,  and  94-100. 

61.  MORL  Studies  Office,  "Intermediate  Orbital  Workshop  System  Study,"  28  June  1968, 
viewgraph  LRC-8-2. 

62.  Ezell,  NASA  Historical  Data  Book,  3:100-101. 


501 


Notes  for  Chapter  1 0 

63.  See  Robert  R.  Gilruth,  "Manned  Space  Stations,"  in  Proceedings  of  the  Fourth  Interna- 
tional Symposium  on  Bioastronautics  and  the  Exploration  of  Space,  San  Antonio,  Tex., 
24-27  Apr.   1968,   pp.   153-177;   also  William  J.  Normyle,    "NASA  Aims  at  100-Man 
Station,"  Aviation  Week  &  Space  Technology,  24  Feb.  1969,  pp.  14-17. 

64.  Hook,  "Space  Stations,"  p.  6.    For  insights  into  the  Apollo  Applications  Program  and 
the  plans  leading  to  Skylab,  see  W.  David  Compton  and  Charles  D.  Benson,  Living  and 
Working  in  Space:  A  History  of  Skylab,  NASA  SP-4208  (Washington,  1983). 

65.  Compton  and  Benson,    Living  and   Working  m  Space.     See  McCurdy,    Space  Station 
Decision,  pp.  23-24;  and  Ezell,  NASA  Historical  Data  Book,  3:100-101.  See  also  North 
American  Rockwell,  Space  Division,  "Space  Station  Program  Phase  B  Definition,  Second 
Quarterly  Progress  Report,"    Report  SD  70-146,   13  Mar.   1970,  and  North  American 
Rockwell,  "Modular  Space  Station  Phase  B  Extension,  NASA  Administrator's  Review," 
Report  SD  71-582,  3  Dec.  1971.    Copies  of  both  reports  can  be  found  in  the  Langley 
Technical  Library. 

66.  McCurdy,  Space  Station  Decision,  pp.  27-28. 

67.  Gardner  interview  with  McMillan,  23  Sept.  1991. 

68.  Hook  interview  with  McMillan,  Hampton,  Va.,  16  Sept.  1991. 

69.  Gardner  interview  with  McMillan,  23  Sept.  1991. 

Chapter  10 
To  Behold  the  Moon:    The  Lunar  Orbiter  Project 

1.  See  R.   Cargill  Hall,    Lunar  Impact:     A  History  of  Project  Ranger,   NASA  SP-4210 
(Washington,    1977),    p.   285,   and  Michael  Collins,    Liftoff:      The  Story  of  America's 
Adventure  in  Space  (New  York:    NASA/Grove  Press,  1988),  p.  117. 

2.  Leonard  Roberts,  "Exhaust  Jet-Dust  Layer  Interaction  during  a  Lunar  Landing,"  unclas- 
sified report  presented  at  the  13th  International  Aeronautical  Congress,  Varna,  Bulgaria, 
24-28  Sept.  1962;   "The  Interaction  of  a  Rocket  Exhaust  with  the  Lunar  Surface,"  un- 
classified report  presented  at  the  Specialists'  Meeting  of  the  Fluid  Dynamical  Aspects 
of  Space  Flight,  Marseilles,  France,  20-24  Apr.  1964.   Copies  of  both  papers  are  in  the 
Langley  Technical  Library. 

3.  See  Hall's  comprehensive  and  well  told,  Lunar  Impact. 

4.  A  complete  history  of  the  Surveyor  program  has  not  been  published.    In  addition  to 
Hall's  Lunar  Impact,  see  Homer  E.  Newell,  Beyond  the  Atmosphere:   Early  Years  of  Space 
Science,  NASA  SP-4211  (Washington,  1980),  pp.  263-272.  A  fine  summary  of  Surveyor 
is  contained  in  Linda  Neuman  Ezell,  NASA  Historical  Data  Book,  vol.  2,  Programs  and 
Projects  1958-1968,  NASA  SP-4012  (Washington,  1988),  pp.  325-331.  For  an  analysis  of 
NASA's  problems  in  the  management  of  the  Surveyor  project,  see  Erasmus  H.  Kloman, 
Unmanned  Space  Project  Management:  Surveyor  and  Lunar  Orbiter,  NASA  SP-4901 
(Washington,  1972). 

5.  The  chairman  of  this  Lunar  Photographic  Mission  Study  Group  was  Capt.  Lee  R.  Scherer, 
a  naval  officer  on  assignment  to  NASA  headquarters  and  a  program  engineer  with  the 
Surveyor  project.  In  Oct.  1962,  NASA  gave  him  the  job  of  heading  the  Office  of  Space 
Sciences/Office  of  Manned  Space  Flight  working  group,   which  was  to  identify  what 
information  about  the  moon  would  be  most  essential  to  the  landing  mission.  See  Scherer, 

502 


Notes  for  Chapter  10 

Study  of  Agena-based  Lunar  Orbiters,  NASA  headquarters,  Office  of  Space  Sciences, 
25  Oct.  1962,  copy  in  Langley  Technical  Library.  See  also  Bruce  K.  Byers,  Destination 
Moon:  A  History  of  the  Lunar  Orbiter  Program,  NASA  TM  X-3487  (Washington,  1977, 
multilith),  pp.  20-23.  Byers'  study  is  the  most  complete  history  to  date  of  Lunar  Orbiter. 

6.  Kloman,  Unmanned  Space  Project  Management. 

7.  Clinton  E.  Brown  interview  with  author,  Hampton,  Va.,  17  July  1989.    Brown  acted 
as  Langley's  spokesman  in  early  discussions  at  NASA  headquarters  in  1963  regarding 
Langley 's  proposed  management  of  the  Lunar  Orbiter  program. 

8.  Hall,  Lunar  Impact,  p.  157. 

9.  "NASA  Minutes  of  the  First  Senior  Council  for  the  Office  of  Space  Sciences,"  7  June  1962, 
E1-2A,  LCF.    Also  quoted  in  Hall,  Lunar  Impact,  p.  157. 

10.  Penetrometer  Feasibility  Study  Group,  Langley  Research  Center,  "Preliminary  Project 
Development  Plan  for  a  Lunar  Penetrometer  Experiment  for  the  Follow-On  Ranger  Pro- 
gram,"  18  Aug.  1961,  Code  N- 101, 24218,  Langley  Technical  Library.    See  also  the  fol- 
lowing memoranda:    E.  M.  Cortright  (for  Abe  Silverstein,  director  of  Space  Flight  Pro- 
grams), to  Ira  H.  Abbott,  director  of  Advanced  Research  Programs,   "Lunar  Surface 
Hardness  Experiments,"  29  June  1961;    E.  C.  Kilgore,  assistant  chief,  Engineering  Di- 
vision, to  Charles  J.  Donlan,  associate  director,  "Meeting  at  Jet  Propulsion  Laboratory 
on  July  18,  1961,  attended  by  representatives  of  Langley  Research  Center,  Aeronutronics 
Division  of  Ford  Motor  Company,  NASA  Headquarters,  and  JPL,  to  discuss  a  penetrom- 
eter  experiment  for  the  follow-on  Ranger  program,"   21  July  1961;  John  L.  McCarty 
and  George  W.  Brooks,   "Visit  to  NASA  Headquarters,  Washington,  D.C.,  by  George 
W.  Brooks  and  John  L.  McCarty,  Vibration  and  Dynamics  Section,  DLD  [Dynamic 
Loads  Division],  June  28,  1961,  to  discuss  possible  Lunar  Penetrometer  Experiment  on 
Ranger  Spacecraft,"  7  July  1961;  Floyd  L.  Thompson  to  NASA  headquarters,  "Lunar 
surface  hardness  experiments,"  27  July  1961;  and  George  W.  Brooks  to  Associate  Di- 
rector Charles  J.  Donlan,  "Langley  action  on  lunar  penetrometer  experiment  for  Ranger 
follow-on  program,"  27  July  1961.  All  these  memos  are  in  LCF,  A200-1B. 

11.  See  John  L.  McCarty  and  Huey  D.  Garden,  "Impact  Characteristics  of  Various  Materials 
Obtained  by  an  Acceleration-Time-History  Technique  Applicable  to  Evaluating  Remote 
Targets,"  NASA  TN  D-1269,  June  1962. 

12.  Penetrometer  Feasibility  Study  Group,   "Preliminary  Project  Development  Plan  for  a 
Lunar  Penetrometer  Experiment,"  p.  1. 

13.  Kilgore  to  Donlan,  "Meeting  at  Jet  Propulsion  Laboratory." 

14.  Hall,  Lunar  Impact,  p.  158. 

15.  Ibid.,  pp.  156-163. 

16.  Oran  Nicks  quoted  in  Byers,  Destination  Moon,  p.  26. 

17.  Floyd  L.  Thompson  to  Dr.  Eugene  Emme,  NASA  historian,  "Comments  on  draft  of  Lunar 
Orbiter  History  dated  November  4,  1969,"  22  Dec.  1969,  in  folder  labeled  "Lunar  Orbiter 
Historical  Notes"  in  Floyd  L.  Thompson  Collection,  LHA.  See  also  Byers,  Destination 
Moon,  p.  25. 

18.  Clinton  E.  Brown  interview,  17  July  1989. 

19.  Ibid.     On  12  July  1961,  six  weeks  after  President  Kennedy's  lunar  landing  speech, 
three  men  in  Clint  Brown's  Theoretical  Mechanics  Division — William  H.  Michael,  Jr., 
Robert  H.  Tolson,  and  John  P.  Gapcynski — put  forward  a  "Preliminary  Proposal  for  a 

503 


Notes  for  Chapter  10 

Circumlunar  Photographic  Experiment."  The  idea  essentially  was  to  support  the  Apollo 
program  by  adapting  the  Ranger  spacecraft  so  that  it  could  perform  a  circumlunar  mission 
that  could  take  high-resolution  color  photographs  during  the  lunar-approach  phase  of  a 
"single-pass,"  circumlunar  trajectory.  In  the  cover  memorandum  to  this  unpublished 
proposal,  one  of  its  authors,  Bill  Michael,  wrote  that  "a  desirable  situation  would  be 
that  of  Langley  having  prime  responsibility  for  the  photographic  experiment,  in  a  role 
similar  to  that  of  chief  experimenter  in  other  specific  experiments  carried  by  the  Ranger 
and  other  vehicles."  See  William  H.  Michael,  Jr.,  head,  Mission  Analysis  Section,  to 
Langley  Associate  Director  Charles  J.  Donlan,  "Preliminary  Proposal  for  a  Circumlunar 
Photographic  Mission,"  17  July  1961.  See  also  C.  I.  Cummings,  Lunar  Program  Director, 
JPL,  to  Bernard  Maggin,  Office  of  Programs,  NASA  headquarters,  9  Nov.  1961,  and 
Maggin  to  Clinton  E.  Brown,  14  Nov.  1961.  Copies  of  the  above  documents  are  in  LCF, 
A200-1B. 

20.  Israel  Taback  interview  with  author,  Hampton,  Va.,  13  Aug.  1991. 

21.  Floyd  L.  Thompson  to  NASA  headquarters  (Attn:     Capt.  Lee  Scherer),  6  Mar.  1963, 
A200-1B,  LCF.     Attached  to  this  memo  is  a  "system  block  diagram"   for  the  Lunar 
Orbiter  as  well  as  the  data  from  the  mission  reliability  analysis. 

22.  On  the  genesis  of  NASA's  system  for  project  management,   see  Robert  L.  Rosholt, 
An  Administrative  History  of  NASA,  1958-1963,  NASA  SP-4101  (Washington,  1966), 
pp.  145-158  and  273-276. 

23.  NASA  Management  Manual,  pt.   1,    General  Management  Instruction,  Number  4-1-1, 
"Subject:    Planning  and  Implementation  of  NASA  Projects,"  NASA  GMI  (Washington, 
18  Jan.  1961),  chap.  4,  p.  4.  See  also  Edgar  M.  Cortright  interview  with  author,  Yorktown, 
Va.,  July  1988,  transcript,  pp.  8-9,  OHC,  LHA. 

24.  Donald  H.  Ward,  One  Engineer's  Life  Relived  (Utica,  Ky.:  McDowell  Publications,  1990), 
p.  61.  Ward  served  as  head  of  spacecraft  launch  operations  for  LOPO. 

25.  Taback  interview,  13  Aug.  1991. 

26.  See  Byers,  Destination  Moon,  pp.  40-41. 

27.  See  Dr.  A.  K.  Thiel,  STL,  to  Oran  W.  Nicks,  director,  Lunar  and  Planetary  Programs, 
OSSA/NASA,  Washington,  D.C.,  20  Sept.  1962,  copy  in  Thompson  Collection,  LHA.  See 
also  Byers,  Destination  Moon,  pp.  16-17. 

28.  See  Byers,  Destination  Moon,  pp.  43-44. 

29.  Interview  with  Sherwood  Butler,  Hampton,  Va.,  23  Aug.  1991.  Butler  was  Langley's  chief 
procurement  officer  at  the  time.  For  an  analysis  of  the  benefits  of  incentives  contracting 
for  Lunar  Orbiter,  see  Kloman,   Unmanned  Space  Project  Management,  pp.  34-36.   See 
also  Byers,   Destination  Moon,  pp.  39-40.     For  a  contemporary  news  story  covering 
the  details  of  the  novel  incentives  contract  for  Lunar  Orbiter,  see  Richard  G.  O'Lone, 
"Orbiter  is  First  Big  NASA  Incentive  Job,"    Aviation   Week  &  Space   Technology  79 
(7  Oct.  1963):    32. 

30.  Taback  interview,  13  Aug.  1991. 

31.  See  Byers,  Destination  Moon,  pp.  40-47. 

32.  Ibid.,  pp.  43-44. 

33.  Taback  interview,  13  Aug.  1991;    see  Byers,  Destination  Moon,  p.  56. 

34.  Byers,  Destination  Moon,  pp.  57-70. 


504 


Notes  for  Chapter  1 0 

35.  For  the  basics  of  the  Lunar  Orbiter  camera  system,  see  Leon  J.  Kosofsky  and  S.  Calvin 
Broome,   "Lunar  Orbiter:     A  Photographic  Satellite,"   paper  presented  to  the  Society 
of  Motion  Picture  and  Television  Engineers,  Los  Angeles,  28  Mar.-2  Apr.  1965,  copy 
in  Langley  Technical  Library.     Broome  was  head  of  the  photo  subsystem  group  in 
LOPO;  Kosofsky  was  the  camera  expert  in  Lee  Scherer's  office  at  NASA  headquarters. 
For  a  more  general  description  of  the  photographic  mission  and  the  Eastman  Kodak 
camera,  see   The  Lunar  Orbiter  (revised  Apr.   1966),   a  38-page  booklet  prepared  by 
Boeing's  Space  Division  and  published  by  NASA  Langley,  esp.  pp.  18-20  and  22-26, 
and  The  Lunar  Orbiter:  A  Radio- Controlled  Camera,  a  glitzy  14-page  brochure  that  was 
published  by  NASA  Langley  with  Boeing's  assistance  after  the  Lunar  Orbiter  project 
had  ended. 

36.  Kosofsky  and  Broome,    "Lunar  Orbiter:     A  Photographic  Satellite."     See  also  Israel 
Taback,    "A  Description  of  the  Lunar  Orbiter  Spacecraft,"   reprinted  in  Highlights  of 
Astronomy  (Dordrecht,  Netherlands:    D.  Reidel  Publishing  Co.,  1968),  p.  464.   Taback 
presented  this  paper  at  the  13th  General  Assembly  of  the  International  Astronomical 
Union  (IAU)  in  Prague,  Czechoslovakia,  1967.   At  this  meeting,  renowned  astronomers 
from  all  over  the  world  took  off  their  shoes  and  crawled  around  on  the  floor  looking  at  a 
huge  layout  of  photographs  taken  by  the  Lunar  Orbiters. 

37.  Scherer,  Study  of  Agena-based  Lunar  Orbiters;  see  Byers,  Destination  Moon,  pp.  20-23. 

38.  Telephone  interview  with  Thomas  Costello  of  the  Boeing  Co.,  Colorado  Springs,  Colo., 
15  Aug.  1961.    Costello  was  an  engineer  with  Boeing  who  worked  on  the  company's 
proposal  for  Lunar  Orbiter  and  then  served  as  one  of  its  project  engineers  from  1963  to 
1966. 

39.  See  "Boeing  to  Build  Lunar  Orbiter,"  Aviation  Week  &  Space  Technology  79  (30  Dec. 
1963):  22;    "NASA  To  Negotiate  with  Boeing  for  Lunar  Orbiter  Spacecraft,"   NASA 
Langley  Researcher,  2  June  1963;  and  "NASA  Explains  Choice  of  Boeing  over  Hughes  in 
Lunar  Orbiter  Award,"  Missiles  and  Rockets  14  (9  Mar.  1964):    13. 

40.  Dr.  Trutz  Foelsche,    "Remarks  on  Doses  Outside  the  Magnetosphere,  and  on  Effects 
Especially  on  Surfaces  and  Photographic  Films,"   paper  presented  at  the  Meeting  to 
Discuss  Charged  Particle  Effects,  OART,  19-20  Mar.  1964,  Washington,  D.C.,  p.  8,  copy 
in  the  Langley  Technical  Library. 

41.  Byers,  Destination  Moon,  pp.  72-74. 

42.  See  Lee  R.  Scherer,  The  Lunar  Orbiter  Photographic  Missions,  p.  2.   This  is  a  20-page 
typescript  booklet  published  by  NASA  Langley  in  late  1967. 

43.  Erasmus  H.  Kloman,  "Organizational  Framework:  NASA  and  Langley  Research  Center," 
p.  7.  This  is  a  chapter  draft  from  Kloman's  comment  copy  of  his  subsequent  Unmanned 
Space  Project  Management.  In  draft  form,  Kloman's  book  contained  many  more  details 
including  the  names  of  responsible  individuals,  which  were  omitted  in  the  shortened 
version  published  by  NASA.    The  author  wishes  to  thank  Thomas  R.  Costello,  former 
engineer  in  the  Boeing  Lunar  Orbiter  project  office,  for  making  available  a  copy  of 
Kloman's  comment  edition. 

44.  The  portraits  of  Cliff  Nelson  and  James  Martin  are  derived  from  statements  made  to  the 
author  by  several  people  associated  with  LOPO. 

45.  Kloman,  "Organizational  Framework:   NASA  and  Langley  Research  Center,"  pp.  10-11. 

46.  Kloman,  Unmanned  Space  Project  Management,  pp.  18  19  and  38. 

505 


Notes  for  Chapter  10 

47.  Taback  interview,  13  Aug.  1991. 

48.  Norman  L.  Crabill  interview,  Hampton,  Va.,  28  Aug.  1991. 

49.  Kloman,  "Organizational  Framework,"  p.  9. 

50.  Taback  interview,  13  Aug.  1991.  See  also  the  press  release  from  News  Bureau,  the  Boeing 
Co.,  Seattle,  "Robert  J.  Helberg,  Lunar  Orbiter  Program  Manager,"  27  Sept.  1965. 

