Skip to main content

Full text of "Models of the Atom, Project Physics Text and Handbook Volume 5"

See other formats


The  Project  Physics  Course  Text  and  Handbook 


Models  of  the  Atom 


/•.  •• 


The  Project  Physics  Course 


Text  and  Handbook 


UNIT 


5 


Models  of  the  Atom 


A  Component  of  the 
Project  Physics  Course 


Published  by 

HOLT,  RINEHART  and  WINSTON,  Inc. 

New  York,  Toronto 


Directors  of  Harvard  Project  Physics 

Gerald  Holton,  Department  of  Physics,  Harvard 

University 
F.  James  Rutherford,  Capuchino  High  School, 

San  Bruno,  Cahfornia,  and  Harvard  University 
Fletcher  G.  Watson,  Harvard  Graduate  School 

of  Education 


Acknowledgments,  Text  Section 

The  authors  and  publisher  have  made  every  effort 
to  trace  the  ownership  of  all  selections  found  in  this 
book  and  to  make  full  acknowledgment  for  their  use. 
Many  of  the  selections  are  in  the  public  domain. 

Grateful  acknowledgment  is  hereby  made  to  the 
following  authors,  pubUshers,  agents,  and  individ- 
uals for  use  of  their  copyrighted  material. 


Special  Consultant  to  Project  Physics 

Andrew  Ahlgren,  Harvard  Graduate  School  of 
Education 


A  partial  Ust  of  staff  and  consultants  to  Harvard 
Project  Physics  appears  on  page  iv. 

This  Text-Handbook,  Unit  5  is  one  of  the  many  in- 
structional materials  developed  for  the  Project 
Physics  Course.  These  materials  include  Texts, 
Handbooks,  Teacher  Resource  Books,  Readers, 
Programmed  Instruction  booklets,  Film  Loops, 
Transparencies,  16mm  films,  and  laboratory 
equipment. 


Copyright  ©  1970  Project  Physics 

All  Rights  Reserved 

SBN  03-084501-7 

1234        039        98765432 

Project  Physics  is  a  registered  trademark 


P.  3  Excerpts  from  The  Way  Things  Are:  The  De 
Rerum  Natura  of  Titus  Lucretius  Caius,  a  transla- 
tion by  Rolfe  Humphries,  copyright  ©  1969  by 
Indiana  University  Press. 

P.  5  From  'The  First  Chapter  of  Aristotle's 
'Foundations  of  Scientij&c  Thought'  (Metaphysica, 
Liber  A),"  translated  by  Daniel  E.  Gershenson  and 
Daniel  A.  Greenburg,  in  The  Natural  Philosopher, 
Vol.  II,  copyright  ©  1963  by  the  Blaisdell  Pub- 
lishing Company,  pp.  14—15. 

P.  7  From  The  Life  of  the  Honorable  Henry 
Cavendish,  by  George  Wilson,  printed  for  the 
Cavendish  Society,  1851,  pp.  186-187. 

Pp.  7-8  From  "Elements  of  Chemistry"  by  Antoine 
Laurent  Lavoisier,  translated  by  Robert  Kerr  in  Great 
Books  of  the  Western  World,  Vol.  45,  copyright  1952 
by  Encyclopaedia  Britannica,  Inc.,  pp.  3-4. 

P.  11  From  "The  Atomic  Molecular  Theory"  by 
Leonard  K.  Nash  in  Harvard  Case  Histories  in 
Experimental  Science,  Case  4,  Vol.  1,  copyright  1950 
by  Harvard  University,  p.  228. 

P.  21  From  The  Principles  of  Chemistry  by  Dmitri 
Mendeleev,  translated  by  George  Kamensky,  copy- 
right 1905  by  Longmans,  Green  and  Company, 
London,  p.  27. 

P.  22  Mendeleev,  Dmitri,  1872. 

P.  29  From  "Experimental  Researches  in  Elec- 
tricity" by  Michael  Faraday  from  Great  Books  of  the 
Western  World,  Vol.  45,  copyright  1952  by 
Encyclopaedia  Britannica,  Inc.,  pp.  389-390. 

Pp.  43-44  Einstein,  Albert,  trans,  by  Professor 
Irving  Kaplan,  Massachusetts  Institute  of  Tech- 
nology. 

P.  48  Roentgen,  W.  K. 

P.  57  From  "Opticks"  by  Isaac  Newton  from  Great 
Books  of  the  Western  World,  Vol.  34,  copyright  1952 
by  Encyclopaedia  Britannica,  Inc.,  pp.  525-531. 

P.  67  From  Background  to  Modeim  Science, 
Needham,  Joseph  and  Pagel,  Walter,  eds.,  copyright 
1938  by  The  Macmillan  Company,  pp.  68-69. 

P.  88  Letter  from  Rutherford  to  Bohr,  March  1913. 

P.  91  From  "Opticks"  by  Isaac  Newton  from  Great 
Books  of  the  Western  World,  Vol.  34,  copyright  1952 
by  Encyclopaedia  Britannica,  Inc.,  p.  541. 

P.  113  From  Atom,ic  Physics  by  Max  Bom,  copy- 
right 1952  by  Blackie  &  Son,  Ltd.,  p.  95. 


p.  114  Letter  from  Albert  Einstein  to  Max  Bom, 
1926. 

P.  119  From  A  Philosophical  Essay  on  Possibilities 
by  Pierre  Simon  Laplace,  translated  by  Frederick  W. 
Truscott  and  Frederick  L.  Emory,  copyright  1951 
by  Dover  Publications,  Inc.,  p.  4. 


Picture  Credits,  Text  Section 

Cover  photo:  Courtesy  of  Professor  Erwin  W. 
Mueller,  The  Pennsylvania  State  University. 

P.  1  (top)  Merck  Sharp  &  Dohme  Research 
Laboratories;  (center)  Loomis  Dean,  LIFE 
MAGAZINE,  ©  Time  Inc. 

P.  2  (charioteer)  Hirmer  Fotoarchiv,  Munich; 
(architectural  ruins)  Greek  National  Tourist 
Office,  N.Y.C. 

P.  4  Electrum  pendant  (enlarged).  Archaic. 
Greek.  Gold.  Courtesy,  Museum  of  Fine  Arts, 
Boston.  Henry  Lillie  Pierce  Residuary  Fund. 

P.  7  Fisher  Scientific  Company,  Medford,  Mass. 

P.  10  from  Dalton,  John,  A  New  System  of 
Chemical  Philosophy,  R.  BickerstafF,  London, 
1808-1827,  as  reproduced  in  A  History  of 
Chemistry  by  Charles-Albert  Reichen,  c  1963, 
Hawthorn  Books  Inc.,  70  Fifth  Ave.,  N.Y.C. 

P.  13  Engraved  portrait  by  Worthington  from  a 
painting  by  Allen.  The  Science  Museum,  London. 

P.  15  (drawing)  Reprinted  by  permission  from 
CHEMICAL  SYSTEMS  by  Chemical  Bond  Approach 
Project.  Copyright  1964  by  Earlham  College  Press, 
Inc.  Published  by  Webster  Division,  McGraw-Hill 
Book  Company. 

P.  20  Moscow  Technological  Institute. 

P.  26  (portrait)  The  Royal  Society  of  London. 

P.  27  Courtesy  of  Aluminum  Company  of  America. 

P.  32  Science  Museum,  London.  Lent  by  J.  J. 
Thomson,  M.A.,  Trinity  College,  Cambridge. 

P.  35  Courtesy  of  Sir  George  Thomson. 

P.  39    (top)  California  Institute  of  Technology. 

P.  45  (left,  top)  Courtesy  of  The  New  York  Times; 
(left,  middle)  American  Institute  of  Physics; 
(middle,  right)  Courtesy  of  California  Institute  of 
Technology  Archives;  (left,  bottom)  Courtesy  of 
Europa  Verlag,  Zurich. 

P.  47  (left,  top)  Dr.  Max  F.  Millikan;  (right,  top) 
Harper  Library,  University  of  Chicago;  (right 
margin)  R.  Diihrkoop  photo. 

P.  48  The  Smithsonian  Institution. 

P.  49  Burndy  Library,  Norwalk,  Conn. 

P.  51  Eastman  Kodak  Company,  Rochester,  N.Y. 

P.  52  High  Voltage  Engineering  Corp. 

P.  53  (rose)  Eastman  Kodak  Company;  (fish) 


American  Institute  of  Radiology;  (reactor  vessel) 
Nuclear  Division,  Combustion  Engineering,  Inc. 

P.  58  Science  Museum,  London.  Lent  by  Sir 
Lawrence  Bragg,  F.R.S. 

P.  64  Courtesy  of  Dr.  Owen  J.  Gingerich, 
Smithsonian  Astrophysical  Observatory. 

P.  67  Courtesy  of  Professor  Lawrence  Badash, 
Dept.  of  History,  University  of  California, 
Santa  Barbara. 

P.  76  (top)  American  Institute  of  Physics; 
(bottom,  right)  Courtesy  of  Niels  Bohr  Library, 
American  Institute  of  Physics. 

P.  80  (ceremony)  Courtesy  of  Professor  Edward 
M.  Purcell,  Harvard  University;  (medal)  Swedish 
Information  Service,  N.Y.C. 

P.  93  Science  Museum,  London.  Lent  by  Sir 
Lawrence  Bragg,  F.R.S. 

P.  94  from  the  P.S.S.C.  film  Matter  Waves. 

P.  100  American  Institute  of  Physics. 

P.  102  Professor  Harry  Meiners,  Rensselaer 
Polytechnic  Institute. 

P.  106  American  Institute  of  Physics. 

P.  107  (de  Broglie)  Academic  des  Sciences,  Paris; 
(Heisenberg)  Professor  Werner  K.  Heisenberg; 
(Schrodinger)  Ameriq^n  Institute  of  Physics. 

P.  109    (top,  left)  Perkin-Elmer  Corp. 

P.  112  Orear,  Jay,  Fundamental  Physics,  ©  1961 
by  John  Wiley  &  Sons,  Inc.,  New  York. 

P.  115  The  Graphic  Work  of  M.  C.  Escher, 
Hawthorn  Books  Inc.,  N.Y.  "Lucht  en  Water  2." 

Picture  Credits,  Handbook  Section 

Cover:  Drawing  by  Saul  Steinberg,  from 
The  Sketchbook  for  1967,  Hallmark  Cards,  Inc. 

P.  130  These  tables  appear  on  pp.  122,  157  and 
158  of  Types  of  Graphic  Representation  of  the 
Periodic  System  of  Chemical  Elements  by 
Edmund  G.  Mazurs,  published  in  1957  by  the 
author.  They  also  appear  on  p.  8  of  Chemistry 
magazine,  July  1966. 

P.  136  Courtesy  L.  J.  Lippie,  Dow  Chemical 
Company,  Midland.  Michigan. 

P.  149  From  the  cover  of  The  Science  Teacher, 
Vol.  31,  No.  8,  December  1964,  illustration  for 
the  article,  "Scientists  on  Stamps;  Reflections  of 
Scientists'  Public  Image, "  by  Victor  Showalter, 
The  Science  Teacher,  December  1964,  pp.  40—42. 

All  photographs  used  with  film  loops  courtesy 
of  National  Film  Board  of  Canada. 

Photographs  of  laboratory  equipment  and  of 
students  using  laboratory  equipment  were  supplied 
with  the  cooperation  of  the  Project  Physics  staff 
and  Damon  Corporation. 


Partial  List  of  Staff  and  Consultants 

The  individuals  listed  below  (and  on  the  following  pages)  have  each  contributed  in  some  way  to  the 
development  of  the  course  materials.  Their  periods  of  participation  ranged  from  brief  consultations  to 
full-time  involvement  in  the  team  for  several  years.  The  affiliations  indicated  are  those  just  prior  to 
or  during  the  period  of  participation. 


Advisory  Committee 

E.  G.  Begle,  Stanford  University,  Calif. 

Paul  F.  Brandwein,  Harcourt,  Brace  &  World, 

Inc.,  San  Francisco,  Calif. 
Robert  Brode,  University  of  California,  Berkeley 
Erwin  Hiebert,  University  of  Wisconsin,  Madison 
Harry  Kelly,  North  Carolina  State  College,  Raleigh 
William  C.  Kelly,  National  Research  Council, 

Washington,  D.C. 
Philippe  LeCorbeiller,  New  School  for  Social 

Research,  New  York,  N.Y. 
Thomas  Miner,  Garden  City  High  School,  New 

York. 
Philip  Morrison,  Massachusetts  Institute  of 

Technology,  Cambridge 
Ernest  Nagel,  Columbia  University,  New  York, 

N.Y. 
Leonard  K.  Nash,  Harvard  University 
I.  I.  Rabi,  Columbia  University,  New  York.  N.Y. 

Staff  and  Consultants 

L.  K.  Akers,  Oak  Ridge  Associated  Universities, 

Tenn. 
Roger  A.  Albrecht,  Osage  Community  Schools, 

Iowa 
David  Anderson,  Oberlin  College,  Ohio 
Gary  Anderson,  Harvard  University 
Donald  Armstrong,  American  Science  Film 

Association,  Washington,  D.C. 
Arnold  Arons,  University  of  Washington 
Sam  Ascher,  Henry  Ford  High  School,  Detroit, 

Mich. 
Ralph  Atherton,  Talawanda  High  School,  Oxford, 

Ohio 
Albert  V.  Baez,  UNESCO,  Paris 
William  G.  Banick,  Fulton  High  School.  Atlanta, 

Ga. 
Arthur  Bardige,  Nova  High  School,  Fort 

Lauderdale,  Fla. 
Rolland  B.  Bartholomew,  Henry  M.  Gunn  High 

School,  Palo  Alto,  Calif. 
O.  Theodor  Benfey,  Earlham  College,  Richmond, 

Ind. 
Richard  Berendzen,  Harvard  College  Observatory 
Alfred  M.  Bork,  Reed  College,  Portland,  Ore. 

F.  David  Boulanger,  Mercer  Island  High  School, 
Washington 

Alfred  Brenner,  Harvard  University 
Robert  Bridgham,  Harvard  University 
Richard  Brinckerhoff,  Phillips  Exeter  Academy, 
Exeter.  N.H. 


Donald  Brittain,  National  Film  Board  of  Canada. 

Montreal 
Joan  Bromberg,  Harvard  University 
Vinson  Bronson,  Newton  South  High  School, 

Newton  Centre,  Mass. 
Stephen  G.  Brush,  Lawrence  Radiation  Laboratory, 

University  of  California.  Livermore 
Michael  Butler.  CIASA  Films  Mundiales.  S.  A.. 

Mexico 
Leon  Callihan,  St.  Mark's  School  of  Texas.  Dallas 
Douglas  Campbell,  Harvard  University 
J.  Arthur  Campbell,  Harvey  Mudd  College, 

Claremont,  California 
Dean  R.  Casperson.  Harvard  University 
Bobby  Chambers.  Oak  Ridge  Associated 

Universities.  Tenn. 
Robert  Chesley.  Thacher  School,  Ojai,  Calif. 
John  Christensen.  Oak  Ridge  Associated 

Universities,  Tenn. 
David  Clarke.  Browne  and  Nichols  School. 

Cambridge.  Mass. 
Robert  S.  Cohen.  Boston  University.  Mass. 
Brother  Columban  Francis.  F.S.C..  Mater  Christi 

Diocesan  High  School.  Long  Island  City.  N.Y. 
Arthur  Compton.  Phillips  Exeter  Academy, 

Exeter.  N.H. 
David  L.  Cone,  Los  Altos  High  School,  CaUf. 
William  Cooley.  University  of  Pittsburgh.  Pa. 
Ann  Couch.  Harvard  University 
Paul  Cowan,  Hardin-Simmons  University. 

Abilene,  Tex. 
Charles  Davis.  Fairfax  County  School  Board. 