51.  Kloman,  Unmanned  Space  Project  Management,  p.  25. 

52.  Taback  interview,  13  Aug.  1991. 

53.  The  author  wishes  to  thank  LOPO  member  Norman  L.  Crabill  for  his  careful  explanation 
of  the  mission  planning  for  Lunar  Orbiter.     Crabill  interview,   28  Aug.   1991.     For  a 
lengthy  written  account  of  mission  planning  in  relation  to  the  design  of  the  Lunar 
Orbiter  spacecraft  system,  see  Thomas  T.  Yamauchi,  the  Boeing  Co.,  and  Israel  Taback, 
NASA  Langley,  "The  Lunar  Orbiter  System,"  paper  presented  at  the  6th  International 
Symposium  on  Space  Technology  and  Science,  Tokyo,  1965,  copy  in  the  Langley  Technical 
Library. 

54.  For  the  details  about  the  evolution  of  the  early  mission  plans  for  Lunar  Orbiter,  see 
Byers,  Destination  Moon,  pp.  177-194. 

55.  Crabill  interview,  28  Aug.  1991. 

56.  Ibid. 

57.  Ibid. 

58.  Ibid. 

59.  Ibid. 

60.  Ibid.;  also  see  A.  Thomas  Young  to  Crabill,  "Mission  Reliability  Analyses  and  Compari- 
son for  the  BellComm  Mission  and  TBC's  S-110  Mission,"  A200-1B,  LCF. 

61.  By  the  end  of  the  project,  Boeing  earned  nearly  an  extra  $2  million  dollars  in  incentives. 
See  Newport  News  (Va.)  Daily  Press,  "Orbiter  Incentive  Award  is  $1.9  Million,"  28  Jan. 
1967;   "Lunar  Orbiter  Successes  Earn  Boeing  $1,811,611,"    17  Nov.  1967;  and  "Lunar 
Success  Pays  Dividend  for  Boeing  Co.,"  18  Nov.  1967. 

62.  Crabill  interview,  28  Aug.  1991. 

63.  See  L.  C.  Rowan,   "Orbiter  Observations  of  the  Lunar  Surface,"   AAS  (American  As- 
tronomical Society)  Paper,  29  Dec.  1966;  and  L.  R.  Scherer  and  C.  H.  Nelson,   "The 
Preliminary  Results  from  Lunar  Orbiter  I,"  in  Spacecraft  Systems,  vol.  1,  International 
Astronautical  Federation,   17th  International  Astronautical  Congress,   Madrid,   Spain, 
9-15  Oct.  1966. 

64.  See  Byers,  Destination  Moon,  pp.  241-243. 

65.  Taback  interview,  13  Aug.  1991. 

66.  I  have  not  been  able  to  track  down  the  exact  source  of  the  phrase,  "the  picture  of  the 
century,"  which  came  to  be  used  generally  to  describe  the  historic  first  picture  of  the  earth 
from  deep  space.   A  number  of  journalists  used  it  in  the  weeks  following  the  release  of 
the  photographs  taken  by  Lunar  Orbiter  I.   One  might  guess  that  a  NASA  public  affairs 
officer  invented  it,  but  there  is  better  reason  to  think  that  someone  at  Eastman  Kodak 
coined  the  phrase.  In  January  1967,  the  company  unveiled  a  rendition  of  the  remarkable 
photograph  on  the  huge  Kodak  Colorama  inside  Grand  Central  Station  in  New  York  City. 
The  Kodak  caption  indeed  called  it  "the  picture  of  the  century."   This  phrase,  however, 
was  also  used  to  describe  other  Lunar  Orbiter  photographs.    For  example,  Boeing  and 
NASA  Langley  used  it  in  "The  Lunar  Orbiter/A  Radio-Controlled  Camera,"  a  brochure 

506 


published  in  1968  (copy  in  LHA,  Ames  Collection,  box  6),  not  to  caption  the  earth 
shot,  but  to  dramatize  a  stereoscopic  picture  of  the  Copernicus  crater  (see  p.  350  of  this 
book).  Veterans  of  the  Lunar  Orbiter  project  team  at  Langley  remember  the  earth  shot 
as  "the  real  picture  of  the  century,"  however,  partly  because  they  know  the  story  of  how  it 
almost  did  not  get  taken.  See  Langley  Researcher  News,  "Reliving  a  Moment  in  History," 
6  Sept.  1991,  p.  5. 

67.  See  Lunar  Orbiter  Photo  Data  Screening  Group,   "Preliminary  Geological  Evaluation 
and  Apollo  Landing  Analysis  of  Areas  Photographed  by  Lunar  Orbiter  II,"  LWP-363, 
Mar.  1967,  copy  in  Milton  Ames  Collection,  LHA. 

68.  Byers,  Destination  Moon,  pp.  243-244. 

69.  Crabill  interview,  28  Aug.  1991. 

70.  Collins,  Liftoff,  p.  8. 

71.  Kloman,  Unmanned  Space  Project  Management,  p.  7. 

72.  Ibid.,  p.  33. 

73.  Ibid.,  p.  22. 

74.  Crabill  interview,  28  Aug.  1991. 

Chapter  11 
In  the  Service  of  Apollo 

1.  President  John  F.  Kennedy,  quoted  in  John  M.  Logsdon,   The  Decision  to  Go  to  the 
Moon:    Project  Apollo  and  the  National  Interest  (Chicago:    University  of  Chicago  Press, 
1976),  p.  128. 

2.  As  one  might  imagine,  the  literature  on  the  Apollo  lunar  landing  program  is  extensive, 
involving  much  more  than  just  historical  treatments.     This  diverse  literature  will  be 
cited  where   most   appropriate.      For  the  details  of  the  involvement   of  the  various 
NASA  centers  in  the  Apollo  program,   the  only  adequate  sources  are  in  the  official 
NASA  History  Series:    Charles  D.  Benson  and  William  Barnaby  Faherty,  Moonport:    A 
History  of  Apollo  Launch  Facilities  and  Operations,  NASA  SP-4204  (Washington,  1978); 
Courtney  G.  Brooks,  James  M.  Grimwood,  and  Loyd  S.  Swenson,  Jr.,    Chariots  for 
Apollo:    A  History  of  Manned  Lunar  Spacecraft,  NASA  SP-4205  (Washington,  1979); 
Roger  E.  Bilstein,  Stages  to  Saturn:  A  Technological  History  of  the  Apollo/Saturn  Launch 
Vehicles,  NASA  SP-4206  (Washington,  1980);  Arnold  Levine,  Managing  NASA  in  the 
Apollo  Era,  NASA  SP-4102  (Washington,  1982);  W.  David  Compton,   Where  No  Man 
Has  Gone  Before:     A  History  of  Apollo  Lunar  Exploration  Missions,  NASA  SP-4214 
(Washington,  1989);  and  Sylvia  Doughty  Fries,  NASA  Engineers  and  the  Age  of  Apollo, 
NASA  SP-4104  (Washington,  1992).  For  more  complete  pictures  of  the  roles  of  Houston 
and  Huntsville  in  the  Apollo  program,  readers  should  see  two  new  NASA-sponsored  books 
on  the  histories  of  these  key  NASA  facilities:    Henry  C.  Dethloff,  Suddenly,  Tomorrow 
Came  .  .  .:  A  History  of  the  Johnson  Space  Center,  NASA  SP-4307  (Lyndon  B.  Johnson 
Space  Center,  Tex.,  1993),  and  Andrew  Dunar  and  Stephen  Waring,  History  of  Marshall 
Space/light  Center  (forthcoming  in  the  NASA  History  Series). 

3.  Laurence  K.  Loftin,  Jr.,  interview  with  author,  Newport  News,  Va.,  8  July  1990. 

4.  Axel  T.  Mattson  interview  with  author,  Hampton,  Va.,  21  July  1989;  only  notes  exist  for 
this  interview. 

507 


Notes  for  Chapter  11 

5.  In  the  LHA  is  a  notebook  labeled  "Research  Organization  Changes/ Announcements," 
which  contains  a  chronological  listing  of  all  the  major  changes  in  the  Langley  organization 
from  1943  to  the  present.  At  one  time  this  book  was  kept  by  the  Langley  staff  office;  for 
the  past  several  years,  it  has  been  updated  by  the  Langley  historical  programs  monitor, 
Richard  T.  Layman.    The  entry  for  28  Mar.  1962  reads:     "Axel  T.  Mattson  appointed 
Research  Assistant  for  Manned  Spacecraft  Projects  reporting  to  Associate  Director — 
relieved  of  duties  as  Assistant  Chief  of  Pull- Scale  Research  Division." 

6.  "Mattson  Named  to  New  Position,"  Langley  Air  Scoop,  6  Apr.  1962,  p.  1. 

7.  Mattson  interview,  21  July  1989,  and  Mattson  interview,  Hampton,  Va.,  14  Aug.  1989, 
transcript,  pp.  52-53,  OHC,  LHA.  There  is  a  recording  and  transcript  for  the  latter 
interview  only. 

8.  Mattson  interview,  14  Aug.  1989,  pp.  53-54. 

9.  Ibid.,  pp.  54-55. 

10.  Ibid.,  p.  76. 

11.  Ibid.,  pp.  55-56. 

12.  Ibid.,  pp.  56-58. 

13.  Mattson  interview,  21  July  1989. 

14.  None  of  the  published  histories  on  Apollo  have  much  to  say  about  the  impact  studies 
conducted  by  North  American  in  Downey.    However,  numerous  technical  reports  have 
been  written  on  this  work  and  can  be  found  in  NASA  technical  libraries  at  Houston 
and  elsewhere.     In  the  LCF,  code   "Apollo  Project,"    I  have  found  numerous  letters 
and  memos  concerning  such  studies.    Several  references  to  this  R&D  aspect  of  Apollo 
are  made  in  the  following  excellent  chronologies:      Ivan  D.   Ertel  and  Mary  Louise 
Morse,    The  Apollo  Spacecraft:     A    Chronology,   vol.    1,    Through  November  7,    1962, 
NASA  SP-4009  (Washington,  1969);  Mary  Louise  Morse  and  Jean  Kernahan  Bays,  The 
Apollo  Spacecraft:    A  Chronology,  vol.  2,  November  8,  1962-September  30,  1964,  NASA 
SP-4009  (Washington,  1973);  and  Courtney  G.  Brooks  and  Ivan  D.  Ertel,   The  Apollo 
Spacecraft:    A  Chronology,  vol.  3,  October  1,  1964~January  20,  1966,  NASA  SP-4009 
(Washington,  1973). 

See  also  the  transcript  of  my  14  Aug.  1989  interview  with  Axel  Mattson  for  a  rather 
complete  personal  account  of  North  American's  (and  Langley's)  work  on  the  impact 
dynamics  of  the  Apollo  capsule. 

15.  Mattson  interview,  14  Aug.  1989,  p.  50. 

16.  Ibid.,  pp.  51-54;  also  Sandy  M.  Stubbs's  telephone  interview  with  author,  29  June  1993. 
Several  memos  pertaining  to  Stubbs's  work  on  the  return  landing  characteristics  of  the 
Apollo  spacecraft  are  in  the  LCF,  "Apollo  Project."    There  is  a  microfiche  copy  of  this 
entire  file  in  the  LHA. 

17.  Mattson  interview,  14  Aug.  1989,  pp.  57-58. 

18.  Ibid.,  p.  55. 

19.  Ibid.,  p.  56. 

20.  Mattson  interview,  14  Aug.  1989,  pp.  56-58.  See  Sandy  M.  Stubbs,  aerospace  engineer, 
Structures  Research  Division,  to  Langley  associate  director,  "Transmittal  of  information 
on  a  water  pressure  landing  investigation  of  a  1/4-scale  model  of  the  Apollo  spacecraft," 
12  Oct.  1964,  Apollo  project  files  (microfiche),  LHA.  For  Stubbs's  work  on  Apollo's 
landing  impact  characteristics,  see  "Landing  Characteristics  of  the  Apollo  Spacecraft 

508 


With  Deployed-Heat-Shield  Impact  Attenuation  Systems,"  NASA  TN  D-3059,  Jan.  1966; 
and  "Dynamic  Model  Investigation  of  Water  Pressures  and  Accelerations  Encountered 
During  Landings  of  the  Apollo  Spacecraft,"  NASA  TN  D-3980,  Sept.  1967.  These  two 
published  papers  contain  references  to  his  earlier  work. 

21.  Mattson's  reports  can  be  found  in  the  Apollo  project  files  (microfiche),  LHA.   The  dates 
of  the  reports  evaluated  in  this  chap,  were:    4  Dec.  1962,  3  Dec.  1963,  17  Nov.  1964, 
30  Nov.  1965,  23  Mar.  1966,  and  19  Feb.  1968. 

22.  For  basic  information  on  Project  Fire,  see  Linda  Neuman  Ezell,  NASA  Historical  Data 
Book,  vol.  2,  Programs  and  Projects,   1958-1968,  NASA  SP-4012  (Washington,  1988), 
pp.  448-451.    The  rocket  lifting  the  Fire  reentry  payloads  was  an  Atlas- Antares;  this 
made  Fire  unique  in  that  it  was  the  only  NASA  project  ever  to  utilize  the  configuration. 
Herbert  A.  Wilson  served  as  Langley's  project  manager  for  Fire  1,  David  G.  Stone  for 
Fire  2.  Under  contract  to  Langley,  the  Chance  Vought  Corp.  of  LTV  built  the  velocity 
package.  Republic  Aviation  built  the  reentry  vehicle. 

For  complete  details  on  Langley's  management  of  this  program,  see  Project  Devel- 
opment Plan:  Flight  Reentry  Research  Project  at  Hypersonic  Velocities,  Project  Fire 
(Langley  Station,  Hampton,  Va.:  Project  No.  714-00-00,  Mar.  1964),  copy  in  LHA.  Also 
preserved  in  the  LHA  is  a  videotape  of  key  moments  in  the  project's  history  from  1963 
to  1965. 

For  an  excellent  technical  summary  of  the  problems  of  "Reentry  from  Lunar  Missions" 
by  a  leading  Langley  engineer,  see  John  V.  Becker's  presentation  for  Administrator  James 
E.  Webb,  5  Aug.  1961,  LHA.  Another  copy  of  this  document,  along  with  an  entire 
collection  of  Becker's  personal  papers,  is  in  the  Archives  for  Aerospace  Exploration  at 
the  library  at  Virginia  Polytechnic  Institute  and  State  University  in  Blacksburg,  Va., 
Becker's  alma  mater. 

23.  Arthur  W.  Vogeley,    "Piloted  Space-Flight  Simulation  at  Langley  Research  Center," 
unpublished  paper  presented  to  the  American  Society  of  Mechanical  Engineers,  winter 
meeting,  27  Nov.-l  Dec.  1966,  copy  in  file  "Langley  Contributions  to  Manned  Space 
Flight,"   Milton  Ames  Collection,  LHA.  In  this  file  are  numerous  items  pertaining  to 
Langley's  development  of  aerospace  simulators.     For  a  more  brief  and  less  technical 
contemporary  overview  of  Langley's  simulation  work  in  support  of  the  lunar  landing 
program,   see  the  NASA  Langley  press  release,    "Crucial  Lunar  Mission  Events  Are 
Duplicated  by  Langley  Simulators,"    18  May   1964,   copy  in   1964  NASA  Inspection 
materials,  LHA. 

24.  On  Project  Gemini,  see  Barton  C.  Hacker  and  James  M.  Grimwood,  On  the  Shoulders 
of  Titans:    A  History  of  Project  Gemini,  NASA  SP-4203  (Washington,  1977),  which  is 
one  of  the  best  books  in  the  NASA  History  Series. 

25.  Michael  Collins,  Liftoff:    The  Story  of  America's  Adventure  in  Space  (New  York:  NASA/ 
Grove  Press,  1988),  pp.  72-73. 

26.  All  the  basic  information  about  these  facilities  is  in  Martin  A.  Weiner,    "Resume  of 
Research  Facilities  at  the  Langley  Research  Center,"    July  1968,  a  copy  of  which  is 
preserved  in  the  LHA.  Weiner,  an  employee  in  Langley's  Research  Models  and  Facilities 
Division,  had  the  job  of  keeping  this  resume  updated  from  the  mid-1960s  into  the  1970s. 


509 


Notes  for  Chapter  1 1 

27.  Arthur  W.  Vogeley  interview  with  author,   Hampton,   Va.,    19  July  1989;   only  notes 
exist  for  this  interview.     For  the  details  about  this  simulator,  see  Vogeley's  "Discus- 
sion of  Existing  and  Planned  Simulators  for  Space  Research,"  unpublished  paper  given 
at  the  Conference  on  the  Role  of  Simulation  in  Space  Technology,  Blacksburg,  Va., 
17-21  Aug.  1964,  as  well  as  Francis  B.  Smith,    "S  for  Manned  Space  Research,"    un- 
published paper  presented  at  the  1966  IEEE  International  Convention,  21-25  Mar.  1966, 
New  York  City.    Copies  of  both  are  in  the  folder  "Langley's  Contributions  to  Manned 
Space  Flight,"  Milton  Ames  Collection,  LHA.  For  various  documents,  reports,  references, 
newspaper  articles,  and  other  clippings  relevant  to  this  simulator,  see  the  folder  labeled 
"Rendezvous  and  Docking  Simulator,"  LHA. 

28.  Quoted  in  Allan  C.  Hanrahan,  "Rendezvous  and  Docking  Maneuvers:    Critical  to  Apollo 
Flight  Plan,"  Langley  Researcher  News,  30  June  1989,  p.  5. 

29.  Quoted  in  Matthew  Paust,   "Brainstorms  Led  to  Practice's  Realism,"   Newport  News 
(Va.)  Daily  Press,  20  July  1989,  p.  A6. 

30.  W.  Hewitt  Phillips  interview  with  author,  Hampton,  Va.,  27  July  1989;  this  interview 
was  not  taped,  but  the  author  took  notes.  Numerous  documents  and  newspaper  clippings 
are  in  a  folder  labeled  "Lunar  Landing  Facility  and  Apollo,"  LHA. 

31.  Neil  A.  Armstrong,    Wingless  on  Luna  (New  York:     The  Wings  Club,  1988),  pp.  8-9. 
This  book  was  produced  from  the  Wings  Club's  25th  General  Harold  R.  Harris  "Sight" 
Lecture  given  by  Armstrong  at  the  Grand  Hyatt  Hotel,  New  York  City,  18  May  1988. 

32.  Richard  P.  Hallion,   On  the  Frontier:     Flight  Research  at  Dryden,   1946  1981,  NASA 
SP-4303  (Washington,  1984),  p.  146. 

33.  Mattson  interview,  21  July  1989;  see  also  Armstrong,  Wingless  on  Luna,  pp.  6-11. 

34.  Vogeley  interview,  19  July  1989.  According  to  an  official  NASA  statement  from  May  1964, 
LOLA  was  not  built  as  a  training  device  for  astronauts  but  as  "a  research  tool  with 
which  NASA  scientists  can  establish  a  fundamental  understanding  in  the  laboratory  of 
the  problems  associated  with  the  complex  task  of  approaching  the  Moon  and  preparing 
for  a  landing  on  its  surface."  See  Lee  Dickinson,  Langley  Public  Affairs  Office,  to  Mr.  Ken 
Allen,  Beckman  Instruments,  Fullerton,  Calif.,  6  May  1964,  Apollo  project  files,  LHA. 

35.  Gilruth  quoted  in  Michael  Collins,  Liftoff,  p.  61. 

36.  On  Rogallo  and  the  idea  of  using  his  paraglider  concept  for  Project  Gemini,  see  Hacker 
and  Grimwood,   On  the  Shoulders  of  Titans,  esp.  pp.  18-20,  92-93,  98-100,  132-134, 
147-148,  and  170-172. 