Fairfax.  Va. 
Michael  Dentamaro.  Senn  High  School.  Chicago, 

111. 
Raymond  Dittman.  Newton  High  School.  Mass. 
Elsa  Dorfman.  Educational  Services  Inc.. 

Watertown.  Mass. 
Vadim  Drozin.  Bucknell  University.  Lewisburg, 

Pa. 
Neil  F.  Dunn.  Burlington  High  School.  Mass. 
R.  T.  Ellickson.  University  of  Oregon.  Eugene 
Thomas  Embry.  Nova  High  School.  Fort 

Lauderdale.  Fla. 
Walter  Eppenstein.  Rensselaer  Polytechnic 

Institute,  Troy,  N.Y. 
Herman  Epstein.  Brandeis  University.  Waltham. 

Mass. 
Thomas  F.  B.  Ferguson.  National  Film  Board  of 

Canada.  Montreal 
Thomas  von  Foerster.  Harvard  University 


(continued  on  p.  122) 


Science  is  an  adventure  of  the  whole  human  race  to  learn  to  live  in  and 
perhaps  to  love  the  universe  in  which  they  are.  To  be  a  part  of  it  is  to 
understand,  to  understand  oneself,  to  begin  to  feel  that  there  is  a  capacity 
within  man  far  beyond  what  he  felt  he  had,  of  an  infinite  extension  of 
human  possibilities  .  .  . 

I  propose  that  science  be  taught  at  whatever  level,  from  the  lowest  to  the 
highest,  in  the  humanistic  way.  It  should  be  taught  with  a  certain  historical 
understanding,  with  a  certain  philosophical  understanding ,  with  a  social 
understanding  and  a  human  understanding  in  the  sense  of  the  biography,  the 
nature  of  the  people  who  made  this  construction,  the  triumphs,  the  trials,  the 
tribulations. 

I.  I.  RABI 

Nobel  Laureate  in  Physics 


Preface 


Background        The  Project  Physics  Course  is  based  on  the  ideas  and 
research  of  a  national  curriculum  development  project  that  worked  in 
three  phases.  First,  the  authors — a  high  school  physics  teacher,  a 
university  physicist,  and  a  professor  of  science  education — collaborated 
to  lay  out  the  main  goals  and  topics  of  a  new  introductory  physics 
course.  They  worked  together  from  1962  to  1964  with  financial  support 
from  the  Carnegie  Corporation  of  New  York,  and  the  first  version  of 
the  text  was  tried  out  in  two  schools  with  encouraging  results. 

These  preliminary  results  led  to  the  second  phase  of  the  Project 
when  a  series  of  major  grants  were  obtained  from  the  U.S.  Office  of 
Education  and  the  National  Science  Foundation,  starting  in  1964. 
Invaluable  additional  financial  support  was  also  provided  by  the 
Ford  Foundation,  the  Alfred  P.  Sloan  Foundation,  the  Carnegie 
Corporation,  and  Harvard  University.  A  large  number  of  collaborators 
were  brought  together  from  all  parts  of  the  nation,  and  the  group 
worked  together  for  over  four  years  under  the  title  Harvard  Project 
Physics.  At  the  Project's  center,  located  at  Harvard  University, 
Cambridge,  Massachusetts,  the  staff  and  consultants  included  college 
and  high  school  physics  teachers,  astronomers,  chemists,  historians 
and  philosophers  of  science,  science  educators,  psychologists, 
evaluation  specialists,  engineers,  film  makers,  artists  and  graphic 
designers.  The  teachers  serving  as  field  consultants  and  the  students 
in  the  trial  classes  were  also  of  vital  importance  to  the  success  of 
Harvard  Project  Physics.  As  each  successive  experimental  version  of 
the  course  was  developed,  it  was  tried  out  in  schools  throughout  the 
United  States  and  Canada.  The  teachers  and  students  in  those  schools 
reported  their  criticisms  and  suggestions  to  the  staff  in  Cambridge, 
and  these  reports  became  the  basis  for  the  subsequent  revisions  of 
the  course  materials.  In  the  Preface  to  Unit  1  Text  you  will  find  a  list  of  the 
major  aims  of  the  course. 


We  wish  it  were  possible  to  list  in  detail  the  contributions  of  each 
person  who  participated  in  some  part  of  Harvard  Project  Physics. 
Unhappily  it  is  not  feasible,  since  most  staff  members  worked  on  a 
variety  of  materials  and  had  multiple  responsibilities.  Furthermore, 
every  text  chapter,  experiment,  piece  of  apparatus,  film  or  other  item 
in  the  experimental  program  benefitted  from  the  contributions  of  a 
great  many  people.  On  the  preceding  pages  is  a  partial  list  of 
contributors  to  Harvard  Project  Physics.  There  were,  in  fact,  many 
other  contributors  too  numerous  to  mention.  These  include  school 
administrators  in  participating  schools,  directors  and  staff  members 
of  training  institutes  for  teachers,  teachers  who  tried  the  course  after 
the  evaluation  year,  and  most  of  all  the  thousands  of  students  who 
not  only  agreed  to  take  the  experimental  version  of  the  course,  but 
who  were  also  willing  to  appraise  it  critically  and  contribute  their 
opinions  and  suggestions. 

The  Project  Physics  Course  Today.  Using  the  last  of  the  experimental 
versions  of  the  course  developed  by  Harvard  Project  Physics  in 
1964-68  as  a  starting  point,  and  taking  into  account  the  evaluation 
results  from  the  tryouts,  the  three  original  collaborators  set  out  to 
develop  the  version  suitable  for  large-scale  publication.  We  take 
particular  pleasure  in  acknowledging  the  assistance  of  Dr.  Andrew 
Ahlgren  of  Harvard  University.  Dr.  Ahlgren  was  invaluable  because 
of  his  skill  as  a  physics  teacher,  his  editorial  talent,  his  versatility 
and  energy,  and  above  all,  his  commitment  to  the  goals  of  Harvard 
Project  Physics. 

We  would  also  especially  like  to  thank  Miss  Joan  Laws  whose 
administrative  skills,  dependability,  and  thoughtfulness  contributed  so 
much  to  our  work.  The  publisher.  Holt,  Rinehart  and  Winston,  Inc. 
of  New  York,  provided  the  coordination,  editorial  support,  and  general 
backing  necessary  to  the  large  undertaking  of  preparing  the  final 
version  of  all  components  of  the  Project  Physics  Course,  including 
texts,  laboratory  apparatus,  films,  etc.  Damon,  a  company  located  in 
Needham,  Massachusetts,  worked  closely  with  us  to  improve  the 
engineering  design  of  the  laboratory  apparatus  and  to  see  that  it  was 
properly  integrated  into  the  program. 

In  the  years  ahead,  the  learning  materials  of  the  Project  Physics 
Course  will  be  revised  as  often  as  is  necessary  to  remove  remaining 
ambiguities,  clarify  instructions,  and  to  continue  to  make  the  materials 
more  interesting  and  relevant  to  students.  We  therefore  urge  all 
students  and  teachers  who  use  this  course  to  send  to  us  (in  care  of 
Holt,  Rinehart  and  Winston,  Inc.,  383  Madison  Avenue,  New  York, 
New  York  10017)  any  criticism  or  suggestions  they  may  have. 


F.  James  Rutherford 
Gerald  Holton 
Fletcher  G.  Watson 


Contents 


TEXT  SECTION,  Unit  5 


Prologue  1 

Chapter  17    The  Chemical  Basis  of  Atomic  Theory 

Dal  ton's  atomic  theory  and  the  laws  of  chemical  combination     11 

The  atomic  masses  of  the  elements     14 

Other  properties  of  the  elements:   combining  capacity     16 

The  search  for  order  and  regularity  among  the  elements     18 

Mendeleev's  periodic  table  of  the  elements     19 

The  modern  periodic  table     23 

Electricity  and  Matter:  qualitative  studies     25 

Electricity  and  matter:  quantitative  studies     28 

Chapter  18     Electrons  and  Quanta 

The  idea  of  atomic  structure     33 

Cathode  rays     34 

The  measurement  of  the  charge  of  the  electron:  Millikan's  experiment     37 

The  photoelectric  effect     40 

Einstein's  theory  of  the  photoelectric  effect     43 

X  rays     48 

Electrons,  quanta  and  the  atom     54 

Chapter  19    The  Rutherford-Bohr  Model  of  the  Atom 

Spectra  of  gases     59 

Regularities  in  the  hydrogen  spectrum     63 

Rutherford's  nuclear  model  of  the  atom     66 

Nuclear  charge  and  size     69 

The  Bohr  theory :  the  postulates     71 

The  size  of  the  hydrogen  atom     72 

Other  consequences  of  the  Bohr  model     74 

The  Bohr  theory:  the  spectral  series  of  hydrogen     75 

Stationary  states  of  atoms:  the  Franck-Hertz  experiment     79 

The  periodic  table  of  the  elements     82 

The  inadequacy  of  the  Bohr  theory  and  the  state  of  atomic  theory  in  the  early  1920's     86 

Chapter  20     Some  Ideas  from  Modern  Physical  Theories 

Some  results  of  relativity  theory    95 
Particle-hke  behavior  of  radiation    99 
Wave-like  behavior  of  particles     101 
Mathematical  vs  visuahzable  atoms     104 
The  uncertainty  principle     108 
Probabihty  interpretation     111 

Epilogue    116 

Contents — HandbookSection     125 

I  ndex/Text  Section    1 59 

Index/HandbookSection     163 

Answers  to  End-of-Section  Questions     165 

Brief  Answers  to  Study  Guide  Questions     168 


.  *    ^*". 


UNIT 


5 


Models  of  the  Atom 


CHAPTERS 

17  The  Chemical  Basis  of  the  Atomic  Theory 

18  Electrons  and  Quanta 

19  The  Rutherford-Bohr  Model  of  the  Atom 

20  Some  Ideas  from  Modern  Physical  Theories 


PROLOGUE     In  the  earlier  units  of  this  course  we  studied  the  motion 
of  bodies:  bodies  of  ordinary  size,  such  as  we  deal  with  in  everyday  life, 
and  very  large  bodies,  such  as  planets.  We  have  seen  how  the  laws  of 
motion  and  gravitation  were  developed  over  many  centuries  and  how 
they  are  used.  We  have  learned  about  conservation  laws,  about  waves, 
about  light,  and  about  electric  and  magnetic  fields.  All  that  we  have 
learned  so  far  can  be  used  to  study  a  problem  which  has  intrigued 
people  for  many  centuries:  the  problem  of  the  nature  of  matter.  The 
phrase,  "the  nature  of  matter,"  may  seem  simple  to  us  now,  but  its 
meaning  has  been  changing  and  growing  over  the  centuries.  The  kind 
of  questions  and  the  methods  used  to  find  answers  to  these  questions 
are  continually  changing.  For  example,  during  the  nineteenth  century 
the  study  of  the  nature  of  matter  consisted  mainly  of  chemistry:  in  the 
twentieth  century  the  study  of  matter  has  also  moved  into  atomic  and 
nuclear  physics. 

Since  1800  progress  has  been  so  rapid  that  it  is  easy  to  forget  that 
people  have  theorized  about  matter  for  more  than  2,500  years.  In  fact 
some  of  the  questions  for  which  answers  have  been  found  only  during 
the  last  hundred  years  began  to  be  asked  more  than  two  thousand 
years  ago.  Some  of  the  ideas  we  consider  new  and  exciting,  such  as 
the  atomic  constitution  of  matter,  were  debated  in  Greece  in  the  fifth 
and  fourth  centuries  B.C.  In  this  prologue  we  shall  therefore  review 
briefly  the  development  of  ideas  concerning  the  nature  of  matter  up  to 
about  1800.  This  review  will  set  the  stage  for  the  four  chapters  of  Unit  5, 
which  will  be  devoted,  in  greater  detail,  to  the  progress  made  since 
1800  on  understanding  the  constitution  of  matter.  It  will  be  shown  in 
these  chapters  that  matter  is  made  up  of  discrete  particles  that  we  call 
atoms,  and  that  the  atoms  themselves  have  structure. 

Opposite:  Monolith— The  Face  of  Half  Dome  (Photo  by  Ansel  Adams) 


The  photographs  on  these  two 
pages  illustrate  some  of  the  variety 
of  forms  of  matter:  large  and  small, 
stable  and  shifting. 


microscopic  crystals 


condensed  water  vapor 


Greek  Ideas  of  Order 

The  Greek  mind  loved  clarity  and  order,  expressed  in 
a  way  that  still  touches  us  deeply.  In  philosophy,  litera- 
ture, art  and  architecture  it  sought  to  interpret  things  in 
terms  of  humane  and  lasting  qualities.  It  tried  to  discover 
the  forms  and  patterns  thought  to  be  essential  to  an 
understanding  of  things.  The  Greeks  delighted  in  show- 
ing these  forms  and  patterns  when  they  found  them.  Their 
art  and  architecture  express  beauty  and  intelligibility 
by  means  of  balance  of  form  and  simple  dignity. 

These  aspects  of  Greek  thought  are  beautifully  ex- 
pressed in  the  shrine  of  Delphi.  The  theater,  which  could 
seat  5,000  spectators,  impresses  us  because  of  the  size 
and  depth  of  the  tiered  seating  structure.  But  even  more 
striking  is  the  natural  and  orderly  way  in  which  the  theater 
is  shaped  into  the  landscape  so  that  the  entire  landscape 
takes  on  the  aspect  of  a  giant  theater.  The  Treasury  build- 
ing at  Delphi  has  an  orderly  system  of  proportions,  with 
form  and  function  integrated  into  a  logical,  pleasing 
whole.  The  statue  of  the  charioteer  found  at  Delphi,  with 
its  balance  and  firmness,  represents  a  genuine  ideal  of 
male  beauty  at  that  time.  After  more  than  2,000  years  we 
are  still  struck  by  the  elegance  of  Greek  expression. 


v^'K^ 


r^. 


^5><^> 


Prologue  3 

The  Roman  poet  Lucretius  based  his  ideas  of  physics  on  the 
tradition  of  atomism  dating  back  to  the  Greek  philosophers  Democritus 
and  Leucippus.  The  following  passages  are  from  his  poem  De  Rerum 
Natura  (On  the  Nature  of  Things),  an  eloquent  statement  of  atomism: 

...  If  you  think 

Atoms  can  stop  their  course,  refrain  from  movement, 

And  by  cessation  cause  new  kinds  of  motion, 

You  are  far  astray  indeed.  Since  there  is  void 

Through  which  they  move,  all  fundamental  motes 

Must  be  impelled,  either  by  their  own  weight 

Or  by  some  force  outside  them.  When  they  strike 

Each  other,  they  bounce  off;  no  wonder,  either. 

Since  they  are  absolute  solid,  all  compact. 

With  nothing  back  of  them  to  block  their  path. 

...  no  atom  ever  rests 

Coming  through  void,  but  always  drives,  is  driven 

In  various  ways,  and  their  collisions  cause. 

As  the  case  may  be,  greater  or  less  rebound. 

When  they  are  held  in  thickest  combination, 

At  closer  intervals,  with  the  space  between 

More  hindered  by  their  interlock  of  figure. 

These  give  us  rock,  or  adamant,  or  iron. 

Things  of  that  nature.  (Not  very  many  kinds 

Go  wandering  little  and  lonely  through  the  void.) 

There  are  some  whose  alternate  meetings,  partings,  are 

At  greater  intervals;  from  these  we  are  given 

Thin  air,  the  shining  sunlight  .  .  . 

*  *  * 

. . .  It's  no  wonder 

That  while  the  atoms  are  in  constant  motion, 

Their  total  seems  to  be  at  total  rest, 

Save  here  and  there  some  individual  stir. 

Their  nature  lies  beyond  our  range  of  sense. 

Far,  far  beyond.  Since  you  can't  get  to  see 

The  things  themselves,  they're  bound  to  hide  their  moves, 

Especially  since  things  we  can  see,  often 

Conceal  their  movements,  too,  when  at  a  distance. 

Take  grazing  sheep  on  a  hill,  you  know  they  move, 

The  woolly  creatures,  to  crop  the  lovely  grass 

Wherever  it  may  call  each  one,  with  dew 

Still  sparkling  it  with  jewels,  and  the  lambs. 

Fed  full,  play  little  games,  flash  in  the  sunlight. 

Yet  all  this,  far  away,  is  just  a  blue, 

A  whiteness  resting  on  a  hill  of  green. 

Or  when  great  armies  sweep  across  great  plains 

In  mimic  warfare,  and  their  shining  goes 

Up  to  the  sky,  and  all  the  world  around 

Is  brilliant  with  their  bronze,  and  trampled  earth 

Trembles  under  the  cadence  of  their  tread, 

White  mountains  echo  the  uproar  to  the  stars, 

The  horsemen  gallop  and  shake  the  very  ground, 

And  yet  high  in  the  hills  there  is  a  place 

From  which  the  watcher  sees  a  host  at  rest. 

And  only  a  brightness  resting  on  the  plain. 

[translated  from  the  Latin  by  Rolfe  Humphries] 


Models  of  the  Atom 


This  gold  earring,  made  in  Greece 
about  600  B.C.,  shows  the  great  skill 
with  which  ancient  artisans  worked 
metals.  [Museum  of  Fine  Arts,  Boston] 


Early  science  had  to  develop  out  of  the  ideas  available  before 
science  started— ideas  that  came  from  experience  with  snow,  wind, 
rain,  mist  and  clouds;  with  heat  and  cold;  with  salt  and  fresh  water; 
wine,  milk,  blood,  and  honey;  ripe  and  unripe  fruit;  fertile  and  infertile 
seeds.  The  most  obvious  and  most  puzzling  facts  were  that  plants, 
animals,  and  men  were  born,  that  they  grew  and  matured,  and  that  they 
aged  and  died.  Men  noticed  that  the  world  about  them  was  continually 
changing  and  yet,  on  the  whole,  it  seemed  to  remain  much  the  same. 
The  unknown  causes  of  these  changes  and  of  the  apparent  continuity 
of  nature  were  assigned  to  the  actions  of  gods  and  demons  who  were 
thought  to  control  nature.  Myths  concerning  the  creation  of  the  world 
and  the  changes  of  the  seasons  were  among  the  earliest  creative 
productions  of  primitive  peoples  everywhere,  and  helped  them  to  come 
to  terms  with  events  man  could  see  happening  but  could  not  rationally 
understand. 

Over  a  long  period  of  time  men  developed  some  control  over  nature 
and  materials:  they  learned  how  to  keep  warm  and  dry,  to  smelt  ores,  to 
make  weapons  and  tools,  to  produce  gold  ornaments,  glass,  perfumes, 
and  medicines.  Eventually,  in  Greece,  by  the  year  600  B.C.,  philosophers 
—literally  "lovers  of  wisdom"— had  started  to  look  for  rational  explana- 
tions of  natural  events,  that  is,  explanations  that  did  not  depend  on  the 
actions  or  the  whims  of  gods  or  demons.  They  sought  to  discover  the 
enduring,  unchanging  things  out  of  which  the  world  is  made,  and  how 
these  enduring  things  can  give  rise  to  the  changes  that  we  perceive, 
as  well  as  the  great  variety  of  material  things  that  exists.  This  was  the 
beginning  of  man's  attempts  to  understand  the  material  world  rationally, 
and  it  led  to  a  theory  of  the  nature  of  matter. 

The  earliest  Greek  philosophers  thought  that  all  the  different  things 
in  the  world  were  made  out  of  a  single  basic  substance.  Some  thought 
that  water  was  the  fundamental  substance  and  that  all  other  substances 
were  derived  from  it.  Others  thought  that  air  was  the  basic  substance; 
still  others  favored  fire.  But  neither  water,  nor  air,  nor  fire  was  satis- 
factory; no  one  substance  seemed  to  have  enough  different  properties 
to  give  rise  to  the  enormous  variety  of  substances  in  the  world.  According 
to  another  view,  introduced  by  Empedocles  around  450  B.C.,  there  are 
four  basic  types  of  matter— earth,  air,  fire,  and  water— and  all  material 
things  are  made  out  of  them.  These  four  basic  materials  can  mingle 
and  separate  and  reunite  in  different  proportions,  and  so  produce 
the  variety  of  familiar  objects  around  us  as  well  as  the  changes  in 
such  objects.  But  the  basic  four  materials,  called  elements,  were 
supposed  to  persist  through  all  these  changes.  This  theory  was  the 
first  appearance  in  our  scientific  tradition  of  a  model  of  matter, 
according  to  which  all  material  things  are  just  different  arrangements 
of  a  few  external  elements. 

The  first  atomic  theory  of  matter  was  introduced  by  the  Greek 
philosopher  Leucippus,  born  about  500  B.C.,  and  his  pupil  Democritus, 
who  lived  from  about  460  B.C.  to  370  B.C.  Only  scattered  fragments  of 
the  writings  of  these  philosophers  remain,  but  their  ideas  were  dis- 
cussed in  considerable  detail  by  the  Greek  philosophers  Aristotle 
(389-321  B.C.)  and  Epicurus  (341-270  B.C.),  and  by  the  Latin  poet 


Prologue 


Lucretius  (100-55  B.C.).  It  is  to  these  men  that  we  owe  most  of  our 
knowledge  of  ancient  atomism. 

The  theory  of  the  atomists  was  based  on  a  number  of  assumptions: 

(1)  matter  is  eternal— no  material  thing  can  come  from  nothing, 
nor  can  any  material  thing  pass  into  nothing; 

(2)  material  things  consist  of  very  small  indivisible  particles— the 
word  "atom"  meant  "uncuttable"  in  Greek  and,  in  discussing  the  ideas 
of  the  early  atomists,  we  could  use  the  word  "indivisibles"  instead  of 
the  word  "atoms"; 

(3)  atoms  differ  chiefly  in  their  sizes  and  shapes; 

(4)  the  atoms  exist  in  otherwise  empty  space  (the  void)  which  sepa- 
rates them,  and  because  of  this  space  they  are  capable  of  movement 
from  one  place  to  another; 

(5)  the  atoms  are  in  ceaseless  motion,  although  the  nature  and 
cause  of  the  motion  are  not  clear; 

(6)  in  the  course  of  their  motions  atoms  come  together  and  form 
combinations  which  are  the  material  substances  we  know;  when  the 
atoms  forming  these  combinations  separate,  the  substances  decay  or 
break  up.  Thus,  the  combinations  and  separations  of  atoms  give  rise  to 
the  changes  which  take  place  in  the  world; 

(7)  the  combinations  and  separations  take  place  in  accord  with 
natural  laws  which  are  not  yet  clear,  but  do  not  require  the  action  of 
gods  or  demons  or  other  supernatural  powers. 

With  the  above  assumptions,  the  ancient  atomists  were  able  to 
work  out  a  consistent  story  of  change,  of  what  they  sometimes  called 
"coming-to-be"  and  "passing  away."  They  could  not  demonstrate 
experimentally  that  their  theory  was  correct,  and  they  had  to  be  satis- 
fied with  an  explanation  derived  from  assumptions  that  seemed 
reasonable  to  them.  The  theory  was  a  "likely  story."  It  was  not 
useful  for  the  prediction  of  new  phenomena;  but  that  became  an 
important  value  for  a  theory  only  later.  To  these  atomists,  it  was  more 
significant  that  the  theory  also  helped  to  allay  the  unreasonable  fear 
of  capricious  gods. 

The  atomic  theory  was  criticized  severely  by  Aristotle,  who  argued 
logically— from  his  own  assumptions— that  no  vacuum  or  void  could 
exist  and  that  the  ideas  of  atoms  with  their  continual  motion  must  be 
rejected.  (Aristotle  was  also  probably  sensitive  to  the  fact  that  in  his 
time  atomism  was  identified  with  atheism.)  For  a  long  time  Aristotle's 
argument  against  the  void  was  widely  held  to  be  convincing.  One  must 
here  recall  that  not  until  the  seventeenth  century  did  Torricelli's 
experiments  (described  in  Chapter  1 1 )  show  that  a  vacuum  could  indeed 
exist.  Furthermore,  Aristotle  argued  that  matter  is  continuous  and 
infinitely  divisible  so  that  there  can  be  no  atoms. 

Aristotle  developed  a  theory  of  matter  as  part  of  his  grand  scheme 
of  the  universe,  and  this  theory,  with  some  modifications,  was  thought 
to  be  satisfactory  by  most  philosophers  of  nature  for  nearly  two 
thousand  years.  His  theory  of  matter  was  based  on  the  four  basic 
elements.  Earth,  Air,  Fire,  and  Water,  and  four  "qualities,"  Cold,  Hot, 
Moist,  and  Dry.  Each  element  was  characterized  by  two  qualities  (the 


According  to  Aristotle  in  his  Meta- 
physics, "There  is  no  consensus 
concerning  the  number  or  nature  of 
these  fundamental  substances. 
Thales,  the  first  to  think  about  such 
matters,  held  that  the  elementary 
substance  is  clear  liquid.  ...  He 
may  have  gotten  this  idea  from  the 
observation  that  only  moist  matter 
can  be  wholly  integrated  into  an 
object — so  that  all  growth  depends 
on  moisture.  .  .  . 

"Anaximenes  and  Diogenes  held 
that  colorless  gas  is  more  elemen- 
tary than  clear  liquid,  and  that  in- 
deed, it  is  the  most  elementary  of 
all  simple  substances.  On  the  other 
hand  Hippasus  of  Metpontum  and 
Heraclitus  of  Ephesus  said  that  the 
most  elementary  substance  is  heat. 
Empedocles  spoke  of  four  elemen- 
tary substances,  adding  dry  dust  to 
the  three  already  mentioned  .  .  . 
Anaxagoras  of  Clazomenae  says 
that  there  are  an  infinite  number  of 
elementary  constituents  of  mat- 
ter. .  .  ."  [From  a  translation  by 
D.  E.  Gershenson  and  D.  A.  Green- 
berg.] 


6 


Models  of  the  Atom 


FIRE 


WATER 


Laboratory 
chemist. 


of     a     16th-century     al- 


nearer  two  to  each  side,  as  shown  in  the  diagram  at  the  left).  Thus 
the  element 

Earth  is  Dry  and  Cold, 

Water  is  Cold  and  Moist, 

Air  is  Moist  and  Hot, 

Fire  is  Hot  and  Dry. 

According  to  Aristotle,  it  is  always  the  first  of  the  two  qualities  which 
predominates.  In  his  version  the  elements  are  not  unchangeable;  any 
one  of  them  may  be  transformed  into  any  other  because  of  one  or  both 
of  its  qualities  changing  into  opposites.  The  transformation  takes  place 
most  easily  between  two  elements  having  one  quality  in  common;  thus 
Earth  is  transformed  into  Water  when  dryness  changes  into  moistness. 
Earth  can  be  transformed  into  Air  only  if  both  of  the  qualities  of  earth 
(dry  and  cold)  are  changed  into  their  opposites  (moist  and  hot). 

As  we  have  already  mentioned  in  the  Text  Chapter  2,  Aristotle  was 
able  to  explain  many  natural  phenomena  by  means  of  his  ideas.  Like 
the  atomic  theory,  Aristotle's  theory  of  coming-to-be  and  passing-away 
was  consistent,  and  constituted  a  model  of  the  nature  of  matter.  It  had 
certain  advantages  over  the  atomic  theory:  it  was  based  on  elements 
and  qualities  that  were  familiar  to  people;  it  did  not  involve  atoms, 
which  couldn't  be  seen  or  otherwise  perceived,  or  a  void,  which  was 
most  difficult  to  imagine.  In  addition,  Aristotle's  theory  provided  some 
basis  for  further  experimentation:  it  supplied  what  seemed  like  a 
rational  basis  for  the  tantalizing  possibility  of  changing  any  material 
into  any  other. 

Although  the  atomistic  view  was  not  altogether  abandoned,  it  found 
few  supporters  during  the  period  300  A.D.  to  about  1600  A.D.  The  atoms 
of  Leucippus  and  Democritus  moved  through  empty  space,  devoid  of 
spirit,  and  with  no  definite  plan  or  purpose.  Such  an  idea  remained 
contrary  to  the  beliefs  of  the  major  religions.  Just  as  the  Athenians  did 
in  the  time  of  Plato  and  Aristotle,  the  later  Christian,  Hebrew,  and 
Moslem  theologians  considered  atomists  to  be  atheistic  and  "mate- 
rialistic" because  they  claimed  that  everything  in  the  universe  can  be 
explained  in  terms  of  matter  and  motion. 

About  300  or  400  years  after  Aristotle,  a  kind  of  research  called 
alchemy  appeared  in  the  Near  and  Far  East.  Alchemy  in  the  Near  East 
was  a  combination  of  Aristotle's  ideas  about  matter  with  methods  of 
treating  ores  and  metals.  One  of  the  aims  of  the  alchemists  was  to 
change,  or  "transmute"  ordinary  metals  into  precious  metals.  Although 
they  failed  to  do  this,  the  alchemists  found  and  studied  many  of  the 
properties  of  substances  that  are  now  classified  as  chemical  properties. 
They  invented  some  pieces  of  chemical  apparatus,  such  as  reaction 
vessels  and  distillation  flasks,  that  (in  modern  form)  are  still  common 
in  chemical  laboratories.  They  studied  such  processes  as  calcination, 
distillation,  fermentation,  and  sublimation.  In  this  sense  alchemy  may 
be  regarded  as  the  chemistry  of  the  Middle  Ages.  But  alchemy  left 
unsolved  the  fundamental  questions.  At  the  opening  of  the  eighteenth 
century  the  most  important  of  these  questions  were:  (1)  what  is  a 
chemical  element;  (2)  what  is  the  nature  of  chemical  composition  and 
chemical  change,  especially  burning;  and  (3)  what  is  the  chemical 


Prologue 


nature  of  the  so-called  elements,  Earth,  Air,  Fire  and  Water.  Until  these 
questions  were  answered,  it  was  impossible  to  make  real  progress  in 
finding  out  the  structure  of  matter.  One  result  was  that  the  "scientific 
revolution"  of  the  seventeenth  century,  which  clarified  the  problems  of 
astronomy  and  mechanics,  did  not  include  chemistry. 

During  the  seventeenth  century,  however,  some  forward  steps  were 
made  which  supplied  a  basis  for  future  progress  on  the  problem  of 
matter.  The  Copernican  and  Newtonian  revolutions  undermined  the 
authority  of  Aristotle  to  such  an  extent  that  his  ideas  about  matter 
were  also  more  easily  questioned.  Atomic  concepts  were  revived,  and 
offered  a  way  of  looking  at  things  that  was  very  different  from  Aristotle's 
ideas.  As  a  result,  theories  involving  atoms  (or  "particles"  or  "corpus- 
cles") were  again  considered  seriously.  Boyle's  models  were  based  on 
the  idea  of  "gas  particles."  Newton  also  discussed  the  behavior  of  a 
gas  (and  even  of  light)  by  supposing  it  to  consist  of  particles.  In 
addition,  there  was  now  a  successful  science  of  mechanics,  through 
which  one  might  hope  to  describe  how  the  atoms  interacted  with  each 
other.  Thus  the  stage  was  set  for  a  general  revival  of  atomic  theory. 