37.  Francis  M.  Rogallo,    "Parawings  for  Astronautics,"   luncheon  speech  presented  at  the 
Specialist  Meeting  on  Space  Rendezvous,  Rescue  and  Recovery,  co-sponsored  by  the 
American  Astronautical  Society  and  the  Air  Force  Flight  Test  Center,  Edwards  AFB, 
Calif.,  10-12  Sept.  1963,  p.  1,  copy  in  Ames  Collection,  box  6,  LHA. 

38.  Ibid.,  p.  2. 

39.  Ibid. 

40.  Ibid.,  p.  3. 

41.  Ibid.,  p.  4. 

42.  In  the  Milton  Ames  Collection  in  the  LHA  is  a  folder  labeled   "Flexible  Wing"  with 
numerous  items  relevant  to  the  history  of  Rogallo's  concept.  These  include  several  talks 
and  papers  that  cover  this  period  of  the  parawing's  development  delivered  by  Francis 
Rogallo  to  various  professional  organizations. 

510 


Notes  for  Chapter  11 

43.  Francis  Rogallo,  "Flexible  Wings,"  in  Astronautics  and  Aeronautics  (Aug.  1968):    54. 

44.  "NASA  Honors  Rogallos  with  Large  Cash  Award,"  Langley  Researcher,  2  Aug.  1963, 
p.  1. 

45.  Rogallo,  "Flexible  Wings,"  pp.  52-53. 

46.  "Lifting  bodies"   are  exactly  what  their  names  imply;  that  is,   they  are  bodies  that 
provide  lift  without  the  benefit  of  wings.    In  the  1950s,  engineers  in  the  NACA  and 
the  aerospace  industry  started  thinking  about  such  rounded  half-cone  designs,  which 
resembled  bathtubs  or  one-half  of  a  baked  potato,  as  a  means  of  providing  a  minimum 
yet  significant  amount  of  aerodynamic  lift  for  hypersonic  gliders  or  other  manned  reentry 
vehicles.   They  thought  that  sufficient  lift  could  be  generated  by  such  a  shape  to  allow 
the  pilot  to  maneuver  the  craft  down  from  space  through  the  atmosphere  to  a  graceful 
gliding  landing  on  a  runway.  In  addition,  the  lift  could  be  used  to  slow  the  spacecraft  as  it 
reentered  at  a  high  speed  from  orbital  altitude.  Lifting-body  research  continued  through 
the  1960s  to  the  present  day.    For  an  analysis  of  some  of  this  NACA/NASA  work,  see 
Hallion,  On  the  Frontier,  pp.  147-172  and  339-346. 

47.  Robert  R.  Gilruth,  STG,  Langley  Field,  to  NASA  headquarters,  Attn:  George  M.  Low, 
"Development  program  on  spacecraft  landing  systems,"  6  July  1961,  Apollo  project  files 
(microfiche),  LHA. 

48.  Floyd  L.  Thompson  to  NASA,  Code  RAA,  "Proposed  Research  Authorization  entitled 
'Free-Flight  and  Wind- Tunnel  Tests  of  Guided  Parachutes  as  Recovery  Devices  for  the 
Apollo  Type  Reentry  Vehicle,'"  30  Aug.  1961,  Apollo  project  files  (microfiche),  LHA. 
On  the  testing  of  the  Parasev  at  Edwards,  see  Hallion,  On  the  Frontier,  pp.  137-140. 

49.  Collins,  Liftoff,  pp.  82-83. 

50.  Hallion,  On  the  Frontier,  p.  138. 

51.  Collins,  Liftoff,  p.  82. 

52.  For  an  accurate  narrative  and  analytical  summary  of  the  Apollo  fire,  see  W.  David 
Compton,    Where  No  Man  Has  Gone  Before:    A  History  of  Apollo  Lunar  Exploration 
Missions,  pp.  91-112.  For  an  inaccurate  and  sensational  popular  account  of  the  tragedy, 
see  Murder  on  Pad  34  (New  York:    G.  P.  Putnam's  Sons,  1968)  by  Eric  Bergaust.  There 
is  definitely  some  blame  and  much  responsibility  to  assign  in  the  matter  of  the  Apollo 
fire,  but  Bergaust 's  indictment  is  largely  off  the  mark. 

53.  Robert  C.  Seamans,  Jr.,  memorandum  for  the  Apollo  204  Review  Board,  28  Jan.  1967, 
Apollo  project  files  (microfiche),  LHA. 

54.  Report  of  the  Presidential  Commission  on  the  Space  Shuttle  Challenger  Accident,  5  vols. 
(Washington,  1987).  For  an  excellent  historical  summary  and  analysis  of  the  Challenger 
accident  and  the  Rogers  Commission  investigation,  see  Collins,  Liftoff,  pp.  222-238. 

55.  Robert  C.  Seamans,  Jr.,  interview  with  Walter  Bonney,  the  Pentagon,  Washington,  D.C., 
23  Feb.  1973,  transcript,  p.  19,  OHC,  LHA. 

56.  Report  of  Apollo  204  Review  Board  to  the  Administrator,  National  Aeronautics  and  Space 
Administration,  5  Apr.  1967,  copy  in  the  Langley  Technical  Library,  CN-122577. 

57.  Ibid.,  p.  iii. 

58.  Collins,  Liftoff,  pp.  137-138. 

59.  John  C.  Houbolt  quoted  in  James  Schultz,   Winds  of  Change:    Expanding  the  Frontiers 
of  Flight:   Langley  Research  Center's  75  Years  of  Accomplishment,  NP-130  (Washington, 
1992),  p.  98. 

511 


Notes  for  Chapter  12 

Chapter  12 
The  Cortright  Synthesis 

1.  On  Hans  Mark's  term  as  director  of  NASA  Ames  Research  Center  (1969-1977),  see 
Elizabeth  A.  Muenger,   Searching  the  Horizon:     A  History  of  Ames  Research  Center, 
194 0-1976,  NASA  SP-4304  (Washington,  1985),  pp.  146-168.   For  insights  into  Mark's 
Machiavellian  management  philosophy,  see  Arnold  Levine  and  Hans  Mark,   The  Man- 
agement of  Research  Institutions:    A  Look  at  Government  Laboratories,  NASA  SP-481 
(Washington,  1984). 

2.  "Dr.  Thompson  Receives  Special  Assignment,"   Langley  Researcher,  8  Mar.  1968.  p.  1. 
The  Langley  Researcher  article  was  basically  a  reprint  of  a  press  release  prepared  at 
NASA  headquarters. 

3.  Edgar  M.  Cortright  interviews  with  author,  Yorktown,  Va.,  18  July  1988  and  7  Aug.  1989. 
The  author  wishes  to  thank  Mr.  Cortright  for  the  detailed  and  very  candid  interviews. 
Most  of  the  insights  used  for  this  chapter  come  from  the  18  July  1988  interview.    The 
subsequent  interview  covered  history  pertinent  to  Cortright 's  work  at  NASA  headquarters 
earlier  in  the  1960s. 

4.  Cortright  interview,  18  July  1988,  transcript,  pp.  3-4  and  5-6,  OHC,  LHA. 

5.  Ibid.,  pp.  3-4.  On  the  feeling  that  NASA  Langley  was  a  little  "sleepy"  under  Thompson's 
direction  and  needed  some  rejuvenation,  also  see  pp.  9-10,  22,  48,  and  68  70. 

6.  Cortright  interview,  18  July  1988,  pp.  5-6. 

7.  Quoted  from  "Cortright  Assumes  Duties  as  New  Langley  Director,"  Langley  Researcher, 
17  May  1968,  p.  1. 

8.  Cortright  interview,  18  July  1988,  pp.  12-14. 

9.  Ibid.,  pp.  6-7. 

10.  Ibid.,  pp.  7  and  10. 

11.  Ibid.,  pp.  14-15. 

12.  Ibid.,  pp.  27-28. 

13.  Ibid.,  pp.  69  and  19-20. 

14.  Langley  Research  Center  Announcement  No.  73-70,  "Reorganization  of  Langley  Research 
Center,"  24  Sept.  1970,  p.  1,  copy  in  file  labeled  "Cortright  Reorganization,"  LHA. 

15.  Ibid. 

16.  For  an  overview  of  the  work  of  the  Atmospheric  Sciences  Division,  see  "Atmospheric 
Sciences  Division,   Space  Directorate,   Langley  Research  Center,"    10  Apr.   1992,   un- 
published, copy  available  in  the  Langley  Public  Affairs  Office.    For  a  published  sum- 
mary of  Langley's  achievements  in  the  field  of  atmospheric  sciences,  see  James  Schultz, 
Winds  of  Change:  Expanding  the  Frontiers  of  Flight,  NASA  NP-130  (Washington,  1992), 
pp.  119-121. 

17.  "Reorganization  of  Langley  Research  Center,"    p.   7.     See  also   "Langley  Announced 
Manager  of  Viking  Project  Office,"  Langley  Researcher,  13  Dec.  1968,  p.  1. 

18.  Cortright  interview,  18  July  1988,  pp.  45-48. 

19.  "Reorganization  of  Langley  Research  Center,"  p.  2. 

20.  Cortright  interview,  18  July  1988,  p.  16.   See  also  "Dr.  John  E.  Duberg  Named  Acting 
Associate  Director,"  Langley  Researcher,  19  Apr.  1968,  p.  1.  I  conducted  a  long  interview 


512 


Notes  for  Chapter  12 

with  John  E.  Duberg  that  covers  the  history  of  the  Cortright  reorganization  and  his 
ambivalent  feelings  about  it.  A  transcript  of  this  interview  is  located  in  the  OHC,  LHA. 

21.  Cortright  interview,  18  July  1988,  pp.  17-18. 

22.  Ibid.,  pp.  35  and  41. 

23.  "Loftin  Appointed  to  Special  Assistant  to  Air  Force,"  Langley  Researcher,  28  May  1971, 
p.  1.    For  Loftin's  critique  of  the  Cortright  reorganization,  see  interview  with  author, 
5  Aug.  1989,  Newport  News,  Va.,  transcript,  pp.  72-87,  OHC,  LHA.  For  Cortright's 
version,  see  Cortright  interview,  18  July  1988,  pp.  34-35. 

24.  "R.  E.  Bower  Named  Langley  Director  for  Aeronautics,"  Langley  Researcher,  9  July  1971, 
p.  1.  Also  see  Cortright  interview,  18  July  1988,  p.  35. 

25.  For  additional  thoughts  from  Cortright  on  his  youth  movement,  see  Cortright  interview, 
18  July  1988,  pp.  19-20,  42-43,  and  69. 

26.  Ibid.,  pp.  32-33.  Langley  in  the  1960s  and  before  did  actually  recruit  at  many  universities 
outside  of  the  South.  It  is  true,  however,  that  graduates  of  southern  schools  outnumbered 
the  rest. 

27.  Ibid.,   p.  44.     The  phrase   "sweat-shop  conditions"   is  overly  dramatic  and  implies  a 
callousness  toward  manual  labor  on  the  part  of  Langley  management  that  did  not  exist; 
Floyd  Thompson  and  the  old  NACA  crowd  were  extremely  fond  of  the  technical  support 
staff  and  interacted  with  them  much  more  regularly  than  Cortright  and  his  successors 
ever  did.    Furthermore,  into  the  late  1960s,  several  buildings  at  Langley  that  housed 
researchers  were  not  air-conditioned,  leaving  them  to  suffer  through  the  summer  heat 
and  humidity  as  well.  Nevertheless,  Cortright's  initiative  should  not  be  undervalued.  His 
changes  to  the  shops  made  a  big  difference  to  those  who  worked  with  their  hands  in  the 
center's  various  shops,  plants,  and  hangars. 

28.  Cortright  interview,  18  July  1988,  p.  61. 

29.  Ibid.,  p.  56.  See  also  "Contract  Awarded  to  Build  Visitor  Information  Center,"  Langley 
Researcher,  10  July  1970,  p.  1. 

30.  Cortright  interview,  18  July  1988,  p.  60. 

31.  For  the  complete  history  of  Project  Viking  and  Langley's  management  role,  see  Edward 
C.  and  Linda  Neuman  Ezell,  On  Mars:  Exploration  of  the  Red  Planet,  1958-1978,  NASA 
SP-4212  (Washington,  1984). 

32.  Cortright  interview,  18  July  1988,  pp.  49-53.  On  NASTRAN,  see  Thomas  G.  Butler  and 
Douglas  Michel,  NASTRAN:  A  Summary  of  the  Functions  and  Capabilities  of  the  NASA 
Structural  Analysis  Computer  System,  NASA  SP-260  (Washington,  1971).    ICASE  has 
remained  active  from  its  inception  under  Cortright  to  the  present.   Its  director  in  1993 
was  Mohammed  Y.  Hussaini,  Bldg.  1192,  18  West  Taylor  Street,  NASA  Langley  Research 
Center. 

33.  Cortright  interview,  18  July  1988,  pp.  51-52.    For  information  on  the  early  history  of 
IPAD,  see  Robert  E.  Fulton,  "National  Meeting  to  Review  IPAD  Status  and  Goals,"  in 
Astronautics  and  Aeronautics  18  (July/ Aug.  1980):  49-52. 

34.  Cortright  interview,  18  July  1988,  pp.  52-53. 

35.  "Langley  Buys  Boeing  Aircraft   for  Flight   Research  Program,"    Langley  Researcher, 
30  Mar.  1973,  p.  1.  For  all  the  details  on  the  TCV  program,  see  NASA  Langley  Research 
Center,  Terminal  Configured  Vehicle  Program  Plan  (Hampton,  Va.,  1  Dec.  1973).  For  a 


513 


Notes  for  Chapter  12 

complete  history  of  Langley's  use  of  the  Boeing  737  airplane  in  the  TCV/ATOPS  pro- 
gram, see  Lane  E.  Wallace,  Airborne  Trailblazer:  Two  Decades  with  NASA  Langley's  737 
Flying  Laboratory,  NASA  SP-4216  (Washington,  1994). 

36.  Cortright  interview,  18  July  1988,  pp.  65-66. 

37.  Ibid.,    pp.   54-55.      For  an   insightful   history   of  how   Langley   was   able  to  support 
supersonic  research  in  the  1970s  and  1980s  in  spite  of  the  political  and  budgetary  climate 
working  against  it,  see  F.  Edward  McLean,  Supersonic  Cruise  Technology,  NASA  SP-472 
(Washington,  1985),  esp.  pp.  101-170. 

38.  Cortright  interview,  18  July  1988,  pp.  55-56.  For  information  related  to  the  origins  of  the 
NTF  at  Langley,  see  "Langley  Tested  for  NTF  Site,"  Langley  Researcher,  11  July  1975, 
p.  1,  and  "National  Transonic  Facility  to  be  Dedicated,"  Langley  Researcher,  18  July  1977, 
p.  1.  For  an  excellent  technical  and  historical  overview  of  the  development  of  the  NTF, 
see  Donald  D.  Baals  and  William  R.  Corliss,    Wind  Tunnels  of  NASA,  NASA  SP-440 
(Washington,  1981). 

39.  Macon  C.  Ellis,  "Rough  Thoughts  on  Status  and  Future  of  Research  at  Langley,"  p.  2, 
in  file  labeled  "Thoughts  on  Status  of  Research  at  Langley,"  Macon  C.  Ellis  Collection, 
LHA. 

40.  Ibid. 

41.  Ibid.,  pp.  1-2. 

42.  "Some  Thoughts  on  the  Status  of  Research  at  Langley,  1969,"  p.  1,  in  file  labeled  "Reasons 
for  Accepting  Assoc(iate)  Ch(ief) — H(ypersonic)  V(ehicles)  D(ivision)  in  1970/Case  for 
Sp(ace)  Scien(ces)  Div(ision),"  in  Ellis  Collection,  LHA. 

43.  Ellis  notes  to  "So-called  Minutes  of  10/27/69  Mtg.  of  Sp.  Sci.  &  Tech.  Steer.  Comm.," 
10  Nov.  1969,  in  Ellis  Collection,  LHA. 

44.  "Some  Hindsight  Comments  on  Langley  Management  and  Reorganizations  of  1969-1974," 
p.  1,  in  John  V.  Becker  Collection,  folder  22-1,  Archives  of  Aerospace  Exploration,  Main 
Library,  Virginia  Polytechnic  Institute  and  State  University,  Blacksburg,  Va. 

45.  Ibid.,  pp.  1-2. 

46.  Ibid.,  p.  2. 

47.  Ibid. 

48.  See  "Abstract  of  Remarks  for  Session  7,  Airlie  House  Meeting,  3/13/70,"  in  folder  22-3, 
Archives  of  Aerospace  Exploration,  pp.  1-2,  Virginia  Polytechnic  Institute  and  State 
University. 

49.  Ibid.,  p.  1. 

50.  Ibid.,  pp.  1-2. 

51.  Ibid.,  p.  2. 

52.  Ibid.,  pp.  2-3. 

53.  Ibid. 

54.  Ibid.,  p.  3. 

55.  See  the  marginal  notes  to  Becker's  "Abstract  of  Remarks  for  Session  7,"  p.  3,  for  hand- 
written notes  concerning  the  comments  made  by  others  to  his  presentation.  Cortright's 
views  on  keeping  Langley  alive  mirrored  those  of  his  counterpart  at  NASA  Ames,  Dr.  Hans 
Mark.  In  his  1984  study  of  government  laboratories  (coauthored  by  Arnold  Levine),  Mark 
wrote:    "There  is  a  misconception  that  organizational  self-perpetuation  is  somehow  bad 


514 


Notes  for  Epilogue 

or  even  sinister.  But  this  is  not  so.  If  it  were,  there  would  scarcely  be  a  major  corpo- 
ration or  nonprofit  organization  that  could  outlive  its  original  reason  for  being  and  find 
other  uses  for  its  resources  and  experience"  ( The  Management  of  Research  Institutions, 
p.  82).  The  only  way  for  a  government  laboratory  to  survive  hard  times,  Mark  argued, 
was  for  its  management  to  be  boldly  entrepreneurial.  In  summary,  this  meant  carving 
out  a  unique  research  area  or  "area  of  emphasis,"  attracting  new  programs  to  it,  playing 
hardball  bureaucratic  politics,  and  generally  doing  whatever  was  necessary  to  cultivate 
powerful  clients,  sponsors,  and  constituencies.  In  contrast  to  this  Machiavellian  approach, 
Becker's  view  was  not  that  Langley  had  totally  achieved  the  objective  for  which  it  was 
first  organized,  or  ever  could,  and  thus  could  be  put  to  death,  but  that  it  should  live  on 
to  do  what  it  was  set  up  to  do  and,  in  fact,  did  the  best.  For  him,  it  was  a  quality  of  life 
issue,  so  to  speak.  It  made  no  sense  to  keep  Langley  breathing  artificially  for  just  any 
purpose;  the  country  needed  Langley  to  survive  for  basic  applied  research. 

56.  Kenneth  Scheibel,  "Future  Operation  of  LRC  Assured,"  Newport  News  (Va.)  Daily  Press, 
17  Aug.  1969. 

57.  Routing  slip,  J.  V.  Becker  to  Mr.  Cortright,  19  Aug.  1969,  pp.  1-2,  copy  in  folder  22-1, 
Becker  Collection,  Archives  of  Aerospace  Exploration,  Virginia  Polytechnic  Institute  and 
State  University. 

58.  Ibid.;  Cortright  wrote  his  reply  on  top  of  Becker's  memorandum  with  a  thick  black  pen, 
Archives  of  Aerospace  Exploration,  Virginia  Polytechnic  and  State  University. 