In  the  eighteenth  century,  chemistry  became  more  quantitative; 
weighing  in  particular  was  done  more  frequently  and  more  carefully. 
New  substances  were  isolated  and  their  properties  examined.  The 
attitude  that  grew  up  in  the  latter  half  of  the  century  was  exemplified  by 
that  of  Henry  Cavendish  (1731-1810),  who,  according  to  a  biographer, 
regarded  the  universe  as  consisting 


One  of  those  who  contributed 
greatly  to  the  revival  of  atomism 
was  Pierre  Gassendi  (1592 — 1655),  a 
French  priest  and  philosopher.  He 
avoided  the  criticism  of  atomism 
as  atheistic  by  saying  that  God  also 
created  the  atoms  and  bestowed 
motion  upon  them.  Gassendi  ac- 
cepted the  physical  explanations  of 
the  atomists,  but  rejected  their  dis- 
belief in  the  immortality  of  the  soul 
and  in  Divine  Providence.  He  was 
thus  able  to  provide  a  philosophical 
justification  of  atomism  which  met 
some  of  the  serious  religious 
objections. 


.  .  .  solely  of  a  multitude  of  objects  which  could  be  weighed, 
numbered,  and  measured;  and  the  vocation  to  which  he  con- 
sidered himself  called  was  to  weigh,  number,  and  measure 
as  many  of  those  objects  as  his  alloted  threescore  years  and 
ten  would  permit.  ...  He  weighed  the  Earth;  he  analysed  the 
Air;  he  discovered  the  compound  nature  of  Water;  he  noted 
with  numerical  precision  the  obscure  actions  of  the  ancient 
element  Fire. 


It  was  Cavendish,  remember,  who 
designed  the  sensitive  torsional 
balance  that  made  it  possible  to 
find  a  value  for  the  gravitational 
constant  G.  (Text  Sec.  8.8.) 


Eighteenth-century  chemistry  reached  its  peak  in  the  work  of 
Antoine  Lavoisier  (1743-1794),  who  worked  out  the  modern  views  of 
combustion,  established  the  law  of  conservation  of  mass,  explained  the 
elementary  nature  of  hydrogen  and  oxygen,  and  the  composition  of 
water,  and  above  all  emphasized  the  quantitative  aspects  of  chemistry. 
His  famous  book,  Traite  Elementaire  de  Chimie  (or  Elements  of 
Chemistry),  published  in  1789,  established  chemistry  as  a  modern 
science.  In  it,  he  analyzed  the  idea  of  an  element  in  a  way  which  is  very 
close  to  our  modern  views: 

...  if,  by  the  term  elements  we  mean  to  express  those  simple 
and  indivisible  atoms  of  which  matter  is  composed,  it  is  ex- 
tremely probable  that  we  know  nothing  at  all  about  them;  but 
if  we  apply  the  term  elements,  or  principles  of  bodies,  to 
express  our  idea  of  the  last  point  which  analysis  is  capable 
of  reaching,  we  must  admit  as  elements  all  the  substances 
into  which  we  are  capable,  by  any  means,  to  reduce  bodies 
by  decomposition.  Not  that  we  are  entitled  to  affirm  that 


Lavoisier's  work  on  the  conserva- 
tion of  mass  was  described  in  Text 
Chapter  9. 


Models  of  the  Atom 


T  R  A  I  T  E 

ELEMENTAIRE 

D  E   CHI  MIE, 

PRfeSENTt  DANS  UN  ORDRE  NOUVEAU 

ET  d'aPR^S  LES  D^COUVERTES  UODERNES} 

Avec  Figures  : 

Tar  M.  Lavo  ist  EA  ,  de  CAcaJimU  dit 
Sc'uncts,  de  la  Socieii  RoyaU  de  Medccme ,  dtt 
Socieus  d' Agriculture  de  Paris  O  d'OrUan.s  ,  de 
la  Societe  RoyaU  de  Londres  ,  de  I'lnftiiut  de 
Bologiie  ,  de  la  Societe  Helvitique  de  Bajle  ,  dt 
celtes  de  PhUadelphle ,  Harlem  ,  Manchefler  , 
Padoue  ,  &c. 


|. 


TOME    PREMIER. 


A     PARIS, 

Ch«  CuCHET,  Libraire,  rue  &  hotel  Serpente. 

M.    D  C  C.    L  X  X  X  I  X. 

Sma  It  PriviUgt  de  TAcaidrnit  dtt  Scieru-ei  6  dt  U 
SociM  RoyaU  dt  Midteint 

Title  page  of  Lavoisier's  Iratte  Ele- 
mentaire  de  Chimie  (1789) 


these  substances  we  consider  as  simple  may  not  be  com- 
pounded of  two,  or  even  of  a  greater  number  of  principles; 
but  since  these  principles  cannot  be  separated,  or  rather 
since  we  have  not  hitherto  discovered  the  means  of  sepa- 
rating them,  they  act  with  regard  to  us  as  simple  substances, 
and  we  ought  never  to  suppose  them  compounded  until  ex- 
periment and  observation  have  proved  them  to  be  so. 

During  the  latter  half  of  the  eighteenth  century  and  the  early  years  of 
the  nineteenth  century  great  progress  was  made  in  chemistry  because 
of  the  increasing  use  of  quantitative  methods.  Chemists  found  out  more 
and  more  about  the  composition  of  substances.  They  separated  many 
elements  and  showed  that  nearly  all  substances  are  compounds— 
combinations  of  a  fairly  small  number  of  chemical  elements.  They 
learned  a  great  deal  about  how  elements  combine  to  form  compounds 
and  how  compounds  can  be  broken  down  into  the  elements  of  which 
they  are  composed.  This  information  made  it  possible  for  chemists  to 
establish  many  empirical  laws  of  chemical  combination.  Then  chemists 
sought  an  explanation  for  these  laws. 

During  the  first  ten  years  of  the  nineteenth  century,  the  English 
chemist  John  Dalton  introduced  a  modified  form  of  the  old  Greek 
atomic  theory  to  account  for  the  laws  of  chemical  combination.  It  is 
here  that  the  modern  story  of  the  atom  begins.  Dalton's  atomic  theory 
was  an  improvement  over  that  of  the  Greeks  because  it  opened  the 
way  for  the  quantitative  study  of  the  atom  in  the  nineteenth  century. 
Today  the  existence  of  the  atom  is  no  longer  a  topic  of  speculation. 
There  are  many  kinds  of  experimental  evidence,  not  only  for  the 
existence  of  atoms  but  also  for  their  inner  structure.  In  this  unit  we 
shall  trace  the  discoveries  and  ideas  that  provided  this  evidence. 

The  first  convincing  modern  idea  of  the  atom  came  from  chemistry. 
We  shall,  therefore,  start  with  chemistry  in  the  early  years  of  the  nine- 
teenth century;  this  is  the  subject  of  Chapter  17.  Then  we  shall  see  that 
chemistry  raised  certain  questions  about  atoms  which  could  only  be 
answered  by  physics.  Physical  evidence,  accumulated  in  the  nineteenth 
century  and  the  early  years  of  the  twentieth  century,  made  it  possible 
to  propose  models  for  the  structure  of  atoms.  This  evidence  will  be 
discussed  in  Chapters  18  and  19.  Some  of  the  latest  ideas  about  atomic 
theory  will  then  be  discussed  in  Chapter  20. 


Chemical  laboratory  of  the  18th  century 


17.1  Dalton's  atomic  theory  and  the  laws  of  chemical  combination      11 

17.2  The  atomic  masses  of  the  elements  14 

17.3  Other  properties  of  the  elements:  combining  capacity  16 

17.4  The  search  for  order  and  regularity  among  the  elements  18 

17.5  Mendeleev's  periodic  table  of  the  elements  19 

17.6  The  modern  periodic  table  23 

17.7  Electricity  and  matter:  qualitative  studies  25 

17.8  Electricity  and  matter:  quantitative  studies  28 


OCD^O® 

oo®®© 


©®©®o 


Dalton's  symbols  for  'elements  "  (1808) 


CHAPTER  SEVENTEEN 


The  Chemical  Basis  of 
Atomic  Theory 


17.1     Dalton's  atomic  theory  and  the  laws  of  chemical  combination 

The  atomic  theory  of  John  Dalton  appeared  in  his  treatise,  A 
New  System  of  Chemical  Philosophy,  published  in  two  parts,  in 
1808  and  1810.  The  main  postulates  of  his  theory  were: 

(1)  Matter  consists  of  indivisible  atoms. 

.  .  .  matter,  though  divisible  in  an  extreme  degree,  is 
nevertheless  not  infinitely  divisible.  That  is,  there  must 
be  some  point  beyond  which  we  cannot  go  in  the  division 
of  matter.  The  existence  of  these  ultimate  particles  of 
matter  can  scarcely  be  doubted,  though  they  are  probably 
much  too  small  ever  to  be  exhibited  by  microscopic  im- 
provements. I  have  chosen  the  word  atom  to  signify  these 
ultimate  particles.  .  .  . 

(2)  Each  element  consists  of  a  characteristic  kind  of  identical 
atoms.  There  are  consequently  as  many  different  kinds  of  atoms  as 
there  are  elements.  The  atoms  of  an  element  "are  perfectly  alike  in 
weight  and  figure,  etc." 

(3)  Atoms  are  unchangeable. 

(4)  When  different  elements  combine  to  form  a  compound,  the 
smallest  portion  of  the  compound  consists  of  a  grouping  of  a  definite 
number  of  atoms  of  each  element. 

(5)  In  chemical  reactions,  atoms  are  neither  created  nor 
destroyed,  but  only  rearranged. 

Dalton's  theory  really  grew  out  of  his  interest  in  meteorology 
and  his  research  on  the  composition  of  the  atmosphere.  He  tried  to 
explain  many  of  the  physical  properties  of  gases  in  terms  of  atoms 
(for  example,  the  fact  that  gases  readily  mix,  and  the  fact  that  the 
pressures  of  two  gases  add  simply  when  both  are  combined  in  a 
fixed  enclosure).  He  thought  of  the  atoms  of  different  elements  as 
being  different  in  size  and  in  mass.  In  keeping  with  the  quantitative 
spirit  of  the  time,  he  tried  to  determine  the  numerical  values  for  their 
relative  masses.  This  was  a  crucial  step  forward.  But  before  consider- 
ing how  to  determine  the  relative  masses  of  atoms  of  the  different 
elements,  let  us  see  how  Dalton's  postulates  make  it  possible  to  ac- 
count for  the  experimentally  known  laws  of  chemical  combination. 

11 


SG  17.1 


Meteorology  is  a  science  that  deals 
with  the  atmosphere  and  its 
phenomena — weather  forecasting 
is  one  branch  of  meteorology. 


12 


The  Chemical  Basis  of  the  Atomic  Theory 


Recall  that  empirical  laws  (such  as 
these,  or  Kepler's  laws  of  planetary 
motion)  are  just  summaries  of 
experimentally  observed  facts.  They 
cry  out  for  some  theoretical  base 
from  which  they  can  be  shown  to 
follow  as  necessary  consequences. 
Physical  science  looks  for  these 
deeper  necessities  that  describe 
nature,  and  is  not  satisfied  with 
mere  summaries  of  observation, 
useful  though  these  may  be  initially. 


Dalton's  atomic  theory  accounts  in  a  simple  and  direct  way  for 
the  law  of  conservation  of  mass.  According  to  Dalton's  theory 
(postulates  4  and  5),  chemical  changes  are  only  the  rearrangements 
of  unions  of  atoms.  Since  atoms  are  unchangeable  (according  to 
postulate  3)  rearranging  them  cannot  change  their  masses.  Hence, 
the  total  mass  of  all  the  atoms  before  the  reaction  must  equal  the 
total  mass  of  all  the  atoms  after  the  reaction. 

Another  well  known  empirical  law  which  could  be  explained 
easily  with  Dalton's  theory  is  the  law  of  definite  proportions.  This 
law  states  that  any  particular  chemical  compound  always  contains 
the  same  elements,  and  they  are  united  in  the  same  proportions  of 
weight.  For  example,  the  ratio  of  the  masses  of  oxygen  and  hy- 
drogen which  combine  to  form  water  is  always  7.94  to  1 : 

mass  of  oxygen    _  7.94 
mass  of  hydrogen         1 

If  there  is  more  of  one  element  present  than  is  needed  for  full 
combination  in  a  chemical  reaction,  say  10  grams  of  oxygen  and 
one  gram  of  hydrogen,  only  7.94  grams  of  oxygen  will  combine 
with  the  hydrogen.  The  rest  of  the  oxygen,  2.06  grams,  remains 
uncombined. 

The  fact  that  elements  combine  in  fixed  proportions  implies  that 
each  chemical  compound  will  also  decompose  into  definite  propor- 
tions of  elements.  For  example,  the  decomposition  of  sodium 
chloride  (common  salt)  always  gives  the  results:  39  percent 
sodium  and  61  percent  chlorine  by  weight. 

Now  let  us  see  how  Dalton's  model  can  be  applied  to  a  chemical 
reaction,  say,  to  the  formation  of  water  from  oxygen  and  hydrogen. 
According  to  Dalton's  second  postulate,  all  the  atoms  of  oxygen 
have  the  same  mass;  and  all  the  atoms  of  hydrogen  have  the  same 
mass,  which  is  different  from  the  mass  of  the  oxygen  atoms.  To 
express  the  total  mass  of  oxygen  entering  into  the  reaction,  we 
multiply  the  mass  of  a  single  oxygen  atom  by  the  number  of  oxygen 
atoms: 


SG  17.2, 17.3 


„  /     mass  of     \ 

mass  of  oxygen  =  (oxygen  atom  ) 


number  of    ^ 
oxygen  atoms, 


Similarly,  for  the  total  mass  of  hydrogen  entering  into  the  reaction: 

„ ,     ,  /       mass  of       \      /      number  of      \ 

mass  of  hydrogen  -  (^hydrogen  atom]  ^  \hydrogen  atoms) 

We  can  find  the  ratio  of  the  mass  of  oxygen  to  the  mass  of  hydrogen 
by  dividing  the  first  equation  by  the  second  equation  as  shown  at 
the  top  of  the  next  page: 


Section  17.1 


13 


mass  of 
mass  of  oxygen    _    oxygen  atom 


mass  of  hydrogen 


mass  of 
hydrogen  atom 


number  of 

oxygen  atoms 

number  of 
hydrogen  atoms 


If  the  masses  of  the  atoms  do  not  change  (postulate  3),  the  first 
ratio  on  the  right  side  of  the  equation  has  a  certain  unchangeable 
value.  According  to  postulate  4,  if  the  smallest  portion  of  the  com- 
pound water  consists  of  a  definite  number  of  atoms  of  each  element 
(postulate  4),  the  second  ratio  on  the  right  side  of  the  equation  has 
a  certain  unchangeable  value  also.  The  product  of  the  two  ratios  on 
the  right  side  will  always  have  the  same  value.  This  equation, 
based  on  an  atomic  theory,  thus  tells  us  that  the  ratio  of  the 
masses  of  oxygen  and  hydrogen  that  combine  to  form  water  will 
always  have  the  same  definite  value.  But  this  is  just  what  the 
experimental  law  of  definite  proportions  says.  Dalton's  theory 
accounts  for  this  law  of  chemical  combination— and  this  success 
tends  to  confirm  Dalton's  conception.  Dalton's  theory  was  also 
consistent  with  another  empirical  law  of  chemical  combination,  the 
law  of  multiple  proportion.  For  some  combinations  of  elements 
there  are  a  set  of  possible  values  for  their  proportions  in  forming  a 


SG  17.4 


^J^3r^ 


A  page  from  Dalton's  notebook, 
showing  his  representation  of  two 
adjacent  atoms  (top)  and  of  a  mole- 
cule or    compound  atom'  (bottom) 


John  Dalton  (1766-1844).  His  first 
love  was  meteorology,  and  he  kept 
careful  daily  weather  records  for 
46  years— a  total  of  200,000  observa- 
tions. He  was  the  first  to  describe 
color  blindness  in  a  publication  and 
was  color-blind  himself,  not  exactly 
an  advantage  for  a  chemist  who  had 
to  see  color  changes  in  chemicals. 
(His  color  blindness  may  help  to 
explain  why  Dalton  is  said  to  have 
been  a  rather  clumsy  experimenter.) 
However,  his  accomplishments  rest 
not  on  successful  experiments,  but 
on  his  ingenious  interpretation  of 
the  known  results  of  others.  Dalton's 
notion  that  all  elements  were  com- 
posed of  extremely  tiny,  indivisible 
and  indestructible  atoms,  and  that 
all  substances  are  composed  of 
combinations  of  these  atoms  was 
accepted  soon  by  most  chemists 
with  surprisingly  little  opposition. 
There  were  many  attempts  to  honor 
him,  but  being  a  Quaker  he  shunned 
any  form  of  glory.  When  he  received 
a  doctor's  degree  from  Oxford,  his 
colleagues  wanted  to  present  him  to 
King  William  IV.  He  had  always 
resisted  such  a  presentation  be- 
cause he  would  not  wear  court 
dress.  However,  his  Oxford  robes 
satisfied  the  protocol. 


14 


The  Chemical  Basis  of  the  Atomic  Theory 


Dalton's    visualization    of    the    com- 
position of  various  compounds. 


set  of  compounds.  Dalton  showed  that  these  cases  could  all  be 
accounted  for  by  different  combinations  of  whole  numbers  of  atoms. 

There  are  other  laws  of  chemical  combination  which  are 
explained  by  Dalton's  theory.  Because  the  argument  would  be 
lengthy  and  relatively  little  that  is  new  would  be  added,  we  shall  not 
elaborate  on  them  here. 

Dalton's  interpretation  of  the  experimental  facts  of  chemical 
combination  made  possible  several  important  conclusions:  (1)  that 
the  difference  between  one  chemical  element  and  another  would 
have  to  be  described  in  terms  of  the  differences  between  the  atoms 
of  which  these  elements  were  made  up;  (2)  that  there  were,  there- 
fore, as  many  different  types  of  atoms  as  there  were  chemical 
elements;  (3)  that  chemical  combination  was  the  union  of  atoms  of 
different  elements  into  molecules  of  compounds.  Dalton's  theory 
also  implied  that  the  analysis  of  a  large  number  of  chemical  com- 
pounds could  make  it  possible  to  assign  relative  mass  values  to 
the  atoms  of  different  elements.  This  possibility  will  be  discussed 
in  the  next  section. 

Q1    What  did  Dalton  assume  about  the  atoms  of  an  element? 
Q2    What  two  experimental  laws  did  Dalton's  theory  explain? 
What  follows  from  these  successes? 


17.2    The  atomic  masses  of  the  elements 


The  first  good  estimates  of 
molecular  size  came  from  the  kinetic 
theory  of  gases  and  indicated  that 
atoms  (or  molecules)  had  diameters 
of  the  order  of  10'"  meter.  Atoms 
are  thus  much  too  small  for  ordinary 
mass  measurements  to  be  made  on 
single  atoms. 


SG  17.5 
SG  17.6 


One  of  the  most  important  concepts  to  come  from  Dalton's  work 
is  that  of  atomic  mass  and  the  possibility  of  determining  numerical 
values  for  the  masses  of  the  atoms  of  different  elements.  Dalton 
had  no  idea  of  the  actual  absolute  mass  of  individual  atoms. 
Reasonably  good  estimates  of  the  size  of  atoms  did  not  appear  until 
about  50  years  after  Dalton  published  his  theory.  Nevertheless,  as 
Dalton  was  able  to  show,  relative  values  of  atomic  masses  can  be 
found  by  using  the  law  of  definite  proportions  and  experimental 
data  on  chemical  reactions. 

To  see  how  this  could  be  done  we  return  to  the  case  of  water, 
for  which,  the  ratio  of  the  mass  of  oxygen  to  the  mass  of  hydrogen 
is  found  by  experiment  to  be  7.94:1.  If  one  knew  how  many  atoms 
of  oxygen  and  hydrogen  are  contained  in  a  molecule  of  water  one 
could  calculate  the  ratio  of  the  mass  of  the  oxygen  atom  to  the  mass 
of  the  hydrogen  atom.  But  Dalton  didn't  know  the  numbers  of 
oxygen  and  hydrogen  atoms  in  a  molecule  of  water  so  he  made  an 
assumption.  As  is  done  often,  Dalton  made  the  simplest  possible 
assumption:  that  a  molecule  of  water  consists  of  one  atom  of 
oxygen  combined  with  one  atom  of  hydrogen.  By  this  reasoning 
Dalton  concluded  that  the  oxygen  atom  is  7.94  times  more  massive 
than  the  hydrogen  atom.  Actually,  the  simplest  assumption  proved 
in  this  case  to  be  incorrect:  two  atoms  of  hydrogen  combine  with 
one  atom  of  oxygen  to  make  a  molecule  of  water.  The  oxygen  atom 
has  7.94  times  the  mass  of  the  two  hydrogen  atoms,  and  therefore 
has  15.88  times  the  mass  of  a  single  hydrogen  atom. 

More  generally,  Dalton  assumed  that  when  only  one  compound 


Section  17.2 


15 


of  any  two  elements  is  known  to  exist,  molecules  of  the  compound 
always  consist  of  one  atom  of  each.  With  this  assumption  Dalton 
could  find  values  for  the  relative  masses  of  different  atoms — but 
later  work  showed  that  Dalton's  assumption  of  one-to-one  ratios  was 
often  as  incorrect  as  it  was  for  water.  By  studying  the  composition 
of  water  as  well  as  many  other  chemical  compounds,  Dalton  found 
that  the  hydrogen  atom  appeared  to  have  a  smaller  mass  than  the 
atoms  of  any  other  element.  Therefore,  he  proposed  to  express  the 
masses  of  atoms  of  all  other  elements  relative  to  the  mass  of  the 
hydrogen  atom.  Dalton  defined  the  atomic  mass  of  an  element  as 
the  mass  of  an  atom  of  that  element  compared  to  the  mass  of  a 
hydrogen  atom.  For  example,  the  masses  of  chlorine  and  hydrogen 
gas  that  react  to  form  hydrogen  chloride  (the  only  hydrogen  and 
chlorine  compound)  are  in  the  ratio  of  about  35V2  to  1 ;  therefore 
the  chlorine  atom  would  be  supposed  to  have  an  atomic  mass  of 
35 V2  atomic  mass  units.  This  definition  could  be  used  by  chemists 
in  the  nineteenth  century  even  before  the  actual  values  of  the 
masses  of  individual  atoms  (say  in  kilograms)  could  be  measured 
directly. 

During  the  nineteenth  century  chemists  extended  and  improved 
Dalton's  ideas.  They  studied  many  chemical  reactions  quantita- 
tively, and  developed  highly  accurate  methods  for  determining 
relative  atomic  and  molecular  masses.  Because  oxygen  combined 
readily  with  many  other  elements  chemists  decided  to  use  oxygen 
rather  than  hydrogen  as  the  standard  for  atomic  masses.  Oxygen 
was  assigned  an  atomic  mass  of  16  so  that  hydrogen  would  have 
an  atomic  mass  close  to  one.  The  atomic  masses  of  other  elements 
could  be  obtained  by  applying  the  laws  of  chemical  combination  to 
the  compounds  of  the  elements  with  oxygen.  Throughout  the  nine- 
teenth century  more  and  more  elements  were  identified  and  their 
atomic  masses  determined.  For  example,  the  table  on  the  next  page 
lists  63  elements  found  by  1872,  together  with  the  modern  values 
for  the  atomic  masses.  This  table  contains  much  valuable  informa- 
tion, which  we  shall  consider  at  greater  length  in  Sec.  17.4.  (The 
special  marks  on  the  table— circles  and  rectangles— will  be  useful 
then.) 

Q3    Was  the  simplest  chemical  formula  for  the  composition  of 
a  molecule  necessarily  the  correct  one? 

Q4    Why  did  Dalton  choose  hydrogen  as  the  unit  of  atomic  mass? 


SG  17.7 
SG  17.8 


The  system  of  atomic  masses  used 
in  modern  physical  science  is  based 
on  this  principle,  although  it  differs 
in  details  (and  the  standard  for 
comparison  by  international  agree- 
ment is  now  carbon  instead  of 
hydrogen  or  oxygen.) 


The  progress  made  in  identifying 
elements  in  the  19th  century  may 
be  seen  in  the  following  table. 

Total  number  of 


Year 

elements  identified 

1720 

14 

1740 

15 

1760 

17 

1780 

21 

1800 

31 

1820 

49 

1840 

56 

1860 

60 

1880 

69 

1900 

83 

Some  of  the  current  representations 
of  a  water  molecule. 


16 


The  Chemical  Basis  of  the  Atomic  Theory 


Elements  known  by  1872,  in  order  of 
increasing  relative  atomic  mass. 


Elements  known  by  1872 

Atomic 

Atomic 

Name 
hydrogen 

Symbol 
H 

Mass* 
1.0 

Name 

Symbol 

Mass* 

cadmium 

Cd 

112.4 

D  lithium 

Li 

6.9 

indium 

In 

114.8(113) 

beryllium 

Be 

9.0 

tin 

Sn 

118.7 

boron 

B 

10.8 

antimony 

Sb 

121.7 

carbon 

C 

12.0 

tellurium 

Te 

127.6(125) 

nitrogen 

N 

14.0 

O  iodine 

1 

126.9 

oxygen 

0 

16.0 

D  cesium 

Cs 

132.9 

O  fluorine 

F 

19.0 

barium 

Ba 

137.3 

D  sodium 

Na 

23.0 

didymium(**) 

Di 

(138) 

magnesium 

Mg 

24.3 

cerium 

Ce 

140.1 

aluminum 

Al 

27.0 

erbium 

Er 

167.3(178) 

silicon 

Si 

28.1 

lanthanum 

La 

138.9(180) 

phosphorus 

P 

31.0 

tantalum 

Ta 

180.9(182) 

sulfur 

S 

32.1 

tungsten 

W 

183.9 

O  chlorine 

CI 

35.5 

osmium 

Os 

190.2(195) 

D  potassium 

K 

39.1 

iridium 

Ir 

192.2(197) 

calcium 

Ca 

40.1 

platinum 

Pt 

195.1(198) 

titanium 

Ti 

47.9 

gold 

Au 

197.0(199) 

vanadium 

V 

50.9 

mercury 

Hg 

200.6 

chromium 

Cr 

52.0 

thallium 

TI 

204.4 

manganese 

Mn 

54.9 

lead 

Pb 

207.2 

iron 

Fe 

55.8 

bismuth 

Bi 

209.0 

cobalt 

Co 

58.9 

thorium 

Th 

232.0 

nickel 

Ni 

58.7 

uranium 

U 

238.0(240) 

copper 

Cu 

63.5 

zinc 

Zn 

65.4 

arsenic 
selenium 

As 

Se 

74.9 

79  0 

D  alkaline  metals 

O  bromine 
n  rubidium 

Br 
Rb 

79.9 
85.5 

O  halogens 

strontium 

Sr 

87.6 

•Atomic  masses  g 

ven  are  modern  values.  Where    1 

yttrium 

Yt 

88.9 

these   differ  greatly   from    those 

accepted  In 

zirconium 

Zr 

91.2 

1872.  the  old  val 

ues  are  given  In 

parentheses. 

niobium 
molybdenum 
ruthenium 
rhodium 

Nb 
Mo 
Ru 
Rh 

92.9 

95.9 
101.1(104) 
102.9(104) 

•*Didymium  (Di)  was  later  shown  to  be  a  mixture 
of  two  different  elements,  namely  praseodym- 
ium (Pr;  atomic  mass  140.9)  and  neodymium 
(Nd:  atomic  mass  144.2). 

palladium 

Pd 

106.4 

silver 

Ag 

107.9 

The  standard  international  chemical 
symbols  are  derived  from  languages 
other  than  English.  The  Latin  name 
for  sodium  is  natrium,  hence  the 
symbol  is  Na. 


17.3     Other  properties  of  the  elements:  combining  capacity 

As  a  result  of  studies  of  chemical  compounds,  chemists  were 
able  to  design  chemical  formulas  that  indicate  by  a  kind  of  symbolic 
shorthand  the  number  of  atoms  in  each  molecule  of  a  compound. 
For  example,  water  has  the  familiar  formula  H2O,  which  indicates 
that  a  molecule  of  water  contains  two  atoms  of  hydrogen  (H)  and 
one  atom  of  oxygen  (O).  (Dalton  thought  it  was  HO.)  Hydrogen 
chloride  (hydrochloric  acid  when  dissolved  in  water)  had  the  formula 
HCl,  signifying  that  one  atom  of  hydrogen  combines  with  one  atom 
of  chlorine  (CI).  Common  salt  may  be  represented  by  the  formula 
NaCl;  this  indicates  that  one  atom  of  sodium  (Na)  combines  with 
one  atom  of  chlorine  to  form  one  molecule  of  sodium  chloride — 
common  table  salt.  Another  salt,  calcium  chloride  (often  used  to 
melt  ice  on  roads),  has  the  formula  CaCl,;  one  atom  of  calcium 
(Ca)  combined  with  two  atoms  of  chlorine  to  form  this  compound. 
Carbon  tetrachloride,  a  common  compound  of  chlorine  used  for  dry 


Section  17.3 


17 


cleaning,  has  the  formula  CCI4  where  C  stands  for  a  carbon  atom 
that  combines  with  four  chlorine  atoms.  Another  common  sub- 
stance, ammonia,  has  the  formula  NH3;  in  this  case  one  atom  of 
nitrogen  (N)  combines  with  three  atoms  of  hydrogen. 

There  are  especially  significant  examples  of  combining  capacity 
among  the  gaseous  elements.  For  example,  the  gas  hydrogen  occurs 
in  nature  in  the  form  of  molecules,  each  of  which  contains  two 
hydrogen  atoms.  The  molecule  of  hydrogen  consists  of  two  atoms 
and  has  the  formula  Hg.  Similarly,  chlorine  has  the  molecular 
formula  CI2.  Chemical  analysis  always  gives  these  results.  It  would 
be  inconsistent  with  experiment  to  assign  the  formula  H3  or  H4  to  a 
molecule  of  hydrogen,  or  CI,  CI3,  or  CI4  to  a  molecule  of  chlorine. 
Moreover,  each  element  shows  great  consistency  in  its  combining 
proportions  with  other  elements.  For  example,  calcium  and  oxygen 
seem  to  have  twice  the  combining  capacity  of  hydrogen  and 
chlorine — one  atom  of  hydrogen  is  enough  for  one  atom  of  chlorine, 
but  two  hydrogens  are  needed  to  combine  with  oxygen  and  two 
chlorines  are  required  to  combine  with  calcium. 

The  above  examples  indicate  that  different  elements  have 
different  capacities  for  chemical  combination.  It  appeared  that 
each  species  of  atom  is  characterized  by  some  definite  combining 
capacity  (which  is  sometimes  called  valence).  At  one  time  combin- 
ing capacity  was  considered  as  though  it  might  represent  the 
number  of  "hooks"  possessed  by  a  given  atom,  and  thus  the  number 
of  links  that  an  atom  could  form  with  others  of  the  same  or  different 
species.  If  hydrogen  and  chlorine  atoms  each  had  just  one  hook 
(that  is,  a  combining  capacity  of  1)  we  would  readily  understand 
how  it  is  that  molecules  like  H2,  CI2,  and  HCl  are  stable,  while 
certain  other  species  like  H3,  H2CI,  HCI2,  and  CI3  don't  exist  at  all. 
And  if  the  hydrogen  atom  is  thus  assigned  a  combining  capacity 
of  1,  the  formula  of  water  (H2O)  requires  that  the  oxygen  atom  has 
two  hooks  or  a  combining  capacity  of  2.  The  formula  NH3  for 
ammonia  leads  us  to  assign  a  combining  capacity  of  three  to  nitro- 
gen; the  formula  CH4  for  methane  leads  us  to  assign  a  capacity  of 
4  to  carbon;  and  so  on.  Proceeding  in  this  fashion,  we  can  assign 
a  combining  capacity  number  to  each  of  the  known  elements. 
Sometimes  complications  arise  as,  for  example,  in  the  case  of 
sulfur.  In  H2S  the  sulfur  atom  seems  to  have  a  combining  capacity 
of  2,  but  in  such  a  compound  as  sulfur  trioxide  (SO3),  sulfur  seems 
to  have  a  combining  capacity  of  6.  In  this  case  and  others,  then, 
we  may  have  to  assign  two  (or  even  more)  different  possible  capaci- 
ties to  an  element.  At  the  other  extreme  of  possibilities  are  those 
elements  like  helium  and  neon  which  have  not  been  found  as  parts 
of  compounds — and  to  these  elements  we  may  appropriately  assign 
a  combining  capacity  of  zero. 