Epilogue 

1.  The  Apollo  program  cost  a  total  of  $25  billion — $21  billion  on  the  first  lunar  landing.  In 
contrast,  in  fiscal  year  1961,  the  U.S.  government  spent  a  total  of  only  $98  billion.  In  1961 
the  entire  NASA  budget  was  still  less  than  $1  billion;  by  1964  it  reached  over  $5  billion 
and  stayed  at  that  level  for  the  next  four  years.  For  data  on  the  funding  history  of  the 
Apollo  program,  see  Linda  Neuman  Ezell,  NASA  Historical  Data  Book,  vol.  2,  Programs 
and  Projects,  1958  1968,  NASA  SP-4012  (Washington,  1988),  pp.  128-132,  and  Ezell, 
NASA  Historical  Data  Book,  vol.  3,  Programs  and  Projects,  1969-1978,  NASA  SP-4012 
(Washington,  1988),  pp.  71-63. 

2.  A  vast  historical  literature  exists  that  treats  the  remarkable  developments  in  American 
history  in  the  late  1960s  and  early  1970s.    For  contrasting  contemporary  insights  into 
these  turbulent  years,   see  W.   L.   O'Neill,    Coming  Apart:     An  Informal  History  of 
America  in  the  1960s  (Chicago:  Quadrangle  Books,  1971),  and  Morris  Dickstein,  Gates 
of  Eden:    American  Culture  in  the  Sixties  (New  York:    Basic  Books,  1977).  For  a  more 
balanced  treatment,  I  recommend  T.  Gitlin,  The  Sixties:    Years  of  Hope,  Days  of  Rage 
(Toronto  and  New  York:    Bantam  Books,  1987).    On  public  policy  in  the  years  of  the 
waning  spaceflight  revolution,  see  A.  J.  Reichley,  Conservatives  in  an  Age  of  Change:   The 
Nixon  and  Ford  Administrations  (Washington:    Brookings  Institution,   1981).    On  the 
policies  of  detente,  see  Raymond  L.  Garthoff,  Detente  and  Confrontation:    American- 
Soviet  Relations  from  Nixon  to  Reagan  (Washington:    Brookings  Institution,  1985). 

3.  Michael  Collins,  Liftoff:    The  Story  of  America's  Adventure  in  Space  (New  York:    Grove 
Press,  1988),  p.  269. 


515 


Notes  for  Epilogue 

4.  For  a  thoughtful  retrospective  on  the  ambitions  of  1969,  both  for  and  against  the  space 
program,  read  John  Tierney's  feature  story,  "Earthly  Worries  Supplant  Euphoria  of  Moon 
Shots,"  The  New  York  Times,  20  July  1994,  sec.  A,  pp.  1  and  11. 

5.  Ibid.    To  revisit  the  days  of  the  "counterculture,"  peruse  Charles  Reich,   The  Greening 
of  America:     How  the    Youth  Revolution  Is   Trying  to  Make  America  Livable  (New 
York:     Random  House,  1971).    For  a  detailed  account  of  the  Nixon  presidency  during 
the  year  1969,  see  Stephen  E.  Ambrose,  Nixon:    The  Triumph  of  a  Politician,  1969-1972 
(New  York  and  London:    Simon  and  Schuster,  1989),  pp.  223-321. 

6.  Norman  Mailer,  Of  a  Fire  on  the  Moon  (New  York:    Signet,  1969),  p.  337. 

7.  Marcus,  Picasso,  and  Mumford  quoted  in  Tierney,  "Earthly  Worries  Supplant  Euphoria 
of  Moon  Shots,"  p.  11. 

8.  Von  Braun  quoted  in  Tierney,   "Earthly  Worries  Supplant  Euphoria  of  Moon  Shots," 
p.  11. 

9.  Howard  McCurdy,  The  Space  Station  Decision:    Incremental  Politics  and  Technological 
Choice  (Baltimore  and  London:    Johns  Hopkins  University  Press,  1990),  pp.  25  26. 

10.  Memorandum  from  Peter  Flanagan  to  Thomas  O.  Paine  and  Robert  Mayo,  6  Jan.  1970, 
in  NASA  HQA;  quoted  in  McCurdy,  The  Space  Station  Decision,  p.  26. 

11.  Richard  M.  Nixon,  memorandum  for  the  Vice-President  et  al.,  13  Feb.  1969,  taken  from 
Space  Task  Group,  The  Post- Apollo  Space  Program:    Directions  for  the  Future,  report  to 
the  president  (Washington,  D.C.:    Executive  Office  of  the  President,  Sept.  1969). 

12.  McCurdy,  The  Space  Station  Decision,  p.  23. 

13.  Richard  M.  Nixon,  "The  Future  of  the  United  States  Space  Program,"   Weekly  Compila- 
tion of  Presidential  Documents,  1  Mar.  1970,  p.  329. 

14.  McCurdy,  The  Space  Station  Decision,  p.  27. 

15.  William  David  Compton  and  Charles  D.  Benson,  Living  and  Working  in  Space:  A  History 
of  Skylab,  NASA  SP-4208  (Washington,  1983),  pp.  52-56. 

16.  Fletcher  quoted  in  McCurdy,  The  Space  Station  Decision,  p.  27. 

17.  McCurdy,  The  Space  Station  Decision,  pp.  22-23. 

18.  "America's  Future  in  Space,"  speech  given  by  Daniel  S.  Goldin,  NASA  administrator, 
at  "  'The  Eagle  Has  Landed'  at  Auburn:    A  25- Year  Retrospective  on  the  First  Lunar 
Landing,"  a  symposium  held  at  the  Auburn  University  Hotel  and  Conference  Center, 
Auburn  University,  Auburn,  Ala.,  27  Oct.  1994. 

19.  Walter  A.  McDougall,  The  Heavens  and  the  Earth:   A  Political  History  of  the  Space  Age 
(New  York:    Basic  Books,  1985),  p.  5. 

20.  Ibid.,  p.  459. 

21.  For  an  analysis  of  President  Reagan's  1984  State  of  the  Union  address  and  its  support  of 
Space  Station  Freedom,  see  McCurdy,  The  Space  Station  Decision,  pp.  177-196.  On  the 
history  of  SDI — or  at  least  that  part  of  the  top-secret  program  that  is  open  to  scholars — 
see  Donald  R.  Baucom,  The  Origins  of  SDI,  1944~1983  (Lawrence:   University  of  Kansas 
Press,  1992). 

22.  Howard  McCurdy,  "The  Graying  of  Space,"  Space  World,  12  Oct.  1988. 

23.  For  an  excellent   profile  of  NASA   Administrator   Daniel   S.   Goldin's  philosophy   for 
"reinventing"  NASA,  see  William  J.  Broad,  "Scientist  at  Work:    Daniel  S.  Goldin:    Bold 
Remodeler  of  a  Drifting  Agency,"  The  New  York  Times,  21  Dec.  1993,  sec.  B,  pp.  5  and  8. 


516 


Notes  for  Epilogue 

For  NASA's  "Vision,  Missions,  and  Goals"  under  Administrator  Goldin,  see  the  "NASA 
Strategic  Plan"  of  May  1994.  A  copy  is  available  from  any  NASA  public  affairs  office. 

24.  Quoted  in  H.  J.  P.  Arnold's  chapter,   "The  Shuttle:     Tragedy  and  Recovery,"  in  Man 
in  Space:    An  Illustrated  History  of  Space/light  (New  York:    Smithmark,  1993),  p.  159. 
For  the  complete  findings  and  recommendations,  see  U.S.  House  of  Representatives,  The 
Future  of  the  U.S.  Space  Program.   Hearings  before  the  Subcommittee  on  Space  Science 
and  Applications  of  the  Committee  on  Science,  Space,  and  Technology,  101st  Cong.,  2d 
sess.,  23  July  1990. 

25.  Edgar  M.  Cortright  quoted  in  James  Schultz,  Winds  of  Change:  Expanding  the  Frontiers 
of  Flight:    Langley  Research  Center's  75  Years  of  Accomplishment,   1917-1992,  NASA 
NP-130  (Washington,  1992),  p.  111. 

26.  For  a  summary  of  Langley's  wind-tunnel  testing  in  support  of  the  Shuttle's  development, 
see  Donald  D.  Baals  and  William  R.  Corliss,    Wind  Tunnels  of  NASA,  NASA  SP-440 
(Washington,  1981),  pp.  117-121.    A.  M.  Whitnah  and  E.  R.  Hillje  provide  a  detailed 
engineering  report  on  Shuttle  wind-tunnel  testing  in  Space  Shuttle  Wind  Tunnel  Testing 
Program  Summary,    NASA   RP-1125   (Washington,    1984).      For  a  discussion  of  the 
development  of  the  special  tiles  used  to  protect  the  Shuttle  during  reentry,  see  Paul 
Cooper  and  Paul  F.  Holloway,  "The  Shuttle  Tile  Story,"  Astronautics  &  Aeronautics  19 
(Jan.  1981):    24-36.  Cooper  and  Holloway  were  Langley  researchers  directly  involved  in 
the  Shuttle  tile  program.    In  1991,  Holloway  became  the  director  of  Langley  Research 
Center,  succeeding  Richard  H.  Petersen.    It  is  also  worth  noting  that  former  Langley 
engineer  Max  Faget  played  a  critical  role  at  the  NASA  center  in  Houston  in  the  conceptual 
design  of  the  Space  Shuttle.     For  an  important  article  that  looks  at  early  plans  for 
the  Space  Shuttle,  see  Faget's  "Space  Shuttle:    A  New  Configuration,"  Astronautics  & 
Aeronautics  8  (Jan.  1970):    52-61.  Starting  in  June  1973,  the  journal  Spaceflight  began 
running  a  series  of  articles  by  David  Baker  on  the  design  evolution  of  the  Space  Shuttle; 
the  last  in  the  series  appeared  in  Mar.  1978. 

Langley  remained  deeply  involved  with  the  STS  program  even  after  the  maiden  flight 
of  Columbia  in  1981.  The  center  developed  simulations  to  solve  problems  in  the  orbiter's 
flight  control  and  guidance  systems  and  conducted  additional  landing  tests  on  tires  and 
brake  systems.  After  the  Challenger  accident  in  Jan.  1986,  Langley  pitched  in  by  helping 
to  redesign  the  solid  rocket  boosters  and  develop  new  crew  emergency  escape  systems. 

27.  On  the  potential  of  the  aerobrake  in  the  mid-1980s,  see  Thomas  O.  Paine  et  al.,  Pioneering 
the  Space  Frontier:    The  Report  of  the  National  Commission  on  Space  (Toronto  and  New 
York:    Bantam  Books,  1986),  pp.  121-122.    For  a  fascinating  science-fiction  account  of 
the  use  of  aerobraking  to  accomplish  a  Mars  orbital  insertion  (or  MOI)  for  humankind's 
first  permanent  colonizing  of  the  Red  Planet,  I  highly  recommend  Kim  Stanley  Robinson, 
Red  Mars  (New  York:    Bantam  Spectra  Books,  1993),  pp.  70-91. 

28.  Several  ambitious  proposals,  NASA  and  non-NASA,  surfaced  in  the  1970s  to  continue 
lifting-body  research,  but  they  failed  to  rejuvenate  the  program.  With  the  Shuttle  well 
into  its  design  phase  by  the  mid-1970s,  NASA  could  not  afford  to  vigorously  pursue  what 
amounted  to  a  rival  technology.    For  an  analytical  look  into  the  history  of  the  lifting- 
body  concept  in  the  United  States,  see  Richard  P.  Hallion,   On  the  Frontier:     Flight 
Research  at  Dryden,  1946-1981,  NASA  SP-4303  (Washington,  1984),  pp.  161-172.   See 


517 


Notes  for  Epilogue 

also  Stephan  Wilkinson,  "The  Legacy  of  the  Lifting  Body,"  Air  &  Space  Smithsonian  6 
(Apr./May  1991):    62. 

On  ambitions  for  the  HL-20  by  the  early  1990s,  see  Theodore  A.  Talay  and  Howard 
H.  Stone,  "The  Personal  Launch  System — A  Lifting  Body  Approach,"  a  paper  presented 
at  the  42d  Congress  of  the  International  Astronautical  Federation,  5-11  Oct.  1991, 
Montreal,  Canada.  The  IAF  published  this  paper  as  IAF-91-202. 

29.  Donald  P.  Hearth  and  Albert  E.  Preyss,  "Hypersonic  Technology:    Approach  to  an  Ex- 
panded Program,"  Astronautics  &  Aeronautics  (Dec.  1976):    pp.  20-37.   Hearth  served 
as  director  of  Langley  Research  Center  from  1975  to  1985.  On  the  hopes  for  the  scram- 
jet,  see  The  National  Commission  on  Space,  Paine  et  al.,  Pioneering  the  Space  Frontier, 
pp.  101-103.    The  technical  literature  on  the  scramjet  is  extensive,  and  much  of  it  is 
classified.    For  summaries  of  Langley's  work,  see  O.  A.  Buchanon,   "Development  and 
Fabrication  of  Structural  Components  for  a  Scramjet  Engine"  (Hampton,  Va.:    NASA 
Langley  Research  Center,  1990),  microfiche;  "Shock  Tunnel  Studies  of  Scramjet  Phenom- 
ena" (Hampton,  Va.:    NASA  Langley  Research  Center,  1988),  microfiche;  and  Stanley 
W.  Candebo,  "Scramjet  Test  Bridges  Hypersonic  Efforts,"  Aviation  Week  &  Space  Tech- 
nology 140  (28  Mar.  1994):    52-54.    The  Candebo  article  reports  on  planned  tests  of  a 
large-scale  ramjet/scramjet  power  plant  in  Langley's  8-Foot  High- Temperature  Tunnel 
that  will  allow  NASP  investigators  to  determine,  for  the  first  time,  how  scale  effects  have 
influenced  the  design  calculations  used  in  scramjet  R&D. 

30.  On  the  status  of  NASP  in  early  1994,  see  Stanley  W.  Candebo,  "Redirected  NASP  Pro- 
gram to  Focus  on  Technology,"  Aviation  Week  &  Space  Technology  140  (10  Jan.  1994):  29. 
See  also  Candebo's  story  from  the  previous  summer,  "NASP  Canceled,  Program  Redi- 
rected," Aviation  Week  &  Space  Technology  138  (14  June  1993):    32. 

31.  Baals  and  Corliss,    Wind   Tunnels  of  NASA,   pp.   96  and   133  134.      See  also  Keith 
F.  Mordoff,  "New,  Wind  Tunnel  Improves  Accuracy,"  Aviation  Week  &  Space  Technology 
119  (19  Dec.  1983):    69. 

32.  For  a  popular  treatment  of  the  NTF's  design  features  and  promises  for  important  new 
data,  see   "Mighty  Wind  Roars,"    Popular  Mechanics  162  (Apr.   1985):  65.     There  is 
a  rather  vast  technical  literature  on  the  NTF,  mostly  NASA  reports,  all  of  which  are 
available  in  the  Langley  Technical  Library.  A  history  of  the  NTF  would  make  a  fascinating 
doctoral  dissertation  in  contemporary  technology. 

33.  Baals  and  Corliss,  Wind  Tunnels  of  NASA,  p.  134. 

34.  Mark  Di  Vincenzo,   "Heaven's  Call  Ever  Present;  NASA  Langley  to  Continue  Role  in 
Space  Program,"    Newport  News-Hampton   (Va.)   Daily  Press,   20  July   1994,   sec.  A, 
pp.  1-2. 

35.  On  the  ambitions  for  the  Mission  to  Planet  Earth,  see  Paine  et  al.,  Pioneering  the  Space 
Frontier,  pp.  30-33. 

36.  Di  Vincenzo,  "Heaven's  Call  Ever  Present,"  p.  1. 


518 


Index 


Abbott,  Ira  H.,  19,  199,  251,  29ill. 
ABL.    See  Allegheny  Ballistics  Labora- 
tory. 

Adamson,  David,  129,  130,  142-143,  223 
Administration  Directorate,  404 
Advanced  Man  in  Space  (AMIS)  project, 

272-274 
Advanced  Research  Projects  Agency 

(ARPA),  51,  52,  55 
Advanced  Transport  Operating  System 

(ATOPS),  415 
Advisory  Group  for  Aeronautical  Research 

and  Development  (AGARD),  267 
aerobrake,  435 

Aerojet  General  Corporation,  200 
aeronautics, 

decline  at  Langley,  96-102,  97ill. 
present-day  at  Langley,  438 
the  shift  to  space,  17-22,  451 
Aeronautics  Directorate,  403,  404 
Aero-Physics  Division,  89,  127 
Aero-Space  Mechanics  Division,  89-90 
Agena,  370 
Agena-Class  Lunar  Orbiter  Project 

(ACLOPS),  327 

Agnew's  Space  Task  Group,  430 
Aircraft  Configuration  Branch,  177 
Aircraft  Engine  Research  Laboratory,  7. 

See  also  Lewis  Research  Center. 
Aircraft  Manufacturers'  Conferences,  7, 

Sill. 
air-density  measurement,  156,  158-162, 

166,  191 
Air  Force  Scientific  Advisory  Board,  235, 

236 

Air  Force  Tactical  Air  Command,  60 
Air  Force,  U.S.,  20,  57,  165,  170,  199,  200, 

214,  289,  327,  338,  453 
Airlie  House,  418,  422,  423,  424 
Air  Scoop,  11,  169 
air  traffic,  improvement,  415 


Aldrin,  Edwin  E.,  Jr.,  355,  428 

Alfven,  Hannes,  123 

Algol  motor,  200,  207,  211 

Allario,  Frank,  151 

Allegheny  Ballistics  Laboratory  (ABL), 

X248,  184,  187,  200,  201,  184tH. 