The  atomic  mass  and  combining  capacities  are  numbers  that 
can  be  assigned  to  an  element;  they  are  "numerical  characteriza- 
tions" of  the  atoms  of  the  element.  There  are  other  numbers  which 
represent  properties  of  the  atoms  of  the  elements,  but  atomic  mass 
and  combining  capacity  were  the  two  most  important  to  nineteenth- 


In  the  thirteenth  century  the 
theologian  and  philosopher  Albert 
Magnus  (Albert  the  Great)  intro- 
duced the  idea  of  affinity  to  denote 
an  attractive  force  between  sub- 
stances that  causes  them  to  enter 
into  chemical  combination.  It  was 
not  until  600  years  later  that  it 
became  possible  to  replace  this 
qualitative  notion  by  quantitative 
concepts.  Combining  capacity  is  one 
of  these  concepts. 


Representations  of  molecules  formed 
from  "atoms  with  hooks.  "  Of  course 
this  conception  is  just  a  guide  to  the 
imagination.  There  are  no  such  me- 
chanical linkages  among  atoms. 


SG  17.9 


Since  oxygen  combines  with  a 
greater  variety  of  elements, 
combining  capacity  of  an  element 
was  commonly  determined  by  its 
combination  with  oxygen.  For 
example,  an  element  X  that  is  found 
to  have  an  "oxide  formula"  XO 
would  have  a  combining  capacity 
equal  to  oxygen's:  2. 


18 


The  Chemical  Basis  of  the  Atomic  Theory 


century  chemists.  These  numbers  were  used  in  attempts  to  find 
order  and  regularity  among  the  elements— a  problem  which  will  be 
discussed  in  the  next  section. 

Q5    At  this  point  we  have  two  numbers  which  are  character- 
istic of  the  atoms  of  an  element.  What  are  they? 

Q6  Assume  the  combining  capacity  of  oxygen  is  2.  In  each  of 
the  following  molecules,  give  the  combining  capacity  of  the  atoms 
other  than  oxygen:  CO,  CO2,  N2O5,  Na^O  and  MnO. 


There  were  also  many  false  trails. 
Thus  in  1829  the  German  chemist 
Johann  Wolfgang  Dbbereiner 
noticed  that  elements  often  formed 
groups  of  three  members  with 
similar  chemical  properties.  He 
identified  the  "triads":  chlorine, 
bromine  and  iodine:  calcium, 
strontium  and  barium:  sulfur, 
selenium  and  tellurium:  iron,  cobalt 
and  manganese.  In  each  "triad,"  the 
atomic  mass  of  the  middle  member 
was  approximately  the  arithmetical 
average  of  the  masses  of  the  other 
two  elements.  But  all  this  turned 
out  to  be  of  little  significance. 


17.4    The  search  for  order  and  regularity  among  the  elements 

By  1872  sixty-three  elements  were  known;  they  are  listed  in 
the  table  on  p.  16  with  their  atomic  masses  and  chemical  symbols. 
Sixty-three  elements  are  many  more  than  Aristotle's  four:  and 
chemists  tried  to  make  things  simpler  by  looking  for  ways  of 
organizing  what  they  had  learned  about  the  elements.  They  tried  to 
find  relationships  among  the  elements — a  quest  somewhat  like 
Kepler's  earlier  search  for  rules  that  would  relate  the  motions  of 
the  planets  of  the  solar  system. 

In  addition  to  relative  atomic  masses,  many  other  properties  of 
the  elements  and  their  compounds  were  determined.  Among  these 
properties  were:  melting  point,  boiling  point,  density,  electrical 
conductivity,  heat  conductivity,  heat  capacity  (the  amount  of  heat 
needed  to  change  the  temperature  of  a  sample  of  a  substance  by  1 
C)  hardness,  and  refractive  index.  The  result  was  that  by  1870  an 
enormous  amount  of  information  was  available  about  a  large 
number  of  elements  and  their  compounds. 

It  was  the  English  chemist  J.  A.  R.  Newlands  who  pointed  out 
in  1865  that  the  elements  could  usefully  be  listed  simply  in  the 
order  of  increasing  atomic  mass.  When  this  was  done,  a  curious  fact 
became  evident;  similar  chemical  and  physical  properties  appeared 
over  and  over  again  in  the  list.  Newlands  believed  that  there  was 
in  the  whole  list  a  periodic  recurrence  of  elements  with  similar 
properties:  ".  .  .  the  eighth  element,  starting  from  a  given  one,  is  a 
kind  of  repetition  of  the  first,  like  the  eighth  note  in  an  octave  of 
music."  Newlands'  proposal  was  met  with  skepticism.  One  chemist 
even  suggested  that  Newlands  might  look  for  a  similar  pattern  in 
an  alphabetical  list  of  elements. 

Yet,  existent  relationships  did  indeed  appear.  There  seemed  to 
be  families  of  elements  with  similar  properties.  One  such  family 
consists  of  the  so-called  alkali  metals— hihium.  sodium,  potassium, 
rubidium  and  cesium.  We  have  identified  these  elements  by  a  D  in 
the  table  on  p.  16.  All  these  metals  are  similar  physically.  They  are 
soft  and  have  low  melting  points.  The  densities  of  these  metals  are 
very  low;  in  fact,  lithium,  sodium  and  potassium  are  less  dense 
than  water.  The  alkali  metals  are  also  similar  chemically.  They  all 
have  combining  capacity  1.  They  all  combine  with  the  same  other 
elements  to  form  similar  compounds.  They  form  compounds  readily 
with  other  elements,  and  so  are  said  to  be  highly  "reactive";  conse- 


Section  17.5 


19 


quently,  they  do  not  occur  free  in  nature,  but  are  always  found  in 
combination  with  other  elements. 

Another  family  of  elements,  called  the  halogens,  includes 
fluorine,  chlorine,  bromine  and  iodine.  The  halogens  may  be  found 
in  the  table  on  p.  16  identified  by  small  circles. 

Although  these  four  halogen  elements  exhibit  some  marked 
dissimilarities  (for  example,  at  25  °C  the  first  two  are  gases,  the 
third  a  liquid,  the  last  a  volatile  solid),  they  also  have  much  in  com- 
mon. They  all  combine  violently  with  many  metals  to  form  white, 
crystalline  salts  (halogen  means  "salt-former");  those  salts  have 
similar  formulas,  such  as  NaF,  NaCl,  NaBr  and  Nal,  or  MgFz, 
MgCla,  MgBra  and  Mgla.  From  much  similar  evidence  chemists 
noticed  that  all  four  members  of  the  family  seem  to  have  the  same 
valence  with  respect  to  any  other  particular  element.  All  four  ele- 
ments from  simple  compounds  with  hydrogen  (HF,  HCI,  HBr,  HI) 
which  dissolve  in  water  and  form  acids.  All  four,  under  ordinary 
conditions,  exist  as  diatomic  molecules;  that  is,  each  molecule 
contains  two  atoms.  But  notice:  each  halogen  precedes  an  alkali 
metal  in  the  list,  although  the  listing  was  ordered  simply  by 
increasing  atomic  mass.  It  is  as  if  some  new  pattern  is  coming  out 
of  a  jig-saw  puzzle. 

The  elements  which  follow  the  alkali  metals  in  the  list  also 
form  a  family,  the  one  called  the  alkaline  earth  family;  this  family 
includes  beryllium,  magnesium,  calcium,  strontium  and  barium. 
Their  melting  points  and  densities  are  higher  than  those  of  the 
alkali  metals.  The  alkaline  earths  all  have  a  valence  of  two.  They 
react  easily  with  many  elements,  but  not  as  easily  as  do  the  alkali 
metals. 

Recognition  of  the  existence  of  these  f  amihes  of  elements 
encouraged  chemists  to  look  for  a  systematic  way  of  arranging  the 
elements  so  that  the  members  of  a  family  would  group  together. 
Many  schemes  were  suggested;  the  most  successful  and  far  reach- 
ing was  that  of  the  Russian  chemist  D.  I.  Mendeleev. 

Q7    What  are  those  properties  of  elements  which  recur  system- 
atically with  increasing  atomic  mass? 

17.5    Mendeleev's  periodic  table  of  the  elements 

Mendeleev,  examining  the  properties  of  the  elements,  reached 
the  conclusion  that  the  atomic  mass  was  the  fundamental  "numeri- 
cal characterization"  of  each  element.  He  discovered  that  if  the 
elements  were  arranged  in  a  table  in  the  order  of  their  atomic 
masses— but  in  a  special  way,  a  bit  like  cards  laid  out  in  the  game 
of  solitaire— the  different  chemical  families  turned  out  to  fall  into 
the  different  vertical  columns  of  the  table.  There  was  no  evident 
physical  reason  why  this  should  be  so,  but  it  was  a  hint  toward 
some  remarkable  connection  among  all  elements. 


Modern  chemists  use  the  word 
'valence"  less  and  less  in  the  sense 
we  use  it  here.  They  are  more  likely 
to  discuss  "combining  number"  or 
"oxidation  number."  Even  the 
idea  of  a  definite  valence  number 
for  an  element  has  changed,  since 
combining  properties  can  be  dif- 
ferent under  different  conditions. 


Li    7 

Be  9.4 

B  11 

C  12 

N  14 

O  16 

F  19 

Ma  23 

Mg  24 

Al  27.4 

Si  28 

P  31 

S   32 

CI  35.3 

K39 

Ca  40 

Ti50 

V51 

...etc. 

Although  the  properties  of  elements 
do  recur  periodically  with  increasing 
atomic  weight,  Newlands  had  not 
realized  that  the  separation  of 
similar  elements  in  the  list  becomes 
greater  for  the  heavier  elements. 


In  this  table,  hydrogen  was  omitted 
because  of  its  unique  properties. 
Helium  and  the  other  elements  of 
the  family  of  "noble  gases"  had  not 
yet  been  discovered. 


20 


The  Chemical  Basis  of  the  Atomic  Theory 


Dmitri  Ivanovich  Mendeleev  (men- 
deh-lay>'-ef)  (1834-1907)  received  his 
first  science  lessons  from  a  political 
prisoner  who  had  been  previously 
banished  to  Siberia  by  the  Czar.  Un- 
able to  get  into  college  in  Moscow,  he 
was  accepted  in  St.  Petersburg,  where 
a  friend  of  his  father  had  some  in- 
fluence. In  1866  he  became  a  profes- 
sor of  chemistry  there:  in  1869  he  pub- 
lished his  first  table  of  the  sixty-three 
then  known  elements  arranged  ac- 
cording to  increasing  atomic  mass. 
His  paper  was  translated  into  German 
at  once  and  so  became  known  to  sci- 
entists everywhere.  Mendeleev  came 
to  the  United  States,  where  he  studied 
the  oil  fields  of  Pennsylvania  in  order 
to  advise  his  country  on  the  develop- 
ment of  the  Caucasian  resources.  His 
liberal  political  views  caused  him 
often  to  be  in  trouble  with  the  oppres- 
sive regime  of  the  Czars. 


As  in  the  table  on  the  preceding  page,  Mendeleev  set  down 
seven  elements,  from  lithium  to  fluorine,  in  order  of  increasing 
atomic  masses,  and  then  put  the  next  seven,  from  sodium  to 
chlorine,  in  the  second  row.  The  periodicity  of  chemical  behavior  is 
already  evident  before  we  go  on  to  write  the  third  row.  In  the  first 
column  on  the  left  are  the  first  two  alkali  metals.  In  the  seventh 
column  are  the  first  two  members  of  the  family  of  halogens.  Indeed, 
within  each  of  the  columns  the  elements  are  chemically  similar, 
having,  for  example,  the  same  characteristic  combining  capacity. 

When  Mendeleev  added  a  third  row  of  elements,  potassium  (K) 
came  below  elements  Li  and  Na,  which  are  members  of  the  same 
family  and  have  the  same  oxide  formula,  X2O,  and  the  same 
combining  capacity  1.  Next  in  the  row  is  Ca,  oxide  formula  XO  as 
with  Mg  and  Be  above  it.  In  the  next  space  to  the  right,  the  element 
of  next  higher  atomic  mass  should  appear.  Of  the  elements  known 
at  the  time,  the  next  heavier  was  titanium  (Ti),  and  it  was  placed  in 
this  space,  right  below  aluminum  (Al)  and  boron  (B)  by  various 
workers  who  had  tried  to  develop  such  schemes.  Mendeleev,  how- 
ever, recognized  that  titanium  (Ti)  has  chemical  properties  similar 
to  those  of  carbon  (C)  and  silicon  (Si).  For  example,  a  pigment, 
titanium  white,  Ti02,  has  a  formula  comparable  to  CO2  and  Si02. 
Therefore  he  concluded  that  titanium  should  be  put  in  the  fourth 
column.  Then,  if  all  this  is  not  just  a  game  but  has  deeper  meaning. 
Mendeleev  thought,  there  should  exist  a  hitherto  unsuspected  ele- 
ment with  atomic  mass  between  that  of  calcium  (40)  and  titanium 
(50),  and  with  an  oxide  X2O3.  Here  was  a  definite  prediction. 
Mendeleev  found  also  other  cases  of  this  sort  among  the  remaining 
elements  when  they  were  added  to  this  table  of  elements  with  due 
regard  to  the  family  properties  of  elements  in  each  column. 

The  table  below  is  Mendeleev's  periodic  system,  or  "periodic 
table"  of  the  elements,  as  proposed  in  1872.  He  distributed  the  63 
elements  then  known  (with  5  in  doubt)  in  12  horizontal  rows  or 
series,  starting  with  hydrogen  in  a  unique  separated  position  at  the 
top  left,  and  ending  with  uranium  at  the  bottom  right.  All  elements 


Periodic    classification    of    the    ele- 
ments; Mendeleev,  1872. 


GROUP—* 

I 

II 

Ill 

IV 

V          i          VI 

VII                       VIII 

Higher  oxides 
and  hydrides 

R2O 

RO 

RjOa 

RO2 
H4R 

R205 
H3R 

RO3 
H2R 

R207 

HR 

RO, 

Ul 

1 

H(l) 

2 

Li(7) 

Be(9.4) 

B(ll) 

C(12) 

N(14) 

0(16) 

F(19) 

3 

Na(23) 

Mg(24) 

Al(27.3) 

Si(28) 

P(3I) 

S(32) 

Cl(35.5) 

4 

K(39) 

Ca(40) 

-(44) 

Ti(48) 

V(SI) 

Cr(52) 

Mn(55) 

Fe(S6),  Co(59). 
Ni(59).  Cu(63) 

5 

iCu(63)l 

Zn(6S) 

-(68) 

-(72) 

A«(75) 

Se(78) 

Br{80) 

6 

Rb(85) 

Sr(87) 

?Yt(88) 

Zr(90) 

Nb(94) 

Mo(96) 

-(100) 

Ru(104),Rh(104), 
Pd(106)..\«(108) 

7 

lAg(108)! 