X254,  201 

X259,  201 

Allen,  H.  Julian,  53,  393 
Altair  motor,  201,  207 
American  Telephone  and  Telegraph 

(AT&T),  176,  194 
Ames  Aeronautical  Laboratory,  7 
Ames,  Joseph  S.,  7,  Sill. 
Ames  Research  Center,  7,  32,  232,  393 
Ampere,  Andre-Marie,  135 
Analysis  and  Computation  Division 

(ACD),  86,  111,  414 
Antares  motor,  201,  209 
Apollo  11,  268,  346,  350,  391,  429,  495 
Apollo  13,  238,  389 
Apollo  204  fire,  373,  387-391,  455 
Apollo  Applications  Program,  306-307 
Apollo  Configuration  Control  Board,  389 
Apollo  Extension  System,  270,  305-306 
Apollo  Orbital  Research  Laboratory 

(AORL),  301 

Apollo  project,  77,  98,  100,  113,  141,  245, 
253,  369,  370,  371,  423 

command  module,  237,  362-366,  364z//. 

flight  plan,  237 

funding,  358 

Langley  contributions,  366-369,  368HI. 
See  also  simulation  research  and 
Apollo  204  fire. 

management  of,  356 

mission  modes.  See  lunar  landing 
mission  modes. 

reentry,  380-381,  385,  386 

scope  of  effort,  355 
Apollo  Spacecraft  Program  Office,  366 


519 


Spaceflight  Revolution 


Apollo  Technical  Liaison  Plan,  233 
Applied  Materials  and  Physics  Division, 

89,  91,  367,  422 
arc-jet,  133 
Armstrong,  Neil  A.,  41,  268,  346,  350,  355, 

378-379,  428,  392z//. 
Army  Ballistic  Missile  Agency  (ABMA), 

xxvii,  34,  42,  102 
Army,  U.S.,  22 
Arnold  Engineering  Development  Center, 

60,  133 

Arnold,  Henry,  7 

astronauts,  Mercury,  39-45,  77,  456 
Atlas-Agena,  293,  313,  321,  344 
Atlas-Centaur,  344 
Atlas-Mercury  vehicle,  47 
Atlas  rocket,  38,  46,  50,  53,  54,  59,  236, 

367 

atmospheric  sciences,  404 
atomic  bomb,  113,  135,  147,  171 
Atomic  Energy  Commission,  12,  132,  147, 

289 

AT&T,  176,  194 
Augustine,  Norman,  433 
aurora  borealis,  122,  145,  146 
avionics,  415 

A.V.  Roe  (AVRO),  248,  455 
Back  River,  45,  72,  94,  62ill. 
Bainbridge,  William  Sims,  xxx-xxxi 
Barium  Cloud  Experiment,  142-147 
Bay  of  Pigs,  249 
Becker,  John  V.,  19,  31,  89,  98,  116,  127, 

132,  150,  151,  232,  420-425,  434, 


Bell  aircraft, 
P-59,  ISill. 
X-l,  7,  202 
X-14A,  379 
Bell  Communications  (BellComm),  108, 

325,  350 
Bell  Telephone  Laboratories,  65,  163,  176, 

178,  186,  259 
Bendix  Corporation,  65 
Bennett,  Willard  H.,  148 
Berglund,  Rene  A.,  281,  283,  295,  297, 

278  ill. 
B-58,  20 
Biermann,  Ludwig,  143-144 


Biggins,  Virginia,  456 

Big  Joe,  46-47,  59,  77 

Big  Shot,  113,  193-194 

Bikle,  Paul  F.,  31 

Billingsley,  Henry  E.,  30 

Bird,  John  D.,  230,  232,  233,  242,  246,  257, 

229UI. 

Bisplinghoff,  Raymond  L.,  97,  Will. 
black  hole,  216 
Boeing  aircraft, 

P-26A  Peashooter,  6ill. 
737-100,  415,  416z//. 
757,  416 
767,  416 
Boeing  Company,  294,  295-297,  302, 

313-314,  327 
Lunar  Orbiter,  328-332,  334-336, 

340-342,  344-346 

Terminal  Configured  Vehicle  and  Avion- 
ics program  (TCVA),  415-416, 

416HI. 

Bomarc  Missile  program,  295,  327 
Bond,  Aleck  C.,  257 
Bonney,  Walter,  12,  31UI. 
Borman,  Frank,  387 
Bostick,  Winston  H.,  149 
Bower,  Robert,  411 

Boyer,  William  J.,  57,  65,  67,  324,  332 
Braun,  Wernher  von.     See  von  Braun, 

Wernher. 

Brayton  cycle,  288 
Brayton,  George  B.,  288 
Bressette,  Walter  E.,  167,  169,  171, 

174-175,  186,  191,  175UL,  183UI. 
Brewer,  Gerald,  323 
Brissendon,  Rcy  R.,  371 
British  National  Physical  Laboratory,  17 
Brockman,  Philip,  134i//. 
Brookings  Institution,  10 
Brooks,  Overton,  285,  286 
Brown,  C.  T.,  Jr.,  202 
Brown,  Clinton  E.,  42,  86,  158,  222, 

223-224,  228,  232,  233,  236-237, 

241,  242,  246,  276,  320,  327,  88UL, 

227UI. 

Brummer,  Edmund  A.,  323 
Buckley,  Edmond  C.,  57,  68,  69 
"Buck  Rogers",  17 


520 


Index 


budget, 

Langley,  423 

space  program,  428 
Buglia,  James  J.,  53 
Bulkeley,  Rip,  447 
Bullpup  missile,  210 
Bureau  of  the  Budget,  10,  104 
Burgess,  George  K.,  Sill. 
Burlock,  Joseph,  140 
Burns  and  Roe,  65 

Busemann,  Adolf,  129-130,  135,  139,  276 
Bush,  George,  433 
Bush,  Vannevar,  7,  dill. 
Butler,  Sherwood,  112,  325 
Butler,  T.  Melvin,  404,  90tH.,  279z//., 

39631 

Cable  News  Network  (CNN),  195 
Cambridge  Research  Center,  164 
Cape  Canaveral,  xxvii,  xxix,  64,  66,  102, 

170,  179,  187,  188,  189,  293,  367 
Cape  Kennedy,  356 
Carpathian  mountains,  348?//. 
Carpenter,  Malcolm  Scott,  40,  41,  42,  76, 

40tH. 

Case  Institute  of  Technology,  11,  13 
Castor,  201 
Castro,  Fidel,  249 
Centaur  rocket,  236,  283,  313,  351 
Center  Development  Directorate,  404 
Central  Intelligence  Agency  (CIA),  172, 

249 

centrifuge,  42,  45 

Centro  Italiano  Ricerche  Aerospaziali,  217 
CF-105  Arrow,  455 
Chaffee,  Roger  B.,  387,  455,  390iH. 
Chamberlin,  James,  235,  248,  257 
Champine,  Robert  A.,  41 
Chance  Vought  Astronautics,  228,  276 
Chance  Vought  Corporation,  201,  203,  228, 
260 

F8U-3,  27 

Chapman,  Sydney,  122 
Cherokee,  198 
CIA,  172,  249 
Clarke,  Arthur  C.,  162-163,  175,  176,  194, 

195,  271 

Clear  Lake,  Texas,  74 
Clemmons,  Dewey  L.,  Jr.,  387 


Clinton,  Bill,  433 

cockpit  development,  415 

Coffee,  Claude  W.,  Jr.,  175 

cold  war,  xxviii,  83,  172,  427,  433 

Colliers,  271-272,  381-383 

Collins,  Michael,  370,  391,  428 

comets,  143,  144,  145 

command  module  (CM),  237,  362-366, 

364iU. 
committees  on  lunar  exploration,  table, 

225 
Communications  Satellite  Corporation 

(ComSatCorp),  194,  195,  200 
communications  satellites  (comsats),  480, 

481 

communications  system,  global,  189 
computers,  111,  413-416,  lllill 
Congress,  51,  430 

Apollo,  392 

appropriations,  437 

House  Independent  Offices  Appropria- 
tions Committee,  74 

House  Select  Committee  on  Science  and 
Astronautics,  176,  285 

House  Space  Committee,  74,  451 

NACA,  2,  9,  10,  23 

NASA  confirmation  hearings,  15 

personnel  authorizations,  104,  106,  109 

space  station,  306 

supersonic  transport  (SST),  101,  416 
contracting, 

for  Langley  services,  108-112 

Langley  philosophy,  82-84 

Lunar  Orbiter,  325-326,  33 lill. 

Manned  Orbiting  Research  Laboratory 
(MORL),  294-302,  307 

Mercury,  59,  65 

Scout,  203-205,  212-215 
Convair  aircraft,  93 

B-58  delta-winged  bomber,  20 

SF2Y-1  Sea  Dart,  93 
Convair  Astronautics,  170 
Cooper,  L.  Gordon,  Jr.,  41,  42,  77,  194, 

517,  40z//. 

Copernicus  crater,  348UL 
Coriolis  effects,  277,  290 
Corneliussen,  Sarah  and  Steve,  109 
Cortright,  Edgar  M.,  Jr.,  150,  335,  389, 
434,  451,  398z//.,  399UL,  4Q9HI. 


521 


Space/light  Revolution 


Cortright,  Edgar  M.,  Jr.  (continued): 

appointed  Langley  director,  393  397 

critique  of,  418-425 

development  of  electronics,  413-418 

Lunar  Orbiter,  319,  320,  325,  326 

1970  reorganization,  89,  90,  92,  152, 
398-407,  421-422 

public  relations,  412-413 

staff,  408-412 

Crabill,  Norman  L.,  153-156,  181,  185, 
189,  195-196,  324,  334,  338,  339, 
340-341,  349,  351,  183UI. 
Cronkite,  Walter,  377,  378UI. 
Crossfield,  Scott  A.,  41 
Crowley,  John  W.,  166,  176-177,  199 
Curtiss- Wright  Corporation,  81 
Davies,  Merton  E.,  328 
Dawson,  John  R.,  302 
Deep  Space  Network,  328 
DeFrance,  Smith  J.,  31,  36-37 
Department  of  Defense  (DOD),  41,  200, 
216,  255,  297-298,  301,  327,  328, 
338,  437,  453 
detente,  428 
D-558,  7 

diffusion  inhibitor,  149 
Diggs,  Charlie,  137z//. 
Discoverer,  28 
Do'dgen,  John  A.,  232 
Dolan,  Thomas  E.,  228,  237,  260 
Donely,  Philip,  89 

Donlan,  Charles  J.,  19,  52,  63,  86,  458, 
264 ill.,  361UL,  398UI. 

Langley  appointments,  99-100,  394-395, 
397,  409,  421 

Lunar  Orbiter,  316,  317,  317UL 

Manned  Spacecraft  Center,  358,  359, 
360,  363,  365 

Mercury,  41 

NACA  transition  to  NASA,  2 

1959  inspection,  34 

rendezvous,  241,  255 

space  station,  294-295,  305 

Space  Task  Group  (STG),  56,  57-58,  73, 

74,  75 

Doolittle,  James  H.,  8,  12,  451 
Douglas  aircraft, 

D-558,  7 


Douglas  Aircraft  Company,  108,  178,  231, 

294,  295-302 

Draley,  Eugene  C.,  19,  89,  91,  118,  177, 
223,  224,  327,  332,  367,  404,  408, 
88i//.,  Will.,  398UI. 
Dryden  Flight  Research  Center,  7 
Dryden,  Hugh  L.,  10,  31,  52,  70,  71,  75, 

76,  384,  400,  451,  452,  Hi//, 
appointment  to  NASA,  12 
Echo,  166,  176 
NACA  appropriations,  23 
1959  inspection,  30,  34 
Space  Task  Group,  73 
Duberg,  John  E.,  91,  367,  408,  398i//., 

400& 

Durand,  William  F.,  7,  8ill. 
Dynamic  Loads  Division,  89,  91,  93 
Dyna-Soar  project,  20,  132,  200,  273,  327, 

453,  9GHL 

Eagle,  346,  350,  356,  495 
earth, 

first  deep  space  photo,  344-346, 
506-507,  ML,  345«H.,  352UI. 
earth-orbit  rendezvous  (EOR).    See  lunar 

landing  mission  modes. 
East  Area,  45,  72,  94 
Eastman  Kodak,  328,  342,  506 
Echo  1,  188-191,  192,  193,  277 
Echo  1  Passive  Communication  Satellite 

Project,  154 
Echo  2,  189,  193-194 
Echo  project,  278,  280,  321 

air-density  measurement,  156,  158-162, 

166,  191 

Beacon  satellite,  170-176,  174z//. 
construction  material,  160,  167-169 
container,  180,  181  ill. 
design,  158-162 
folding  the  balloon,  187-188 
Goddard,  177-179 
inflation,  154,  160,  169,  180-184,  186, 

155t//.,  182UL,  183UL 
initial  test,  153-156 
management,  112-113 
mapping,  190 
micrometeorites,  188,  189 
100-foot  satelloon,  177-185 
proposal  for,  164-166 
Sub-Satellite,  166-170,  168tM.,  171UI. 


522 


Index 


Echo  project  (continued}: 

Task  Group,  177,  178,  181,  186,  187, 

188,  195,  183iH. 
telecommunications,  global  system, 

162-164 
tracking,  161 

yo-yo  de-spin  system,  184-185,  186 
Edwards  Air  Force  Base  (AFB),  7,  63. 
See  also  Dryden  Flight  Research 
Center. 

Edwards,  Howard  B.,  233 
Eggers,  Alfred  J.,  53 
Eggleston,  John  M.,  232,  233 
E.  I.  du  Pont  de  Nemours  &  Co.,  167,  292 
Eisenhower,  D wight  D.,  xxvi,  xxvii,  173, 

ML 

administration,  198 
creation  of  NASA,  2,  12,  449 
Echo  1,  188 

Mercury  astronauts,  41 
NASA  personnel,  104 
Electronics  Directorate,  403,  404 
Ellis,  Macon  C.,  Jr.,  116,  125,  126-129, 
132,  140,  142,  150,  151,  152, 
418-420,  422,  434,  129UI. 
engineer-in-charge,  10 
English,  Roland  D.,  198,  202,  210,  211, 

213-214,  216 
"Enos",  41 

Environmental  and  Space  Sciences  Divi- 
sion, 404,  420,  422 

Erectable  Torus  Manned  Space  Labora- 
tory, 277-281,  279UI.,  280UI. 
European  Space  Agency,  217 
European  Space  Research  Organization 

(ESRO),  144,  146 
Evans,  John  S.,  140 
Explorer,  123 
Explorer  1,  173 
Explorer  9,  191 
Explorer  10,  125 
Explorer  19,  191 
Explorer  24,  191,  192i//. 
Explorer  39,  191 
F8U-3,  27 
F-101A,  17 
F-104  Starfighter,  20 
F-111A,  IQOill. 
F-111B,  lOOtil. 


Faget,  Maxime,  52,  226,  387,  451,  458,  517 

Mattson,  Axel  T.,  360 

Mercury  capsule,  43,  47,  53-54,  44^/. 

rendezvous,  235,  236-237,  241,  242,  255, 
260,  261-262 

Scout,  197 

Fairburn,  Robert,  323 
Fechet,  James  E.,  Sill. 
Federal  Aviation  Administration  (FAA), 

415 
Federal  Ministry  for  Scientific  Research, 

West  Germany,  145 
Feix,  Marc,  130,  131  iH. 
Fields,  Edison  M.,  37 
Fire  project,  91,  112,  216,  321,  322,  332, 

367-369,  367tH. 
Fleming  Committee,  250 
Fleming,  William  A.,  250 
Fletcher,  James,  430 
Flexi-Kite,  381 

Flight  Instrumentation  Division,  91 
Flight  Mechanics  and  Technology  Division, 

90,  98 

Flight  Reentry  Programs  Office,  91 
Flight  Research  Center,  Edwards  Air  Force 

Base,  232,  378-379,  385,  386 
flying  boats,  94 
Foelsche,  Trutz,  330 
France,  217 
Friendship  7,  260 
Fuhrmeister,  Paul  F.,  86,  91 
Full-Scale  Research  Division,  90,  93,  94, 

98,  380 

Gagarin,  Yuri,  249 
Gapcynski,  John  P.,  230 
Gardner,  William  N.,  294,  303-305,  308, 

3QOUL,  30SUI. 
Garland,  Benjamin  J.,  53 
Garrick,  Isadore  E.,  89,  267 
Gas  Dynamics  Laboratory,  122,  126,  129 
Geer,  E.  Barton,  387 
Geggie,  Jean,  24 

Gemini  project,  232,  257,  293,  301,  369, 
370,  371,  385 

reentry,  380-381,  385 
General  Dynamics  aircraft, 

F-111A,  lOOill. 

F-111B,  IQOm. 
General  Dynamics  Corporation,  305 


523 


Space/light  Revolution 


General  Electric  (G.E.),  108,  176 
General  Mills,  181,  184 
General  Motors,  259 

Geophysics  and  Astronomy  Division,  193 
Germany,  217 
Gibbons,  Howard,  155 
Gilkey,  John  E.,  37 
Gillis,  Clarence  L.,  177 
Gilmore,  William  E.,  Sill. 
Gilruth,  Robert  R.,  19,  31,  37,  43,  50,  202, 
222,  379,  451,  38UL,  79UL,  361i//. 

Echo,  165 

lunar-orbit  rendezvous,  248,  265 

Marshall  Space  Flight  Center,  360,  366 

Mercury  planning,  51-55 

rendezvous,  242-245,  257,  260,  261,  262, 
263 

Space  Task  Group's  move  to  Texas, 
76-77,  80,  79^/. 

Space  Task  Group  staffing,  55-57,  59, 
60-63,  65,  69,  72,  73-74 

Sputnik,  172 
Girouard,  Robert,  323 
Glenn,  Annie  Castor,  46z//.,  79ill. 
Glenn,  John  H.,  Jr.,  40,  42,  43,  76,  77,  222, 

260,  4QUL,  45^/.,  46^.,  79ill. 
Glennan,  T.  Keith,  31,  81,  118,  236,  452, 
Will.,  Uill,  29UL,  31ill. 

appointment  to  NASA,  11-15 

"Message  to  Employees",  12-15 

NASA  personnel,  104-106 

1959  inspection,  30 

Space  Task  Group,  55 
global  telecommunication  system,  175-177 
Goddard  Space  Flight  Center,  65,  194, 
232,  395 

Echo,  177-179 

1959  inspection,  32,  34-37,  36ill 

Scout,  217 

Space  Task  Group,  72-74 
Goett  Committee,  226,  232,  272-274 
Goett,  Harry  J.,  31,  74,  179,  226 
Gold,  Thomas,  311 
Goldin,  Daniel,  433,  438 
Golovin  Committee,  255-257 
Golovin,  Nicholas  E.,  255 
Goodyear  Aerospace  Corporation,  191 
Goodyear  Aircraft  Corporation,  276,  277, 
278,  285,  385 


Graham,  John  B.,  323,  324 

Graves,  G.  Barry,  Jr.,  66,  67,  67ill. 

gravity,  artificial,  281,  295,  297 

Green,  Milt,  203,  215 

Griffith,  Leigh  M.,  394 

Grissom,  Virgil  L,  42,  222,  385,  387,  455, 

40UL,  390UI. 
Grosse,  Aristid  V.,  140 
Group  1,  86,  91,  367,  369 
Group  2,  86,  89,  367 
Group  3,  86,  89,  367 
G.  T.  Schjeldahl  Company,  184,  187 
Guggenheim,  Harry  F.,  7 
Guy-Lussac  Promontory,  348z//. 
Hall,  Edwin  H.,  138 
Hall,  Eldon  W.,  255 
Hall,  Harvey,  255 
Hall,  James  R.,  198,  202,  203,  205, 

206-207,  209,  210,  212,  215,  219, 

203UL 

Hall,  R.  Cargill,  317-319 
Hallion,  Richard,  378-379,  385-386 
"Ham",  41 

Hammack,  Jerome,  37 
Harbridge  House,  118 
Harrison,  Albertis  S.,  77 
Harvard  College  Observatory,  164 
Hasel,  Lowell  E.,  233 
Hearth,  Donald  P.,  417,  435,  437 
Heaton  Committee,  253,  255,  258 
Heaton,  Donald  H.,  253 
heat  shield,  53,  362,  369 
Helberg,  Robert  J.,  335,  340,  342,  344, 

345,  351 

Heldenfels,  Richard  R.,  89,  233,  414 
Heos  .7,  146 

Hercules  Powder  Company,  201 
Hess,  Robert  V.,  129,  131,  138,  139-140, 

144,  151,  139UI. 
Hewes,  Donald,  373,  374,  375,  377,  36H/L 

374UI. 
High-Speed  Flight  Station,  7.  See  also 

Dryden  Flight  Research  Center. 
High-Speed  Hydrodynamics  Tank,  93 
Hilburn,  Earl  D.,  81,  104,  106 
Hill,  Paul  R.,  223,  233,  275,  277,  294, 

275UI. 
HL-20,  435 


524 


Index 


Hodge,  John  D.,  37,  455 
Hoffer,  Eric,  xxix 
Holloway,  517 

Holmes,  Brainerd,  258-259,  260,  263 
Honest  John,  198 
Hook,  W.  Ray,  307 
Hoover,  Herbert,  23 
Horrocks,  Edith  R.,  202 
Horton,  Elmer  A.,  37 
Houbolt,  John  C.,  86,  222,  224,  230-231, 
275,  356,  361-362,  391-392, 
495-496,  244tH.,  247 ill.,  266i//. 
letters  to  Robert  Seamans,  249-251, 

258-260 
lunar  landing  mode  decision,  262  263, 

265-268 
lunar-orbit  rendezvous  (LOR),  233-237, 

238,  239-247,  252,  253-255,  257 
rendezvous  committees,  231-233 
Hubbard,  Harvey,  17 
Huber,  Paul  W.,  129,  140,  152,  131UI. 
Hughes  Aircraft  Company,  194,  319,  327, 

328,  330,  331,  335 
Hunsaker,  Jerome  C.,  7,  9 
Huss,  Carl  R.,  37,  57 
Hydrodynamics  Division,  60,  93 
Hypersonic  Ramjet  Experiment,  112 
hypersonics,  158,  432,  435-436.  See  also 

Dyna-Soar  project, 
boost  glider,  xxviii,  20,  54,  273 
tunnels,  133 

Hypersonic  Vehicles  Division,  420 
ICBM.    See  intercontinental  ballistic 

missiles. 

incinerator,  Hampton,  413 
Inspection,  1959,  51,  52 

exhibits,  27-28,  32-33,  35-38,  39-40,  42, 
46-47,  126,  135,  29tH.,  31tM.,  32i«., 
36tH. 

guests,  31-32 
planning,  30,  33  37 
space  station,  274 
Space  Task  Group,  37-38,  40-45 
Inspections,  NAG  A,  28-30,  33-37 
Institute  for  Computer  Applications  in 

Science  and  Engineering  (ICASE), 
414 

Institute  of  Aeronautical  Sciences  (IAS), 
383,  94 


Instrument  Research  Division  (IRD),  86, 

91 
Integrated  Program  for  Aerospace- Vehicle 

Design  (IPAD),  414-415 
Integrative  Life  Support  System  (ILSS), 

305,  304iH. 
intercontinental  ballistic  missiles  (ICBM), 

132,  140,  200 
International  Business  Machines  (IBM) 

Corporation,  65 
International  Geophysical  Year  (IGY),  157, 

163,  198 
International  Telecommunications  Satellite 

Consortium  (Intelsat),  195 
inventor's  award,  384,  495 
ion  rockets,  132,  150 
Italy,  217 

Jacobs,  Eastman  N.,  149 
Jacobs,  Ken,  203,  214 
Jaffe,  Leonard,  177,  178 
Jalufka,  Nelson,  148 
James,  Robert,  185 
Jastrow,  Robert,  224 
Javelin  rocket,  145 
Jet  Propulsion,  163 
Jet  Propulsion  Laboratory  (JPL),  178, 

224,  232,  255,  312-313,  315,  316, 

319,  321,  334,  335,  351 
1959  inspection,  32,  34,  35-37 
Johnson,  Caldwell, 

lunar-orbit  rendezvous,  235 
NACA,  22-23,  24 
Space  Task  Group,  52,  53,  75,  458 
Johnson,  E.  Townsend,  Jr.,  411 
Johnson,  Lyndon  B.,  10,  15,  74,  172,  249 
Johnson,  Roy,  51,  55 
Jones,  Robert  T.,  139 
Journal  of  the  American  Rocket  Society, 

163 

Juno  II,  173 

Jupiter  C,  xxvi,  xxvii,  173,  198,  199 
Jupiter  Senior,  198,  200 
Kantrowitz,  Arthur,  149 
Karth,  Joseph  E.,  286 
KatzorT,  Samuel,  224,  327 
Kavanau,  Lawrence  L.,  255 
Kennedy,  John  F., 
assassination,  194 


525 


Space/light  Revolution 


Kennedy,  John  F.  (continued): 

lunar  landing  commitment,  xxviii,  xxix, 
77,  106,  221-222,  249,  286,  355,  427, 
428 

lunar-orbit  rendezvous,  265 

management  of  NASA,  106 

New  Frontier,  83 

Kennedy  Space  Center,  102,  356,  389 
Kepler  crater,  31SUI. 
Kilgore,  Edwin,  169,  187-188,  IS3UI. 
Killian,  James  R.,  Jr.,  51,  193,  449 
Kimpton,  Lawrence,  118 
King,  Paul,  IGill. 
Kinzler,  Jack,  47,  57 
Kistiakowsky,  George,  51 
Kloman,  Erasmus  H.,  350-351 
Koelle,  Hermann,  265 
Kondratyuk,  Yuri,  227 
Kotanchik,  Joseph  N.,  363,  365-366 
Kraft,  Christopher  C.,  Jr.,  17,  63,  455 
Krieger,  Robert  L.,  32 
Kubrick,  Stanley,  271 
Kuhn,  Thomas  S.,  xxx 
Kurbjun,  Max  C.,  232,  246,  257,  371 
Kyle,  Howard  C.,  57,  65 
Land,  Emory  S.,  Sill. 
Landing  Loads  Track,  93 
Langley, 

budget,  107-108 

competition  among  divisions,  91-93 

organization,  1962,  85-91,  S7ill. 

policy  of  obscurantism,  91-93,  401-402 

public  relations,  22-25,  412-425 

relationship  with  headquarters,  104-109, 
398-407 

reorganization,  1970,  393-394,  403-418, 
405  ill. 

"shadow"  and  ad  hoc  groups,  85,  92 
Langley  Memorial  Aeronautical  Labora- 
tory, 4 

Langley  Researcher,  396UL,  410UI. 
Langley,  Samuel  P.,  4 
Lansing,  Donald,  17 

Large  Launch  Vehicle  Planning  Group,  255 
Large  Orbiting  Research  Laboratory 

(LORL),  301 
lasers,  138,  149,  151 
Launch  Operations  Center,  102 
Leiss,  Abraham,  202 


Levine,  Arnold  S.,  118 

Lewis  Flight  Propulsion  Laboratory,  7,  13, 

395 
Lewis,  George  W.,  10,  23,  91,  401,  404, 

Sill.,  Qill.,  Uill. 
Lewis  Research  Center,  7,  32,  202,  232, 

246,  261 
Ley,  Willy,  383 

Lidar  In-Space  Technology  (LITE),  438 
Life  magazine,  266UI. 
life-support  system,  287,  289-292,  297, 

302-305 

lifting-body  research,  435,  511,  517 
Lina,  Lindsay  J.,  245 
Lindbergh,  Charles  A.,  7,  38,  58,  91,  97, 

384,  Gill. 
Ling-Temco-Vought  (LTV),  112,  203-205, 

213,  214-215,  216,  217,  204.Ul. 
Little  Joe,  28,  47,  50,  59,  77,  48a//.,  Will. 
Lockheed  aircraft, 

F-104  Starfighter,  20 
Lockheed  Missiles  and  Space  Company, 

277,  327,  331 
Loftin,  Laurence  K.,  Jr.,  71,  89,  92,  99, 

116,  118,  226,  251,  252,  276,  367, 

414,  421,  422,  88HL,  90i//.,  398 ill. 
AMIS,  272-274 

Cortright  reorganization,  409-411 
space  station,  285-286 
Los  Alamos  National  Laboratory,  149 
Love,  Eugene  S.,  233,  275,  421 
Low  Committee,  243,  245 
Low,  George  M.,  226,  241,  243,  245, 

249,  250,  258-259,  262-263,  268, 
283-284,  389,  451,  259z//.,  409z//. 

lunar  excursion  module  (LEM),  230, 

237,  238,  242,  316-317,  373-375, 
377-378,  243i//.,  266z//.,  376UL, 
377UL 

lunar  exploration  committees,  table,  225 

Lunar  Exploration  Working  Group, 
223-224,  226 

lunar  landing  mission  modes, 

compared,  234-241,  255-257,  261-262, 

247z//.,  254z//.,  256HI. 
direct  ascent,  237,  239-241,  242,  246, 

250,  255,  261-262 


526 


Index 


lunar  landing  mission  modes  (continued): 
earth-orbit  rendezvous  (EOR),  xxix, 
231-232,  235,  237,  239,  241,  246, 
247,  253,  255,  263 

lunar-orbit  rendezvous  (LOR),  xxix, 
233-239,  241-247,  250,  252,  253, 
255,  257,  258,  259,  260-270,  356, 
361-362,  495-496 
lunar-surface  rendezvous,  252,  255 
Lunar  Landing  Research  Facility,  371, 
373-379,  375iH.,  376iH.,  377UL, 
392i/J. 

Lunar  Landing  Training  Vehicle,  378-379 
Lunar  Mission  Steering  Group,  232-233 
Lunar  Orbit  and  Letdown  Approach 

Simulator,  371,  379,  380i//. 
Lunar  Orbiter,  112,  196,  295,  297,  356, 

400,  314i/J. 

assignment  to  Langley,  319-321 
basic  research  vs.  project  objectives, 

315-319,  320 
budget,  313,  330 
camera,  324-325,  328-329,  333,  336,  342, 

329*0. 

concurrent  development,  322 
contract,  325-332,  334-336,  340-342, 

344-346,  331»H. 
earth  photos,  344-346,  ML,  345UL 

352UI. 

lunar  gravity  data,  346 
mission  objective,  313,  315,  319,  346 
NASA  headquarters,  335-336 
photographic  targets,  336-342,  346-350, 

337tH. 

photometry  data,  349-350 
project  management,  350-353 
project  start-up,  322-326 
solar  radiation,  330-331,  348 
typical  flight  sequence,  34QUI. 
Lunar  Orbiter  I,  342,  344,  345,  346,  506, 

343tM.,  3Mill. 
Lunar  Orbiter  II,  346,  348 
Lunar  Orbiter  III,  318,  346 
Lunar  Orbiter  IV,  346 
Lunar  Orbiter  V,  346,  347,  349,  352 
Lunar  Orbiter  Project  Office  (LOPO),  91, 

323-324,  328,  332-334 
lunar-orbit  rendezvous  (LOR).    See  lunar 
landing  mission  modes. 


lunar  surface,  311-312,  316-317,  314iH. 

Carpathian  Mountains,  348i//. 

Copernicus  crater,  348i//. 

dark  side,  347i//. 

Guy-Lussac  Promontory,  348z//. 

Kepler  crater,  318i//. 

landing  dynamics  test,  318i//. 

Sea  of  Tranquility,  346,  356,  428 

Tycho  crater,  349i/i. 
lunar  vacation,  428 
Lundin,  Bruce  T.,  202,  250 
Lundin  Committee,  250-253,  255 
Lunik,  125 

Lunney,  Glynn,  57,  64 
Lust,  Riemar,  143-144 
Mace,  William,  232 
Maggin,  Bernard,  92,  245 
Maglieri,  Domenic,  17 
Magnetic  Compression  Experiment,  148 
Magnetohydrodynamics  Section,  129, 

147-149 
magnetoplasmadynamics  (MPD), 

conferences,  130 

definition,  121-122 

early  theories,  122-126 

Inspection  1959  exhibits,  126,  135 

wane  of,  150-152 

Magnetoplasmadynamics  (MPD)  Branch, 
85,  191 

creation,  126 

dissolution,  150-152,  421 

facilities,  133-141 

goals,  132 

organization,  129-130 

staff,  129,  130-131 

Thompson,  Floyd  L.,  126-127 
magnetosphere,  123-124,  142,  144,  145, 

146,  217 

Mailer,  Norman,  429 
Main  Committee,  NACA,  7 
Malley,  George,  387 
Manhattan  Project,  113,  147,  149 
Man-In-Space-Soonest  project  (MISS),  57 
Manned  Lunar  Landing  and  Return 

(MALLAR),  228 

Manned  Lunar  Landing  Involving  Ren- 
dezvous (MALLIR),  247-248,  257, 


Manned  Lunar  Landing  Task  Group,  243 


527 


Spaceflight  Revolution 


Manned  Orbital  Rendezvous  and  Docking 

(MORAD),  247-248,  257 
Manned  Orbiting  Research  Laboratory 

(MORL),  91,  293-302,  307-308, 

296»H.,  298tH.,  300tH.,  3Q2UI. 
contract  bidding,  294-302,  307 
lunar  exploration,  299  300 
Studies  Office,  294,  297,  303,  308 
Manned  Spacecraft  Center,  102,  107,  261, 

262,  263,  264,  316-317,  356 
Mattson,  Axel  T.,  357-366 
Manned  Space  Flight  Management  Coun- 
cil, 263,  265 
Manned  Space  Laboratory  Research 

Group,  231-232,  233,  274-276,  289 
Marcus,  Greil,  429 
Marine  Corps,  U.S.,  210 
Mariner  C,  328 
Mark,  Hans,  393 
Mars,  132,  151,  306,  353,  383,  404,  408, 

413,  423,  429,  432,  434,  151tf/., 

4ff?UL 
Marshall  Space  Flight  Center,  102,  108, 

232,  233,  246,  252,  255,  263, 

264-265,  303,  356,  394 
Martin  aircraft, 

YP6M-1  Seamaster,  93,  95 
Martin  Company,  276,  327 
Martin,  James  V.,  331,  332-333,  406, 

333  ttf. 

Martin  Marietta  Corporation,  433 
Marvin,  Charles  F.,  Sill. 
Massachusetts  Institute  of  Technology 

(MIT),  108 
Millstone  Hill  Radar  Observatory,  170, 

173 

Masursky,  Harold,  339 
Mather,  G.  R.,  149 
Mathews,  Charles  W.,  52,  54,  63,  65,  263, 

387 

Mattson,  Axel  T.,  33-37,  107,  35tM. 
lunar-orbit  rendezvous,  263 
Manned  Spacecraft  Center,  357-366, 

379,  36 lill 
Visitors'  Center,  413 
Max  Planck  Institut,  143,  145,  146 
Mayer,  John  P.,  57 
Maynard,  Owen  E.,  235,  241,  262 
Mayo,  Robert,  429 


Mayo,  Wilbur  L.,  230 
McCauley,  John  F.,  339 
McCurdy,  Howard,  432-433 
McDonnell  aircraft, 

F-101A,  17 

McDonnell  Aircraft  Corporation,  45 
McDonnell  Douglas,  307 
McDougall,  Walter  A.,  xxvii,  171,  431-432 
McHatton,  Austin,  188 
McKinsey  &  Co.,  118 
McNamara,  Robert  S.,  273,  301 
Meintel,  Alfred  J.,  Jr.,  371 
Mercury-Atlas  6,  76 
Mercury-Atlas  7,  76 
Mercury-Atlas  8,  77 
Mercury-Atlas  9,  77 
Mercury  Mark  II,  385 
Mercury  project, 

astronauts,  39-45,  77,  456 

capsule,  45,  46,  52,  53,  60,  247,  458 
29UL,  31UL,  Mill.,  45UL,  46UI. 
See  also  Big  Joe. 

contract  requirements,  59,  65 

control  center,  64,  66 

early  skepticism,  58-59 

flight  concept,  38,  51,  52-55,  39ill. 

impact  tests,  60,  65,  96,  362,  61  tW.,  62ill. 

Langley  support  of,  59-63,  65-66,  77 

orbital  flights,  76,  77 

tracking  range,  63-69,  68 ill. 
Meyer,  Andre,  53 
Michael,  William  H.,  Jr.,  152,  223,  224, 

226-230,  237,  246,  227HI. 
micrometeorites,  188,  189,  216,  278-280, 

281 

microwave  cavity  resonance,  138 
Military  Test  Space  Station  (MTSS),  298 
Miller,  James,  184z/£. 

Millstone  Hill  Radar  Observatory,  170,  173 
Minitrack  Network,  65,  161 
Minzer,  Raymond,  164,  165 
MIR,  307-308 

Mission  Control,  63-64,  268,  356,  391 
Mitchell,  Jesse  L.,  167,  170-171,  181,  193, 

mill. 
Moffett  Field,  7.  See  also  Ames  Research 

Center. 

Moffett,  William  A.,  7,  Sill. 
Moonball  experiment,  316 


528 


Index 


Moore,  William  M.,  202 
Morrison  Planetarium,  172 
Moulton,  Forrest  R.,  223 
Mrazek,  William  A.,  234 
Mueller,  George  E.,  397,  337 ill. 
Mumford,  Lewis,  429 
Muroc  Dry  Lake,  202 
Muroc  Field,  7 
Mutual  Weapons  Defense  Program 

(MWDP),  453 
Mylar,  167 
NACA,  xxviii, 

aeronautical  research,  15-17 

appropriations,  9,  10,  23 

charter,  3-7 

committee  system,  7-10,  449,  Sill. 

conferences.  See  Inspections,  NACA. 

creation  of,  2 

director  of  research,  10 

headquarters,  400,  401 

Lewis,  451 

Nuts,  24,  80 

public  relations,  22-25 

research  management,  23-24 

transition  to  NASA,  1-3,  14-15 

young  Turks,  11,  451 

NACA  Conference  on  High-Speed  Aerody- 
namics, 53 
NASA, 

after  Apollo,  308,  392 

creation  of,  xxviii,  1-2,  3,  10-11 

First  Anniversary  Inspection.  See  In- 
spection, 1959. 

intercenter  relationships,  35-37,  178-179, 
194,  357 

inventions  and  contributions  board,  384, 
495 

logo,  438 

NASA  Flight  Research  Center,  32 
NASA  headquarters,  91,  92,  100,  108, 
395-397 

center  management,  224,  319,  322, 
400-401,  424 

Lunar  Orbiter,  335-336 

personnel,  107 

NASA/Industry  Apollo  Technical  Confer- 
ence, 233,  253 

NASA  Structural  Analysis  (NASTRAN), 
414 


National  Academy  of  Public  Administra- 
tion, 336,  350 

National  Academy  of  Sciences,  130,  158 
National  Aeronautics  and  Space  Act,  11, 

14,  15,  16,  22 
National  Aeronautics  and  Space  Council, 

12 
National  Aero-Space  Plane  (NASP),  432, 

436 
National  Orbiting  Space  Station  (NOSS), 

298 

National  Science  Foundation,  12,  41 
National  Transonic  Facility  (NTF),  418, 

436-437,  417*//. 
NATO,  267,  453 
Naugle,  John  E.,  474 
Naval  Aviation  Medical  Acceleration 

Laboratory,  42,  45 
Naval  Research  Laboratory  (NRL),  32, 

34-37,  166,  169,  178,  395 
Navy,  U.S.,  16,  95 
Polaris,  198,  200 
R6L  biplane  amphibian,  16 
Transit  system,  216 
Nelson,  Clifford  H.,  323,  324,  326,  332-334, 

334-335,  342,  351,  323UL,  333UL, 

345z«.,  398z//. 
Netherlands,  217 
Newcomb,  John  F.,  324 
Newell,  Homer  E.,  144,  319,  325,  395,  401, 

474 

Newport  News  Daily  Press,  74-75,  155 
Nichols,  Mark  R.,  90,  99,  231,  274,  275 
Nicks,  Oran  W.,  319,  320,  325,  326,  395, 

408-409,  411,  345UL 
Nike-Cajun  (CAN),  164 
Nike-Deacon  (DAN),  164 
Nike- Tomahawk,  145 
Nixon,  Richard  M.,  xxviii,  xxix,  148,  151, 

249,  306,  307,  429-431 
North  American  aircraft,  283,  295,  297, 

384,  385,  391 
X-15,  20,  28,  32,  54,  98,  131,  158,  371, 

435,  456,  21ili,  32ill. 
XB-70  Valkyrie,  17,  115 
XLR-99,  32 
North  American  Aviation,  281,  290,  362, 

363,  365 
North  American  Rockwell,  146,  307 


529 


Spaceflight  Revolution 


North  Atlantic  Treaty  Organization 
(NATO),  267,  453 

Nova  rocket,  237,  239-241,  244,  246,  250, 
252,  258,  261-262 

nuclear  power,  288-289 

nuclear  weapons,  200 

Oberth,  Hermann,  223,  271 

oblique  photography,  345 

O'Brien-Joyner,  Kitty,  1Q5HI. 