Cd(112) 

ln(ll3) 

Sn(118) 

Sb(122) 

Te(l25) 

1(127) 

8 

C8(133) 

Ba(I37) 

?Di(138) 

?Ce(140) 







9 















10 





?Er(I78) 

?U(I80) 

Ta(l82) 

W(184) 



08(195),  Ir(197), 
Pt(196).  Au(199) 

11 

|Au(199)l 

Hg(200) 

Tl(204) 

Pb(207) 

Bi(208) 





12 





Th(231) 



U(240) 

Section  17.5 


21 


were  listed  in  order  of  increasing  atomic  mass  (Mendeleev's  values 
given  in  parentheses),  but  were  so  placed  that  elements  with  similar 
chemical  properties  are  in  the  same  vertical  column  or  group. 
Thus  in  Group  VII  are  all  the  halogens;  in  Group  VIII,  only  metals 
that  can  easily  be  drawn  into  wires;  in  Groups  I  and  II,  metals  of 
low  densities  and  melting  points;  and  in  I,  the  family  of  alkali 
metals. 

The  table  at  the  bottom  of  the  previous  page  shows  many  gaps. 
Also,  not  all  horizontal  rows  (series)  have  equally  many  elements. 
Nonetheless,  the  table  revealed  an  important  generalization; 
according  to  Mendeleev, 

For  a  true  comprehension  of  the  matter  it  is  very  impor- 
tant to  see  that  all  aspects  of  the  distribution  of  the 
elements  according  to  the  order  of  their  atomic  weights 
express  essentially  one  and  the  same  fundamental  depen- 
dence— periodic  properties. 

There  is  gradual  change  in  physical  and  chemical  properties  within 
each  vertical  group,  but  there  is  a  more  striking  periodic  change  of 
properties  in  the  horizontal  sequence. 

This  periodic  law  is  the  heart  of  the  matter  and  a  real  novelty. 
Perhaps  we  can  best  illustrate  it  as  Lothar  Meyer  did,  by  drawing 
a  graph  that  shows  the  value  of  some  measureable  physical  quantity 
as  a  function  of  atomic  mass.  Below  is  a  plot  of  the  relative 
atomic  volumes  of  the  elements,  the  space  taken  up  by  an  atom  in 
the  liquid  or  solid  state.  Each  circled  point  on  this  graph  represents 
an  element;  a  few  of  the  points  have  been  labeled  with  the 
identifying  chemical  symbols.  Viewed  as  a  whole,  the  graph 
demonstrates  a  striking  periodicity:  as  the  mass  increases  starting 
with  Li,  the  atomic  volume  first  drops,  then  increases  to  a  sharp 
maximum,  drops  off  again  and  increases  to  another  sharp  maximum, 
and  so  on.  And  at  the  successive  peaks  we  find  Li,  Na,  K,  Rb,  and 
Cs,  the  members  of  the  family  of  alkali  metals.  On  the  left-hand 
side  of  each  peak,  there  is  one  of  the  halogens. 


70 


I  50 


30! 


10 


50  70  90 

Atomic  mass    (amu) 


110 


130 


The   'atomic  volume"  is  defined 
as  the  atomic  mass  divided  by  the 
density  of  the  element  in  its  liquid 
or  solid  state. 


In  1864,  the  German  chemist  Lothar 
Meyer  wrote  a  chemistry  textbook. 
In  this  book,  he  considered  how  the 
properties  of  the  chemical  elements 
might  depend  on  their  atomic 
masses.  He  later  found  that  if  he 
plotted  atomic  volume  against  the 
atomic  mass,  the  line  drawn  through 
the  plotted  points  rose  and  fell  in 
two  long  periods.  This  was  exactly 
what  Mendeleev  had  discovered  in 
connection  with  valence.  Mendeleev 
published  his  first  result  in  1869; 
Meyer,  as  he  himself  later  admitted, 
lacked  the  courage  to  include  provi- 
sion for  empty  spaces  that  would 
amount  to  the  prediction  of  the 
discovery  of  unknown  elements. 
Nevertheless,  Meyer  should  be 
given  credit  for  the  idea  of  the 
periodic  table. 


The    atomic    volumes    of    elements 
graphed  against  their  atomic  masses. 


22 


The  Chemical  Basis  of  the  Atomic  Theory 


Mendeleev's  periodic  table  of  the  elements  not  only  provided  a 
remarkable  correlation  of  the  elements  and  their  properties,  it  also 
enabled  him  to  predict  that  certain  unknown  elements  should  exist 
and  what  many  of  their  properties  should  be.  To  estimate  physical 
properties  of  a  missing  element,  Mendeleev  averaged  the  properties 
of  its  nearest  neighbors  in  the  table:  those  to  right  and  left,  above 
and  below.  A  striking  example  of  Mendeleev's  success  in  using  the 
table  in  this  way  is  his  set  of  predictions  concerning  the  gap  in 
Series  5,  Group  IV.  Group  IV  contains  silicon  and  elements  re- 
sembling it.  Mendeleev  assigned  the  name  "eka-silicon"  (Es)  to  the 
unknown  element.  His  predictions  of  the  properties  of  this  element 
are  listed  in  the  left-hand  column  below.  In  1887,  this  element 
was  isolated  and  identified  (it  is  now  called  "germanium",  Ge);  its 
properties  are  listed  in  the  right-hand  column.  Notice  how  remark- 
ably close  Mendeleev's  predictions  are  to  the  properties  actually 
found. 


"The  following  are  the 
properties  which  this 
element  should  have  on 
the  basis  of  the  known 
properties  of  silicon, 
tin,  zinc,  and  arsenic. 

Its  atomic  mass  is 
nearly  72,  its  forms  a 
higher  oxide  EsOa,  .  .  .  Es 
gives  volatile  organo- 
metallic  compounds;  for 
instance  .  .  .  Es  (€2^2)4, 
which  boils  at  about  160°, 
etc.;  also  a  volatile  and 
liquid  chloride,  EsCl^, 
boiling  at  about  90°  and 
of  specific  gravity  about 
1.9.  ..  .  the  specific  gravity 
of  Es  will  be  about  5.5, 
and  ESO2  will  have  a  spe- 
cific gravity  of  about  4.7, 
etc " 


The  predictions  in  the  left 
column  were  made  by 
Mendeleev  in  1871.  In 
1887  an  element  (german- 
ium) was  discovered  which 
was  found  to  have  the 
following  properties: 

Its  atomic  mass  is  72.5. 
It  forms  an  oxide  GeOa, 
and  an  organo- 
metallic  compound 
Ge(C2H5)4  which  boils  at 
160°  and  forms  a  liquid 
chloride  GeCl4  which 
boils  at  83°  C  and  has  a 
specific  gravity  of  1.9. 
The  specific  gravity  of 
germanium  is  5.5  and  the 
specific  gravity  of 
GeOi  is  4.7. 


The  daring  of  Mendeleev  is  shown  in  his  willingness  to  venture 
detailed  numerical  predictions;  the  sweep  and  power  of  his  system 
is  shown  above  in  the  remarkable  accuracy  of  those  predictions.  In 
similar  fashion,  Mendeleev  described  the  properties  to  be  expected 
for  the  then  unknown  elements  that  he  predicted  to  exist  in  gaps  in 
Group  III,  period  4,  and  in  Group  III.  period  5— elements  now  called 
gallium  and  scandium — and  again  his  predictions  turned  out  to  be 
remarkably  accurate. 

Although  not  every  aspect  of  Mendeleev's  work  yielded  such 
successes,  these  were  indeed  impressive  results,  somewhat 


Section  17.6 


23 


reminiscent  of  the  successful  use  of  Newtonian  laws  to  find  an 
unknown  planet.  Successful  numerical  predictions  like  these  are 
among  the  most  desired  results  in  physical  science— even  if  in 
Mendeleev's  case  it  was  still  mysterious  why  the  table  should  work 
the  way  it  did. 

Q8    Why  is  Mendeleev's  table  called  "periodic  table"? 
Q9    What  was  the  basic  ordering  principle  in  Mendeleev's  table? 
Q10  What  reasons  led  Mendeleev  to  leave  gaps  in  the  table? 
Q11  What  success  did  Mendeleev  have  in  the  use  of  the  table? 


The  discovery  of  Uranus  and  Nep- 
tune is  described  in  Text  Chapter  8. 


17.6    The  modern  periodic  table 

The  periodic  table  has  had  an  important  place  in  chemistry  and 
physics  for  a  century.  It  presented  a  serious  challenge  to  any  theory 
of  the  atom  proposed  after  1880:  the  challenge  that  the  theory 
provide  an  explanation  for  the  wonderful  order  among  the  elements 
as  expressed  by  the  table.  A  successful  model  of  the  atom  must 
provide  a  physical  reason  why  the  table  works  as  it  does.  In  Chapter 
19  we  shall  see  how  one  model  of  the  atom— the  Bohr  model— met 
this  challenge. 

Since  1872  many  changes  have  had  to  be  made  in  the  periodic 
table,  but  they  have  been  changes  in  detail  rather  than  in  general 
ideas.  None  of  these  changes  has  affected  the  basic  periodic  feature 
among  the  properties  of  the  elements.  A  modern  form  of  the  table 
with  current  values  is  shown  in  the  table  below. 


A  modern  form  of  the  periodic  table 
of  the  chemical  elements.  The  number 
above  the  symbol  is  the  atomic  mass, 
the  number  below  the  symbol  is  the 
atomic  number. 


Group—* 
Period 

i 

I 

II 

III 

IV 

V 

VI 

VII 

0 

1.0080 

4.0026 

1 

H 

He 

1 

2 

6.939 

9.012 

10.811 

12.011 

14.007 

15.999 

18.998 

20.183 

2 

U 
3 

Be 

4 

B 
5 

0 
6 

N 

7 

0 

8 

F 

9 

Ne 
10 

22.990 

24.31 

26.98 

28.09 

30.97 

32.06 

35.45 

39.95 

3 

Na 

n 

Mg 

12 

Al 
13 

Si 
14 

P 
15 

S 
16 

Cl 

17 

Ar 

18 

39.10 

40.08 

44.96 

47.90 

50.94 

52.00 

54.94 

55.85 

58.93 

58.71 

63.54 

65.37 

69.72 

72.59 

74.92 

78.96 

79.91 

83.80 

4 

K 

Ca 

Sc 

Ti 

V 

Or 

M 

Fe 

Co 

Ni 

Cu 

Zd 

Ga 

Ge 

As 

Se 

Br 

Kr 

19 

20 

21 

22 

23 

24 

25 

26 

27 

28 

29 

30 

31 

32 

33 

34 

35 

36 

85.47 

87.62 

88.91 

91.22 

92.91 

95.94 

(99) 

101.07 

102.91 

106.4 

107.87 

112.40 

114.82 

118.69 

121.75 

127.60 

126.9 

131.30 

5 

Rb 

Sr 

Y 

Zr 

Nb 

Mo 

Tc 

Ru 

Rh 

Pd 

Ag 

Cd 

In 

Sd 

Sb 

Te 

I 

Xe 

37 

38 

39 

40 

41 

42 

43 

44 

45 

46 

47 

48 

49 

50 

51 

52 

53 

54 

132.91 

137.34 

178.49 

180.95 

183.85 

186.2 

190.2 

1922 

195.09 

196.97 

200.59 

204.37 

207.19 

208.98 

210 

(210) 

222 

6 

Cs 

Ba 

• 

Hf 

Ta 

W 

Re 

0» 

Ir 

Pt 

Au 

Hg 

Tl 

Pb 

Bi 

Po 

At 

Rn 

55 

56 

57-71 

72 

73 

74 

75 

76 

77 

78 

79 

80 

81 

82 

83 

84 

85 

86 

(223) 

226.05 

\ 

7 

Fr 
87 

Ra 

88 

+ 
89-103 

)   J 

•Rarc- 

138.91 

140.12 

140.91 

144.27 

(147) 

150.35 

151.96 

157.25 

158.92 

162.50 

164.93 

167.26 

168.93 

173.04 

174.97 

parth 

U 

Ce 

Pr 

Nd 

Pm 

Sm 

Eu 

Gd 

Tb 

Dy 

Ho 

Er 

Tm 

Yb 

Lu 

metals 

57 

58 

59 

60 

61 

62 

63 

04 

65 

66 

67 

68 

69 

70 

71 

t 

227 

232.04 

231 

238.03 

(237) 

(242) 

(243) 

(245) 

(249) 

(249) 

(253) 

(255) 

(256) 

(253) 

(257) 

\ctmide 

Ac 

Th 

Pa 

U 

Np 

Pu 

Am 

Cm 

Bk 

Of 

E 

Fm 

Mv 

No 

Lw 

metals 

89 

90 

91 

92 

93 

94 

95 

96 

97 

98 

99 

100 

101 

102 

103 

24 


The  Chemical  Basis  of  the  Atomic  Theory 


Although  Mendeleev's  table  had 
eight  columns,  the  column  labelled 
VIII  did  not  contain  a  family  of 
elements.  It  contained  the  "transi- 
tion" elements  which  are  now  placed 
in  the  long  series  (periods)  labelled 
4,  5  and  6  in  the  table  on  p.  23.  The 
group  labelled  "O"  in  that  table  does 
consist  of  a  family  of  elements, 
the  noble  gases,  which  do  have 
similar  properties  in  common. 


Helium  was  first  detected  in  the 
spectrum  of  the  sun  in  1868 
(Chapter  19).  Its  name  comes  from 
helios,  the  Greek  word  for  the  sun. 

In  chemistry,  elements  such  as  gold 
and  silver  that  react  only  rarely  with 
other  elements  were  called  "noble." 


One  difference  between  the  modern  and  older  tables  results  from 
new  elements  having  been  found.  Forty  new  elements  have  been 
identified  since  1872,  so  that  the  table  now  contains  103  or  more 
elements.  Some  of  these  new  elements  are  especially  interesting, 
and  you  will  learn  more  about  them  in  Unit  6. 

Comparison  of  the  modern  form  of  the  table  with  Mendeleev's 
table  shows  that  the  modern  table  contains  eight  groups,  or  famihes, 
instead  of  seven.  The  additional  group  is  labeled  "zero."  In  1894, 
the  British  scientists  Lord  Rayleigh  and  William  Ramsay  discovered 
that  about  1  percent  of  our  atmosphere  consists  of  a  gas  that  had 
previously  escaped  our  detection.  It  was  given  the  name  argon 
(symbol  Ar).  Argon  does  not  seem  to  enter  into  chemical  combina- 
tion with  any  other  elements,  and  is  not  similar  to  any  of  the  groups 
of  elements  in  Mendeleev's  original  table.  Later,  other  elements 
similar  to  argon  were  also  discovered:  helium  (He),  neon  (Ne), 
krypton  (Kr),  xenon  (Xe),  and  radon  (Rn).  These  elements  are 
considered  to  form  a  new  group  or  family  of  elements  called  the 
"noble  gases."  The  molecules  of  the  noble  gases  contain  only  one 
atom,  and  until  recent  years  no  compound  of  any  noble  gas  was 
known.  The  group  number  zero  was  thought  to  correspond  to  the 
chemical  inertness,  or  zero  combining  capacity  of  the  members  of 
the  group.  In  1963,  some  compounds  of  xenon  and  krypton  were 
produced,  so  we  now  know  that  these  elements  are  not  really  inert. 
These  compounds  are  not  found  in  nature,  however,  and  some  are 
very  reactive,  and  therefore  very  difficult  to  keep.  The  noble  gases 
as  a  group  are  certainly  less  able  to  react  chemically  than  any  other 
elements. 

In  addition  to  the  noble  gases,  two  other  sets  of  elements  had  to 
be  included  in  the  table.  After  the  fifty-seventh  element,  lanthanum, 
room  had  to  be  made  for  a  whole  set  of  14  elements  that  are  almost 
indistinguishable  chemically,  known  as  the  rare  earths  or  lantha- 
nide  series.  Most  of  these  elements  were  unknown  in  Mendeleev's 
time.  Similarly,  after  actinium  at  the  eighty-ninth  place,  there  is  a 
set  of  14  very  similar  elements,  forming  what  is  called  the  actinide 
series.  These  elements  are  shown  in  two  rows  below  the  main  table. 
No  more  additions  are  expected  except,  possibly,  at  the  end  of  the 
table.  There  are  no  known  gaps,  and  we  shall  see  in  Chapters  19 
and  20  that  according  to  the  best  theory  of  the  atom  now  available, 
no  new  gaps  are  expected  to  exist  within  the  table. 