Oertel,  Goetz  K.  H.,  130,  148 

Office  for  Flight  Projects,  91,  367 

Office  of  Advanced  Research  and  Technol- 
ogy (OART),  94,  106,  322 

Office  of  Aeronautical  and  Space  Research, 
92 

Office  of  Engineering  and  Technical  Ser- 
vices, 369 

Office  of  International  Programs,  30 

Office  of  Launch  Vehicle  Programs,  202, 
239,  255 

Office  of  Manned  Space  Flight,  301,  319, 
325,  356,  397 

Office  of  Space  Flight  Development,  30,  55, 
72,  178-179,  202 

Office  of  Space  Flight  Programs,  246,  251 

Office  of  Space  Sciences,  177,  178,  315, 
316,  317,  319,  320 

Office  of  Space  Sciences  and  Applications 
(OSSA),  144,  146,  148,  322,  324, 
325,  326,  327 

Olstad,  Walter  B.,  150 

O'Neal,  Robert  L.,  241 

One-Man  Propulsion  Research  Apparatus, 
371 

"One-O- Wonder",  17 

Operation  Paperclip,  130 

Osborne,  Robert  S.,  294,  303-305,  275iH. 

Ostrander,  Donald  R.,  202,  251 

O'Sullivan,  William  J.,  Jr.,  92,  156, 
157-162,  163,  164-166,  169, 
170-171,  176-177,  178,  180-181, 
185,  186,  191,  192-193,  476-477, 
157tW.,  174tH.,  175t//.,  183UI. 

P-26A  Peashooter,  18ill. 

P-59,  Gill. 

Pageos,  193,  193UI. 

Pageos  1,  191-192 

Paine,  Thomas  O.,  395,  424,  428,  429, 
388UI. 


Pan  American  World  Airways,  428 

panto-base  airplane,  93 

Parasev,  385-387,  386UI. 

Parasev  I,  386 

parawing,  381-387,  382UI. 

Parawing  Project  Office,  387 

Parker,  Eugene  N.,  125,  143 

Parkinson,  John  B.,  93-94 

Passive  Geodetic  Earth-Orbiting  Satellite 
(Pageos),  191 

Pearson,  E.  O.,  Jr.,  245 

Pearson,  Henry,  42 

penetrometer,  316-317 

penetrometer  feasibility  study  group,  316 

Peninsula  Chamber  of  Commerce  (PCC), 
461 

Pennington,  Jack  E.,  371 

Perry,  Tom,  211,  217 

personnel, 

agency  ceilings,  104-106 
congressional  authorization,  109 
distribution  among  NASA  centers,  104, 

106,  1QMI. 

growth  at  Langley,  102,  1Q3UI. 
NASA  headquarters,  107 
recruitment  policy,  411-412 

Petersen,  Richard  H.,  391,  517,  417tH. 

Phillips,  W.  Hewitt,  42,  232,  275,  373, 
374121 

Picasso,  Pablo,  429 

Pierce,  John  R.,  163,  175,  176,  177,  178 

Piland,  Robert  O.,  197,  234,  235 

Pilotless  Aircraft  Research  Division 

(PARD),  56-57,  89,  115-116,  167, 
197,  198,  205 

Pilotless  Aircraft  Research  Station,  7.  See 
also  Wallops  Station. 

Pioneer  4,  172 

Pioneer  10,  431 

plasma  accelerators,  133-141 

Plasma  Applications  Section,  129,  140 

Plasma  Physics  Section,  129,  130,  151 

plasma  sheath,  132-133 

Plott,  Anna,  323 

Pope  Paul  VI,  217 

Porter,  Richard  W.,  165,  166 

Powers,  John  A.,  80,  78ill. 

President's  Science  Advisory  Committee 
(PSAC),  51,  176,  449 


Index 


primate  pilots,  41,  47 
Pritchard,  E.  Brian,  150 
Projection  Planetarium,  371 
project  work,  108-109,  112-119 

predominance  of,  315-319,  320,  406-408, 
418-420,  422-425,  434.  See  also 
research  culture  deterioration, 
public  relations,  412-413 
Langley,  357 
NACA,  22-25 

Purser,  Paul,  47,  52,  235,  360 
Queijo,  Manuel  J.,  241 
Radio  Attenuation  Measurements  (RAM), 

112,  141,  216 
radio  blackout,  133,  140 
Radio  Corporation  of  America  (RCA), 

176,  178,  328-330 

radio-transmission  attenuation,  140 
Rand  Corporation,  231,  325,  328 
Ranger,  113,  312,  315-316,  317,  319,  332, 

504 

Reagan,  Ronald,  xxix,  219,  428,  429,  435 
Recant,  Isadore  G.,  345 ill. 
Redstone  Arsenal,  34 
Redstone  rockets,  50,  59 
Reduced  Gravity  Walking  Simulator,  371, 

377 

reentry  experiments,  132-141,  150,  216 
Reid,  Henry  J.  E.,  10,  36-37,  55,  58,  81, 
91,  96,  394,  409,  9ill.,  Uill,  29ill. 
relativity  theory,  216 
Relay  1,  194 
Relay  2,  194 
rendezvous, 
earth-orbit.  See  lunar  landing  mission 

modes, 
importance  of,  230,  232,  233,  235,  248, 

249-250,  251,  253 
launch  window,  231 
lunar-orbit.  See  lunar  landing  mission 

modes. 

Michael's  parking  orbit,  226-230 
simulators,  281-283 
space  station,  286 
Rendezvous  and  Docking  Simulator, 

371-373,  372z//. 
Republic  Aviation,  98,  332 
Research  Airplane  Project  (RAP),  17 


research  culture  deterioration,  59,  70-72, 
76,  116,  152,  315-319,  320,  406-408, 
418-420,  422-425,  434,  439 
research  management,  23-24 
Research  Steering  Committee  on  Manned 

Space  Flight,  226 

Resident  Research  Associates  (RRAs),  130 
retreats,  staff,  418,  422 
Reynolds  Metals  Company,  167-169 
Rickenbacker,  Edward  V.,  7 
Roberts,  Leonard,  224,  311 
Rogallo,  Francis  M.,  380-384,  387,  382UI. 
Rogallo,  Gertrude,  381,  384,  387,  382UI. 
Roosevelt,  Franklin  D.,  23 
Rosen,  Milton  W.,  255 
Ross,  H.  E.,  227 
rotating  hexagon,  273UL 
Rotating  Vehicle  Simulator,  371 
Rowan,  Lawrence,  339,  345z//. 
Royal  Aircraft  Establishment,  130 
R6L  biplane  amphibian,  16 
Rupp,  George,  210 
Ryan  Company,  384,  385 
Saint  project,  235 
Salinger,  Pierre,  249 
Salyuts,  307,  308 
"Sam",  47 

San  Marco,  217,  218UI. 
satellites, 

communication  (comsats),  163,  175-177 

geosynchronous,  162-164 

global  telecommunication  network, 
194-196 

mapping,  189,  192,  216 

passive,  176 

passive  vs.  active,  163,  179,  190-191,  194 

weather  monitoring,  173 
Saturn  C-l,  252,  229UI. 
Saturn  C-3,  252 
Saturn  I,  293,  297 

Saturn  rocket,  98,  236,  239,  245,  246,  249, 
263,  283,  286,  293,  297,  306,  307, 
356 

Saturn  V,  237,  429 
Savage,  Melvyn,  239 

Scherer,  Lee  R.,  325,  326,  328,  345,  330i//. 
Schirra,  Walter  M.,  Jr.,  40,  42,  77,  40i//., 


Schnitzer,  Emanuel,  277-278 


531 


Spaceflight  Revolution 


Schult,  Eugene  D.,  202,  210,  213,  215, 

211$, 

Schy,  Albert  A.,  224 
science  fiction,  162,  163 
Scopinski,  Ted,  57 

Scout  project,  28,  112,  144,  145,  236,  247, 
283,  321,  322,  400,  406,  204i//., 
208i//.,  214*H. 

approval,  199-200 

Blue  Scout,  200,  209 

contractors,  203-205,  212-215 

cost,  217 

Cub  Scout,  205-206 

design,  198,  200-202 

failures,  210-214 

Goddard,  217 

Project  Office,  91,  202-203,  219 

recertification,  212,  214-216 

tests,  206-209,  201tZ/.,  2Q7UL,  212UL, 

213UL,  2Uill. 
scramjets,  152,  436 

Seamans,  Robert  C.,  Jr.,  235,  236,  239, 
249-251,  253,  255,  258,  260,  262, 
295,  327,  387,  388-389,  391 
Sea  of  Tranquility,  346,  356,  428 
Sergeant,  198,  201 

X248,  154 

XM-33,  184 

service  module  (SM),  237 
737-100,  415,  416tH. 
757,  416 
767,  416 

SF2Y-1  Sea  Dart,  93 
Shea,  Joseph  F.,  259-260,  262,  264,  366 
Shepard,  Alan  B.,  Jr.,  40,  42,  222,  249, 


Sherwood  project,  113,  147 
shop  conditions, 

Langley,  513 

Shortal,  Joseph  A.,  89,  167,  176,  177 
Shotput,  113,  179,  184-187,  188,  184tH. 
Shotput  1,  153-156,  160,  185-186 
Siegel,  Gerald,  172 
Sikorsky  aircraft, 

twin-float  Amphibian,  94 
Silverstein,  Abe,  30,  55,  72-74,  178-179, 
202,  246,  251,  258,  395,  411,  451 
Silvertson,  Wilford  E.,  Jr.,  233 
Simpkinson,  Scott,  47 


simulation  research,  369-371 
simulators, 

closed-loop  analog,  43 

Gemini  and  Apollo,  371-380 

lunar  gravity,  282 ill. 

rendezvous,  281-283 
Sinclair,  A.  Wythe,  Jr.,  275 
Skylab,  270-271,  306,  307,  308,  430 
Slayton,  Donald  K.,  40,  42,  4(H//. 
Smith,  Francis  B.,  86,  91,  367 
Smith,  Norman  F.,  37 
Smithsonian  Astrophysical  Laboratory, 

164 

Society  of  Automotive  Engineers.  232,  233 
solar  corona,  147,  148 
solar  cycle,  123 
solar  flares,  125,  146,  148 
solar  physics,  147 
solar  power,  277,  288,  289 
solar  wind,  125,  142,  143 
solid-fuel  rockets,  198 
Soule,  Hartley  A.,  17,  22,  66,  86 
sound  barrier,  41 

Source  Evaluation  Board,  326-329 
Soviet  Union,  10,  157,  189,  198,  307-308, 

433,  440 
Spaatz,  Carl,  7 

Space  Applications  Branch,  85 
Space  Base,  306-307 
Space  Defense  Initiative  (SDI),  432 
Space  Directorate,  403,  404,  422 
Space  Exploration  Program  Council 

(SEPC),  238-241 

spaceflight  revolution,  xxx-xxxi,  355,  381, 
393,  395,  400,  439-440 

changes  at  Langley,  418-420 

present-day,  431-434,  439-440 
Space  Mechanics  Division  (SMD),  369-380 
Space  Physics  Section,  129,  130,  142 
space  program,  NASA, 

budget,  428,  433 

declining  interest  in,  428-429 

present-day  goals,  433-434 
space  race,  76,  82,  83,  108,  202,  223,  269, 

427-428 
space  research, 

present-day  at  Langley,  437-438 
Space  Sciences  and  Technology  Steering 
Committee,  419-420 


532 


Index 


Space  Sciences  Division,  152 
Space  Sciences  Steering  Committee,  144 
Space  Shuttle,  xxviii,  xxix,  54,  141,  158, 
307,  421,  434-435,  517 

Challenger,  219,  388,  389,  391,  432,  448, 
517 

Columbia,  517 

Nixon,  Richard  M.,  430 
space  sickness,  216 
space  station,  xxix,  189,  421.  See  also 

Manned  Orbiting  Research  Labo- 
ratory (MORL). 

air  lock,  281,  299UI. 

Apollo's  influence,  292-293,  305 

Apollo  X,  301 

artificial  gravity,  281,  287,  295,  297 

budget  constraints,  306 

competing  concepts,  301,  306-307 

construction  material,  291-292 

dynamic  control,  280-281,  286-287 

early  budgets,  283-285 

early  concepts,  223,  226,  271-272 

early  proposals,  277,  278UI. 

inflatable  torus,  277-281,  292,  279z//., 
280UL,  2S2UI. 

life-support  system,  287,  289-292,  297, 
302-305 

Nixon,  Richard  M.,  430 

opposition  to,  276 

power  source,  287-289,  297 

rotating  hexagon,  281-283,  292,  295, 
297,  284UL 

skipped  step,  307-309 

testimony  before  congress,  285-286 
Space  Station  Alpha,  308,  435 
Space  Station  Freedom,  308,  432 
Space  Station  Research  Group,  294 
Space  Station  Task  Force,  309 
Space  Systems  Research  Division,  421,  422 
Space  Task  Group  (STG),  178,  202,  222, 
321,  455 

facilities,  72 

Goddard,  72-75 

Hampton  parade,  77-80,  7SUL,  79ill. 

lunar-orbit  rendezvous,  234,  235,  236, 
241-245 

move  to  Texas,  74,  76-80,  461 

1959  inspection,  37-38,  40-45 

promotions,  71 


Space  Task  Group  (STG)  (continued): 
rendezvous,  245,  246,  248,  255,  257,  260, 

261-263 

staffing,  55-59,  60-62,  69,  72,  73-74 
Space  Technologies  Laboratories  (STL), 

321,  325,  326,  327,  328,  330,  331 
Space  Transportation  System  (STS),  219, 

307,  434,  517 

Space  Vehicle  Group,  167,  173,  177,  178 
Special  Committee  on  Space  and  Astro- 
nautics, 10 

Spirit  of  St.  Louis,  384 
Spitzer,  Lyman,  Jr.,  158 
Sputnik,  76,  97,  199,  381,  384,  439,  440 
crisis,  xxvi-xxix,  10,  99,  156,  171-172, 

369,  449-450 

Sputnik  1,  xxxi,  19,  171,  172,  249,  383,  447 
Sputnik  2,  xxvii,  249,  447 
Stack,  John,  20,  22,  89,  98-100,  116,  135, 

451,  453,  35tf/.,  Will. 
"Star  Wars",  432.  See  also  Strategic 

Defense  Initiative  (SDI). 
State  Department,  172 
Stevens  Institute  of  Technology,  149 
Stokes,  Charles  S.,  140 
Stone,  Ralph  W.,  Jr.,  236,  241 
Stoney,  William  E.,  Jr.,  197,  199,  200,  202, 

210 

Strang,  Charles  F.,  387 
Strategic  Defense  Initiative  (SDI),  149 
Stratton,  Samuel  W.,  Sill. 
Strauss,  H.  Kurt,  241 
Structures  Directorate,  403,  404,  414 
Structures  Research  Division,  89,  91,  92 
Stubbs,  Sandy  M.,  363,  365,  366 
Stuhlinger,  Ernst,  394 
Sunflower  Auxiliary  Power  Plant  Project, 

277 
Supersonic  Commercial  Air  Transport 

(SCAT),  101i«. 
supersonic  transport  (SST),  xxix,  20,  27, 

98,  100,  101,  404,  415,  416-418 
Surveyor  Lander,  319 
Surveyor  Orbiter,  319 
Surveyor  project,  113,  252,  312-313,  315, 

327,  334,  335,  350-351 
Swallow  aircraft,  20,  453,  21ill. 
sweptwing  theory,  20,  139 
Swift,  Calvin  T.,  140 


533 


Space/light  Revolution 


Syncom  3,  195 

Systems  Engineering  and  Operations 

Directorate,  404 
Taback,  Israel,  93,  323,  324,  328,  332,  334, 

336,  349,  323*//. 
Tank  No.  1,  45,  60,  94,  95tH. 
Tank  No.  2,  94,  95,  96HL 
Tant,  Larry,  204 
Taub,  Will,  184z//. 
Taylor,  David  W.,  Sill. 
Teague,  Olin  E.,  74 
technical  report,  114,  116 
telecommunications  satellite  (comsat), 

194-196 
Telstar,  65 
Telstar  1,  194 
Telstar  2,  194 
Temple  University  Research  Institute, 

139-140 
Terminal  Configured  Vehicle  and  Avionics 

(TCVA),  415-416 
Theoretical  Mechanics  Division,  86,  223, 

369,  371 

thermonuclear  fusion,  132,  147-149,  150 
Thibodaux,  Joseph  G.,  Jr.,  197 
Thiokol  Company, 

Redstone  Division,  201 
Thorn,  Karlheinz,  130,  148 
Thomas,  Albert  H.,  74 
Thompson,  Floyd  L.,  16,  81-85,  86,  94, 

96,  104,  342,  345,  6iH.,  29z//.,  84tH., 
90tl/.,  264»H.,  279UL,  337UL,  345z//., 
358iH.,  388tH. 

Apollo,  357,  360,  366 

Apollo  204  fire,  387-389 

associate  director,  394-395 

Echo,  165,  167,  176 

Hampton  community,  24 

Lunar  Orbiter,  319-321,  323,  326,  333 

lunar-orbit  rendezvous,  246 

magnetoplasmadynamics,  126  127 

management  style,  91-93,  99-100,  421 

NACA  charter,  4 

personnel,  106-107 

rendezvous,  232 

replacement  of,  395,  397,  396tH. 

Scout,  202,  205 


Thompson,  Floyd  L.  (continued): 
space  station,  274-276,  294 
Space  Task  Group,  55-56,  59-63,  65-66, 

69-72,  75,  76,  77,  80 
Thompson-Ramo-Wooldridge  (TRW),  108, 

277,  321 

Thor-Able,  28,  199 
Thor-Agena,  191 

Thor-Delta,  154,  177,  184,  187,  188 
Tidewater  Virginia,  4,  bill. 
tiger  team,  214-215 
Time-Life,  39,  456 
Titan,  259,  293 
Tolson,  Robert  H.,  230 
Tracking  and  Ground  Instrumentation 

Unit  (TAGIU),  66-69 
Trout,  Otto,  303*H. 
Truscott,  Starr,  94 

Tsiolkovskii,  Konstantin,  223,  226,  271 
Tycho  crater,  349UL 
UFOs,  146,  171 
United  Kingdom,  217 
Upper  Atmosphere  Rocket  Research  Panel, 

156-157,  161,  164,  165 
Urey,  Harold  C.,  224,  315 
U.S.  Army  Map  Service,  192 
U.S.  Coastal  and  Geodetic  Survey,  192 
U.S.  Geological  Survey,  338,  339 
U.S.  National  Committee/International 

Geophysical  Year  (USNC/IGY), 
technical  panel  on  rocketry,  164 
technical  panel  on  the  earth  satellite 

projects,  164,  165,  166,  172-173 
U.S.  Post  Office  Department,  192 
USS  Hornet,  429 
U-2  "spy  plane" ,  20 
Van  Allen,  James,  123,  158 
Van  Allen  radiation  belts,  123-125,  145, 

173,  216,  331,  124i/Z. 
Van  Cleve,  Jon,  204 
Vandenberg  Air  Force  Base  (AFB),  210, 

213,  215,  216 
Vandenberg,  Hoyt,  7 
Van  Dolah,  Robert  W.,  387 
Vanguard  1,  173 
Vanguard  rocket,  166,  170,  173 
Vanguard  satellite,  xxvii,  157-158,  164, 

170,  198 
Variable-Density  Tunnel  (VDT),  305,  436 


534 


Index 


variable  wing  sweep,  100 

Vavra,  Paul  H.,  66 

Vega,  32,  36 

Verne,  Jules,  226 

Vertol  76,  27 

VE-7,  Will. 

Victory,  John  F.,  12,  31,  34,  Sill.,  84z//. 

Vietnam  War,  269,  305 

Viking  Lander,  4Q6UI. 

Viking  Lander  2,  4Q7UI. 

Viking  project,  70,  196,  353,  404-408,  413, 
423,  431,  434 

Viking  Project  Office,  353,  404 

Viking  rocket,  xxvii,  198 

Virginia  Air  and  Space  Center,  413 

Virtual  Image  Rendezvous  Simulator,  371 

Visitors'  Center,  412-413 

Voas,  Robert  T.,  42,  456 

Vogeley,  Arthur  W.,  232,  245,  246,  257, 
261,  369,  371,  373,  111HI 

von  Braun,  Wernher,  xxvi,  34,  50,  198, 
223,  239,  246,  356,  383,  429,  494, 
495,  264.Ul. 

rendezvous,  236,  241,  262,  263-265,  268 
space  station,  271-272 

von  Pirquet,  Guido,  223,  271 

von  Puttkamer,  Jesco,  270 

"Voodoo",  17.  See  also  McDonnell  air- 
craft. 

Vostok  1,  241 

Vought  aircraft, 
VE-7,  16ill. 
XF5U,  58 

Voyager,  431 

VTOL  tests,  95 

V-2  Panel,  156-157,  474 

V-2  rockets,  28,  156,  157 

Walker,  Joseph  A.,  41 

Waller,  Marvin  C.,  371 

Wallops  Station,  7,  27,  28,  47,  60,  89,  107, 
144,  146,  153,  154,  179,  198,  201, 
202,  203,  205,  207,  210,  211,  213, 
215,  367,  383,  206UL,  208UL 

Walter  Kidde  and  Company,  200 

Ward,  Donald  H.,  322 

Water  Immersion  Simulator,  371 

Watson,  William  I.,  323,  324 

Weapons  System  (WS)-llOA,  115,  116, 
276,  Will. 


weather,  controlling,  146 

Webb,  James  E.,  81-82,  84-85,  92,  106, 
118-119,  194,  249,  265,  315,  322, 
331,  357,  397,  421,  84i//.,  279i//., 
331  ill,  358UI. 

West  Area,  72 

Western  Electric  Company,  65,  69 

Whipple,  Fred  L.,  164,  170,  172 

White,  Edward  H.,  373,  387,  455 

White,  George  C.,  Jr.,  387 

Wiesner,  Jerome,  265 

Wilford,  John  Noble,  xxvi,  xxvii 

Williams,  John  J.,  387 

Williams,  Walter  C.,  63 

Wilson,  Herbert  A.,  Jr.,  367,  420 

wind  tunnels, 

Anechoic  Antenna  Test  Facility,  416i//. 
Continuous-Flow  Hypersonic  Tunnel, 


crossed-field  plasma  accelerator,  135-138 
cyanogen  flame  apparatus,  138-140, 


8-Foot  Transonic  Tunnel,  302UI. 
Electro-Magnetic  Hypersonic  Accelera- 

tor Pilot  Model  Including  Arc-Jet 

Ion  Source,  133 

4-Foot  Supersonic  Pressure  Tunnel,  383 
Full-Scale  Tunnel,  74,  384,  Gill.,  ISill., 

Qlill,  95z«.,  386z//. 

Hall-current  plasma  accelerator,  137UI. 
hotshot,  133-134 
impulse,  133,  134 

linear  Hall-current  accelerator,  138 
MPD-arc  plasma  accelerator,  138,  134i//. 
one-megajoule  theta-pinch,  147-148 
plasma-focus  research  facility,  148-149 
7  x  10-Foot  High-Speed  Tunnel,  21  ill., 

6lili,  Will. 
shock  tube,  134-135 
16-Foot  Transonic  Tunnel,  21i//. 
20-  megawatt  plasma  accelerator,  135, 

136UI. 

Unitary  Plan  Wind  Tunnel,  62ill.,  Wlill. 
Variable-Density  Tunnel  (VDT),  ISill. 
Wind  Tunnel  No.  1,  17 
Windier,  Milton  B.,  37 
Wireless  World,  162 
Wolfe,  Elmer  J.,  202 
women  scientists,  IQ5HI. 


535 


Space/light  Revolution 


Wood,  George  P.,  129,  131,  135,  138, 

136UI. 
Working  Group  on  Lunar  and  Planetary 

Surfaces  Exploration,  224 
World  War  I,  2,  4 
World  War  II,  7,  12,  24,  94,  113,  119,  130, 

156,  369 

Wright,  Orville,  7,  Gill,  Sill. 
X-l,  7,  202 
X-14A,  379 
X-15,  20,  28,  32,  54,  98,  131,  132,  158,  371, 

435,  456,  21  ill.,  32ill. 
X-20,  20,  132,  200,  273,  327,  453,  Will 
X-30,  436 

XB-70  Valkyrie,  17,  115 
XF5U,  58 
XLR-99,  32 
X-rays,  217 
Yeager,  Charles  E.,  41 
York,  Herbert,  51,  55 
Young,  Tom,  339,  341,  349 
YP6M-1  Seamaster,  93,  95 
Zimmerman,  Charles  H.,  52,  58,  59,  158, 

458 
Zond  III,  346 


536 


The  Author 


James  R.  Hansen  is  Alumni  Associate  Professor  of  History  and  chair 
of  the  Department  of  History  at  Auburn  University  in  Auburn,  Alabama. 
He  is  the  author  of  three  books  as  well  as  several  book  chapters,  articles, 
and  reviews  in  aerospace  history  and  the  history  of  technology.  His  books 
include:  Engineer  in  Charge:  A  History  of  the  Langley  Aeronautical  Labo- 
ratory, 1917-1958,  which  was  published  by  NASA  in  1987,  and  From  the 
Ground  Up:  The  Autobiography  of  an  Aeronautical  Engineer,  co-authored 
by  American  aviation  pioneer  Fred  E.  Weick  and  published  by  the  Smithso- 
nian Institution  Press  in  1988.  The  latter  book  received  the  History  Book 
Award  of  the  American  Institute  of  Aeronautics  and  Astronautics  (AIAA). 

Professor  Hansen  has  received  a  number  of  citations  for  his  historical 
scholarship,  including  the  Robert  H.  Goddard  Award  from  the  National 
Space  Club  and  distinctions  of  excellence  from  the  Air  Force  Historical 
Foundation  and  the  American  Institute  for  Aeronautics  and  Astronautics. 
Iii  1986-1987  he  gave  historical  talks  around  the  country  as  an  AIAA 
Distinguished  Lecturer.  He  has  served  on  a  number  of  important  advisory 
boards  and  panels,  including  the  Research  Advisory  Board  of  the  National 
Air  and  Space  Museum,  the  Editorial  Advisory  Board  of  the  Smithsonian 
Institution  Press,  the  Board  of  Advisers  for  the  Society  for  the  History 
of  Technology,  and  the  Advisory  Board  for  the  Archives  of  Aerospace 
Exploration  at  Virginia  Polytechnic  Institute  and  State  University.  He  is 
also  a  past  vice-president  of  the  board  of  directors  of  the  Virginia  Air  and 
Space  Center  and  Hampton  Roads  History  Center  in  Hampton,  Virginia. 

At  Auburn  University,  Dr.  Hansen  teaches  courses  on  the  history  of  flight, 
the  history  of  science,  space  history,  as  well  as  a  large  auditorium  class  that 
surveys  the  history  of  technology  from  ancient  times  to  the  present — or,  as 
he  puts  it,  "from  Australopithecus  to  Arthur  Clarke." 

Hansen  earned  an  A.B.  degree,  with  high  honors,  from  Indiana  University 
(1974)  and  an  M.A.  (1976)  and  Ph.D.  (1981)  from  The  Ohio  State  Univer- 
sity. He  was  born  in  Ft.  Wayne,  Indiana,  on  12  June  1952.  He  is  married  to 
Margaret  Anne  Miller-Hansen  and  has  two  children,  Nathaniel  and  Jennifer. 


537 


The  NASA  History  Series 


Reference  Works,  NASA  SP-4000: 

Grimwood,  James  M.  Project  Mercury:    A  Chronology  (NASA  SP-4001,  1963). 

Grimwood,  James  M.,  and  Hacker,  Barton  C.,  with  Vorzimmer,  Peter  J.  Project  Gemini 

Technology  and  Operations:    A  Chronology  (NASA  SP-4002,  1969). 
Link,  Mae  Mills.  Space  Medicine  in  Project  Mercury  (NASA  SP-4003,  1965). 
Astronautics  and  Aeronautics,  1963:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4004,  1964). 
Astronautics  and  Aeronautics,  1964'-    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4005,  1965). 
Astronautics  and  Aeronautics,  1965:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4006,  1966). 
Astronautics  and  Aeronautics,  1966:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4007,  1967). 
Astronautics  and  Aeronautics,  1967:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4008,  1968). 
Ertel,  Ivan  D.,  and  Morse,  Mary  Louise.    The  Apollo  Spacecraft:    A  Chronology,   Volume  I, 

Through  November  7,  1962  (NASA  SP-4009,  1969). 
Morse,  Mary  Louise,  and  Bays,  Jean  Kernahan.     The  Apollo  Spacecraft:     A   Chronology, 

Volume  II,  November  8,  1962-September  30,  1964  (NASA  SP-4009,  1973). 
Brooks,  Courtney  G.,  and  Ertel,  Ivan  D.  The  Apollo  Spacecraft:    A  Chronology,  Volume  III, 

October  1,  1964-January  20,  1966  (NASA  SP-4009,  1973). 
Ertel,  Ivan  D.,  and  Newkirk,  Roland  W.,  with  Brooks,  Courtney  G.  The  Apollo  Spacecraft:  A 

Chronology,  Volume  IV,  January  21,  1966-July  13,  1974  (NASA  SP-4009,  1978). 
Astronautics  and  Aeronautics,  1968:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4010,  1969). 
Newkirk,  Roland  W.,  and  Ertel,  Ivan  D.,  with  Brooks,  Courtney  G.  Skylab:    A  Chronology 

(NASA  SP-4011,  1977). 
Van  Nimmen,  Jane,  and  Bruno,  Leonard  C.,  with  Rosholt,  Robert  L.  NASA  Historical  Data 

Book,  Vol.  I:    NASA  Resources,  1958-1968  (NASA  SP-4012,  1976,  rep.  ed.  1988). 
Ezell,  Linda  Neuman.    NASA  Historical  Data  Book,   Vol  II:    Programs  and  Projects,  1958- 

1968  (NASA  SP-4012,  1988). 
Ezell,  Linda  Neuman.  NASA  Historical  Data  Book,  Vol.  Ill:   Programs  and  Projects,  1969- 

1978  (NASA  SP-4012,  1988). 
Astronautics  and  Aeronautics,  1969:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4014,  1970). 
Astronautics  and  Aeronautics,  1970:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4015,  1972). 
Astronautics  and  Aeronautics,  1971:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4016,  1972). 

539 


Astronautics  and  Aeronautics,  1972:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4017,  1974). 
Astronautics  and  Aeronautics,  1973:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4018,  1975). 
Astronautics  and  Aeronautics,  1974'-    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4019,  1977). 
Astronautics  and  Aeronautics,  1975:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4020,  1979). 
Astronautics  and  Aeronautics,  1976:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4021,  1984). 
Astronautics  and  Aeronautics,  1977:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4022,  1986). 
Astronautics  and  Aeronautics,  1978:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4023,  1986). 
Astronautics  and  Aeronautics,  1979-1984:     Chronology  of  Science,   Technology,  and  Policy 

(NASA  SP-4024,  1988). 
Astronautics  and  Aeronautics,  1985:    Chronology  of  Science,  Technology,  and  Policy  (NASA 

SP-4025,  1990). 
Gawdiak,  Ihor  Y.,  compiler.  NASA  Historical  Data  Book,  Vol.  IV:   NASA  Resources,  1969 

1978  (NASA  SP-4012,  1994). 

Management  Histories,  NASA  SP-4100: 

Rosholt,  Robert  L.  An  Administrative  History  of  NASA,  1958-1963  (NASA  SP-4101,  1966). 

Levine,  Arnold  S.  Managing  NASA  in  the  Apollo  Era  (NASA  SP-4102,  1982). 

Roland,  Alex.    Model  Research:     The  National  Advisory  Committee  for  Aeronautics,  1915- 

1958  (NASA  SP-4103,  1985). 

Pries,  Sylvia  D.  NASA  Engineers  and  the  Age  of  Apollo  (NASA  SP-4104,  1992). 
Glennen,   T.  Keith.      The  Birth  of  NASA:     The  Diary  of  T.  Keith  Glennan,  edited  by 

J.  D.  Hunley  (NASA  SP-4105,  1993). 

Project  Histories,  NASA  SP-4200: 

Swenson,  Loyd  S.,  Jr.,  Grimwood,  James  M.,  and  Alexander,  Charles  C.  This  New  Ocean:  A 

History  of  Project  Mercury  (NASA  SP-4201,  1966). 
Green,  Constance  McL.,  and  Lomask,  Milton.    Vanguard:   A  History  (NASA  SP-4202,  1970; 

rep.  ed.  Smithsonian  Institution  Press,  1971). 
Hacker,  Barton  C.,  and  Grimwood,  James  M.  On  Shoulders  of  Titans:    A  History  of  Project 

Gemini  (NASA  SP-4203,  1977). 
Benson,  Charles  D.,  and  Faherty,  William  Barnaby.  Moonport:   A  History  of  Apollo  Launch 

Facilities  and  Operations  (NASA  SP-4204,  1978). 
Brooks,   Courtney  G.,   Grimwood,   James  M.,   and  Swenson,   Loyd  S.,   Jr.      Chariots  for 

Apollo:    A  History  of  Manned  Lunar  Spacecraft  (NASA  SP-4205,  1979). 
Bilstein,  Roger  E.  Stages  to  Saturn:    A  Technological  History  of  the  Apollo/Saturn  Launch 

Vehicles  (NASA  SP-4206,  1980). 
Compton,  W.  David,  and  Benson,  Charles  D.  Living  and  Working  in  Space:    A  History  of 

Skylab  (NASA  SP-4208,  1983). 

540 


Ezell,  Edward  Clinton,  and  Ezell,  Linda  Neuman.   The  Partnership:  A  History  of  the  Apollo- 

Soyuz  Test  Project  (NASA  SP-4209,  1978). 

Hall,  R.  Cargill.  Lunar  Impact:    A  History  of  Project  Ranger  (NASA  SP-4210,  1977). 
Newell,  Homer  E.  Beyond  the  Atmosphere:    Early  Years  of  Space  Science  (NASA  SP-4211, 

1980). 
Ezell,  Edward  Clinton,  and  Ezell,  Linda  Neuman.   On  Mars:   Exploration  of  the  Red  Planet, 

1958-1978  (NASA  SP-4212,  1984). 
Pitts,  John  A.   The  Human  Factor:     Biomedicine  in  the  Manned  Space  Program  to  1980 

(NASA  SP-4213,  1985). 
Compton,   W.   David.      Where  No  Man  Has  Gone  Before:     A   History  of  Apollo  Lunar 

Exploration  Missions  (NASA  SP-4214,  1989). 
Naugle,  John  E.  First  Among  Equals:    The  Selection  of  NASA  Space  Science  Experiments 

(NASA  SP-4215,  1991). 
Wallace,  Lane  E.  Airborne  Trailblazer:   Two  Decades  with  NASA  Langley's  Boeing  737  Flying 

Laboratory  (NASA  SP-4216,  1994). 

Center  Histories,  NASA  SP-4300: 

Rosenthal,  Alfred.   Venture  into  Space:   Early  Years  of  Goddard  Space  Flight  Center  (NASA 

SP-4301,  1985). 
Hartman,  Edwin  P.  Adventures  in  Research:  A  History  of  Ames  Research  Center,  1940  1965 

(NASA  SP-4302,  1970). 
Hallion,  Richard  P.  On  the  Frontier:  Flight  Research  at  Dryden,  1946-1981  (NASA  SP-4303, 

1984). 
Muenger,  Elizabeth  A.  Searching  the  Horizon:    A  History  of  Ames  Research  Center,  1940 

1976  (NASA  SP-4304,  1985). 
Hansen,  James  R.  Engineer  in  Charge:    A  History  of  the  Langley  Aeronautical  Laboratory, 

1917-1958  (NASA  SP-4305,  1987). 
Dawson,  Virginia  P.  Engines  and  Innovation:    Lewis  Laboratory  and  American  Propulsion 

Technology  (NASA  SP-4306,  1991). 
Dethloff,  Henry  C.  "Suddenly  Tomorrow  Came  .  .  . ":  A  History  of  the  Johnson  Space  Center, 

1957  1990.  (NASA  SP-4307,  1993). 

General  Histories,  NASA  SP-4400: 

Corliss,  William  R.  NASA  Sounding  Rockets,  1958  1968:    A  Historical  Summary  (NASA 

SP-4401,  1971). 
Wells,  Helen  T.,  Whiteley,  Susan  H.,  and  Karegeannes,  Carrie.    Origins  of  NASA  Names 

(NASA  SP-4402,  1976). 
Anderson,  Frank  W.,  Jr.  Orders  of  Magnitude:   A  History  of  N AC  A  and  NASA,  1915-1980 

(NASA  SP-4403,  1981). 

Sloop,  John  L.  Liquid  Hydrogen  as  a  Propulsion  Fuel,  1945-1959  (NASA  SP-4404,  1978). 
Roland,  Alex.    A  Spacefaring  People:    Perspectives  on  Early  Spaceflight  (NASA  SP-4405, 

1985). 
Bilstein,  Roger  E.  Orders  of  Magnitude:    A  History  of  the  NACA  and  NASA,  1915-1990 

(NASA  SP-4406,  1989). 


541 


New  Series  in  NASA  History,  published  by  The  Johns  Hopkins  University  Press: 

Cooper,  Henry  S.  F.,  Jr.  Before  Lift-Off:    The  Making  of  a  Space  Shuttle  Crew  (1987). 
McCurdy,  Howard  E.  The  Space  Station  Decision:    Incremental  Politics  and  Technological 

Choice  (1990). 

Hufbauer,  Karl.  Exploring  the  Sun:    Solar  Science  Since  Galileo  (1991). 
McCurdy,  Howard  E.  Inside  NASA:  High  Technology  and  Organizational  Change  in  the  U.S. 

Space  Program  (1993). 


542