Besides  the  addition  of  new  elements  to  the  periodic  table,  there 
have  also  been  some  changes  of  a  more  general  type.  As  we  have 
seen,  Mendeleev  arranged  most  of  the  elements  in  order  of 
increasing  atomic  mass.  In  the  late  nineteenth  century,  however, 
this  basic  scheme  was  found  to  break  down  in  a  few  places.  For 
example,  the  chemical  properties  of  argon  (Ar)  and  potassium  (K) 
demand  that  they  should  be  placed  in  the  eighteenth  and  nineteenth 
positions,  whereas  on  the  basis  of  their  atomic  masses  alone  (39.948 
for  argon,  39.102  for  potassium),  their  positions  should  be  reversed. 
Other  reversals  of  this  kind  are  also  necessary,  for  example,  for  the 
fifty-second  element,  tellurium  (atomic  mass  =  127.60)  and  the  fifty- 
third,  iodine  (atomic  mass  =  126.90). 


Section  17.7 


25 


The  numbers  that  place  elements  in  the  table  with  the  greatest 
consistency  in  periodic  properties  are  called  the  atomic  numbers 
of  the  elements.  The  atomic  numbers  of  all  the  elements  are  given 
in  the  table  on  p.  23.  The  atomic  number  is  usually  denoted  by  the 
symbol  Z;  thus  for  hydrogen,  Z  =  1,  for  chlorine,  Z  =  17,  for 
uranium,  Z  =  92.  In  Chapter  19  we  shall  see  that  the  atomic  number 
has  a  fundamental  physical  meaning  related  to  atomic  structure, 
and  that  is  the  key  to  both  the  many  puzzhng  successes  and  few 
puzzUng  failures  of  Mendeleev's  scheme.  Since  he  used  atomic 
mass  as  the  basis  for  the  order  of  the  elements,  he  preferred  to 
believe  that  the  apparent  reversals  were  due  to  error  in  the  values 
for  the  atomic  masses. 

The  need  for  reversals  in  mass  order  in  the  periodic  table  of  the 
elements  was  apparent  to  Mendeleev.  He  attributed  it  to  faulty 
atomic  weight  data.  He  confidently  expected,  for  example,  that  the 
atomic  mass  of  tellurium  (which  he  placed  fifty-second),  when 
more  accurately  determined  would  turn  out  to  be  lower  than  that  of 
iodine  (which  he  placed  fifty-third).  And,  in  fact,  in  1872  (see  Table 
p.  20)  he  had  convinced  himself  that  the  correct  atomic  mass  of 
tellurium  was  125!  As  the  figures  in  the  modern  tables  show  how- 
ever, tellurium  does  have  a  greater  atomic  mass  than  iodine — the 
reversal  is  real.  Mendeleev  overestimated  the  applicability  of  the 
periodic  law  in  every  detail,  particularly  as  it  had  not  yet  received 
a  physical  explanation.  He  did  not  realize  that  atomic  mass  was  not 
the  underlying  ordering  principle  for  atomic  numbers — it  was  only 
one  physical  property  (with  slightly  imperfect  periodicity).  Satis- 
factory explanations  for  these  reversals  have  been  found  in  modern 
atomic  physics,  and  will  be  explained  in  Unit  6. 

Q1 2    What  is  the  "atomic  number"  of  an  element?  Give  examples 
of  the  atomic  number  of  several  elements. 


SG  17.10-17.12. 


17.7    Electricity  and  matter:  qualitative  studies 


While  chemists  were  applying  Dalton's  atomic  theory  in  the 
first  decade  of  the  nineteenth  century,  another  development  was 
taking  place  which  opened  an  important  path  to  our  understanding 
of  the  atom.  Humphry  Davy  and  Michael  Faraday  made  discoveries 
which  showed  that  electricity  and  matter  are  intimately  related. 
Their  discoveries  in  "electrochemistry"  had  to  do  with  decomposing 
chemical  compounds  by  passing  an  electric  current  through  them. 
This  process  is  called  electrolysis. 

The  study  of  electrolysis  was  made  possible  by  the  invention  of 
the  electric  cell  in  1800  by  the  Italian  scientist  Alessandro  Volta.  As 
we  saw  in  Unit  4,  Volta's  cell  consisted  of  disks  of  different  metals 
separated  from  each  other  by  paper  moistened  with  a  weak  solution 
of  salt.  As  a  result  of  chemical  changes  occurring  in  such  a  cell,  an 
electric  potential  difference  is  established  between  the  metals.  A 
battery  is  a  set  of  several  similar  cells  connected  together.  A  battery 
usually  has  two  terminals,  one  charged  positively  and  the  other 


Some  liquids  conduct  electricity. 
Pure  distilled  water  is  a  poor  con- 
ductor; but  when  certain  substances 
such  as  acids  or  salts  are  dissolved 
in  water,  the  resulting  solutions  are 
good  electrical  conductors.  Gases 
are  not  conductors  under  normal 
conditions,  but  can  be  made 
electrically  conducting  in  the 
presence  of  strong  electric  fields,  or 
by  other  methods.  The  conduction  of 
electricity  in  gases,  vital  to  the  story 
of  the  atom,  will  be  discussed  in 
Chapter  18. 


26 


The  Chemical  Basis  of  the  Atomic  Theory 


Humphry  Davy  (1778-1829)  was  the 
son  of  a  farmer.  In  his  youth  he  worl<ed 
as  an  assistant  to  a  physician,  but  was 
discharged  because  of  his  lil<ing  for 
explosive  chemical  experiments.  He 
became  a  chemist,  discovered  nitrous 
oxide  (laughing  gas),  which  was  later 
used  as  an  anaesthetic,  and  developed 
a  safety  lamp  for  miners  as  well  as  an 
arc  light.  His  work  in  electrochemistry 
and  his  discovery  of  several  elements 
made  him  world-famous;  he  was 
knighted  in  1812.  In  1813  Sir  Hum- 
phry Davy  hired  a  young  man,  Michael 
Faraday,  as  his  assistant  and  took 
him  along  on  an  extensive  trip  through 
France  and  Italy.  It  became  evident  to 
Davy  that  young  Faraday  was  a  man  of 
scientific  genius.  Davy  is  said  to  have 
been  envious,  at  first,  of  Faraday's 
great  gifts.  He  later  said  that  he  be- 
lieved his  greatest  discovery  was 
Faraday. 


charged  negatively.  When  the  terminals  are  connected  to  each  other 
by  means  of  wires  or  other  conducting  materials,  there  is  an  electric 
current  in  the  battery  and  the  materials.  Thus,  the  battery  can 
produce  and  maintain  an  electric  current.  It  is  not  the  only  device 
that  can  do  so,  but  it  was  the  first  source  of  steady  currents. 

Within  a  few  weeks  after  Volta's  announcement  of  his  discovery 
it  was  found  that  water  could  be  decomposed  into  oxygen  and 
hydrogen  by  the  use  of  electric  currents.  At  the  left  is  a  diagram  of 
an  electrolysis  apparatus.  The  two  terminals  of  the  battery  are 
connected,  by  conducting  wires,  to  two  thin  sheets  of  platinum 
("electrodes").  When  these  platinum  sheets  are  immersed  in  ordinary 
water,  bubbles  of  oxygen  appear  at  one  sheet  and  bubbles  of 
hydrogen  at  the  other.  Adding  a  small  amount  of  certain  acids 
speeds  up  the  reaction  without  changing  the  products.  Hydrogen 
and  oxygen  gases  are  formed  in  the  proportion  of  7.94  grams  of 
oxygen  to  1  gram  of  hydrogen,  which  is  exactly  the  proportion  in 
which  these  elements  combine  to  form  water.  Water  had  previously 
been  impossible  to  decompose,  and  had  long  been  regarded  as  an 
element.  Thus  the  ease  with  which  water  was  separated  into  its 
elements  by  electrolysis  dramatized  the  chemical  use  of  electricity, 
and  stimulated  many  other  investigations  of  electrolysis. 

Among  these  investigations,  some  of  the  most  successful  were 
those  of  the  young  English  chemist  Humphry  Davy.  Perhaps  the 
most  striking  of  Davy's  successes  were  those  he  achieved  in  1807 
when  he  studied  the  effect  of  the  current  from  a  large  electric 
battery  upon  soda  and  potash.  Soda  and  potash  were  materials  of 
commercial  importance  (for  example,  in  the  manufacture  of  glass, 
soap,  and  gunpowder)  and  had  been  completely  resistant  to  every 
earlier  attempt  to  decompose  them.  Soda  and  potash  were  thus 
regarded  as  true  chemical  elements— up  to  the  time  of  Davy's  work. 
(See  Dalton's  symbols  for  the  elements  on  p.  10.)  When  electrodes 
connected  to  a  large  battery  were  touched  to  a  solid  lump  of  soda, 
or  to  a  lump  of  potash,  part  of  the  solid  was  heated  to  its  melting 
point.  At  one  electrode  small  globules  of  molten  metal  appeared 
which  burned  brightly  and  almost  explosively  in  air.  When  the 
electrolysis  was  done  in  the  absence  of  air,  the  metalhc  material 
could  be  collected  and  studied.  The  metallic  elements  discovered  in 
this  way  were  called  sodium  and  potassium.  Sodium  was  obtained 
from  soda  (now  called  sodium  hydroxide),  and  potassium  was 
obtained  from  potash  (now  called  potassium  hydroxide).  In  the 
immediately  succeeding  years,  electrolysis  experiments  made  on 
several  previously  undecomposed  "earths"  yielded  the  first  samples 
ever  obtained  of  such  metallic  elements  as  magnesium,  strontium, 
and  barium.  There  were  also  many  other  demonstrations  of  the 
striking  changes  produced  by  the  chemical  activity  of  electricity. 

Q13    Why  was  the  first  electrolysis  of  water  such  a  surprising 
achievement? 

Q14    What  were  some  other  unexpected  results  of  electrolysis? 


Electrolysis 


Student  laboratory  apparatus  like 
that  shown  in  the  sketch  above  can  be 
used  for  experiments  in  electrolysis. 
This  setup  allows  nneasurement  of  the 
amount  of  electric  charge  passing 
through  the  solution  in  the  beaker, 
and  of  the  mass  of  metal  deposited 
on  the  suspended  electrode. 

The  separation  of  elements  by 
electrolysis  is  important  in  industry, 
particularly  in  the  production  of  alumi- 
num. These  photographs  show  the 
large  scale  of  a  plant  where  aluminum 
is  obtained  from  aluminum  ore  in 
electrolytic  tanks. 

(a)  A  row  of  tanks  where  alumi- 
num is  separated  out  of  aluminum  ore. 

(b)  A  closer  view  of  the  front  of 
some  tanks,  showing  the  thick  copper 
straps  that  carry  the  current  for 
electrolysis. 

(c)  A  huge  vat  of  molten  alumi- 
num that  has  been  siphoned  out  of 
the  tanks  is  poured  into  molds. 


28 


The  Chemical  Basis  of  the  Atomic  Theory 


By  chemical  change  we  mean  here 
the  breaking  up  of  molecules  during 
electrolysis,  as  by  gas  bubbles 
rising  at  the  electrodes,  or  by  metal 
deposited  on  it. 


Mass  «  current  x  time 

charge 
"^^  —r: — 5—  >  time 
time 

<^  charge  transferred 


17.8     Electricity  and  matter:  quantitative  studies 

Davy's  work  on  electrolysis  was  mainly  qualitative.  But 
quantitative  questions  were  also  asked.  How  much  chemical  change 
can  be  produced  when  a  certain  amount  of  electric  charge  is  passed 
through  a  solution?  If  the  same  amount  of  charge  is  passed  through 
different  solutions,  how  do  the  amounts  of  chemical  change  com- 
pare? Will  doubling  the  amount  of  electricity  double  the  chemical 
change  effected? 

The  first  answers  to  these  questions  were  obtained  by  Michael 
Faraday,  who  discovered  two  fundamental  and  simple  empirical 
laws  of  electrolysis.  He  studied  the  electrolysis  of  a  solution  of  the 
blue  salt  copper  sulfate  in  water.  The  electric  current  between 
electrodes  placed  in  the  solution  caused  copper  from  the  solution 
to  be  deposited  on  the  negative  electrode  and  oxygen  to  be  liberated 
at  the  positive  electrode.  Faraday  determined  the  amount  of  copper 
deposited  on  the  cathode  by  weighing  the  cathode  before  the  elec- 
trolysis started  and  again  after  a  known  amount  of  current  had 
passed  through  the  solution.  He  found  that  the  mass  of  copper  de- 
posited depends  on  only  two  things:  the  magnitude  of  the  electric 
current  (measured,  say,  in  amperes),  and  the  length  of  time  that  the 
current  was  maintained.  In  fact,  the  mass  of  copper  deposited  is 
directly  proportional  to  both  the  current  and  the  time.  When  either 
was  doubled,  the  mass  of  copper  deposited  was  doubled.  When  both 
were  doubled,  four  times  as  much  copper  was  deposited.  Similar 
results  were  found  in  experiments  on  the  electrolysis  of  many 
different  substances. 

Faraday's  results  may  be  described  by  stating  that  the  amount 
of  chemical  change  produced  in  electrolysis  is  proportional  to  the 
product  of  the  current  and  the  time.  Now,  the  current  (in  amperes) 
is  the  quantity  of  charge  (in  coulombs)  transferred  per  unit  time 
(in  seconds).  The  product  of  current  and  time  therefore  gives  the 
total  charge  in  coulombs  that  has  moved  through  the  cell  during  the 
given  experiment.  We  then  have  Faraday's  first  law  of  electrolysis: 


The  mass  of  an  element  liberated  at  an  electrode  during 
electrolysis  is  proportional  to  the  amount  of  charge  which 
has  passed  through  the  electrode. 

Next  Faraday  measured  the  mass  of  different  elements  liberated 
from  chemical  compounds  by  equal  amount  of  electric  charge.  He 
found  that  the  amount  of  an  element  liberated  from  the  electrolyte 
by  a  given  amount  of  electricity  depends  on  the  element's  atomic 
mass  and  on  its  combining  capacity  (valence).  His  second  law  of 
electrolysis  states: 


This  experimentally  determined 
amount  of  electric  charge,  96,540 
coulombs,  is  now  called  a  faraday. 


If  A  is  the  atomic  mass  of  an  element,  and  if  v  is  its 
valence,  a  transfer  of  96,540  coulombs  of  electric  charge 
liberate  Alv  grams  of  the  element. 


SG  17.13-17.16 


The  table  on  the  next  page  gives  examples  of  Faraday's  second 


Section  17.8 


29 


Masses  of  elements  that  would  be  electrolyzed 

from  compounds  by 

96,540  coulombs  of  electric  charge. 

COMBINING 

MASS  OF  ELEMENT 

ELEMENT 

ATOMIC  MASS  A 

CAPACITY  V 

LIBERATED  (grams) 

Hydrogen 

1.008 

1 

1.008 

Chlorine 

35.45 

1 

35.45 

Oxygen 

16.00 

2 

8.00 

Copper 

63.54 

2 

31.77 

Zinc 

65.37 

2 

32.69 

Aluminum 

26.98 

3 

8.99 

The  values  of  atomic  mass  in  this 
table  are  based  on  a  value  of 
exactly  16  for  oxygen. 


law  of  electrolysis.  In  each  case  the  mass  of  the  element  produced 
by  electrolysis  is  equal  to  its  atomic  mass  divided  by  its  combining 
capacity. 

The  quantity  Alv  was  recognized  to  have  significance  beyond 
just  electrolysis  experiments.  For  example,  the  values  for  Alv  are 
8.00  for  oxygen  and  1.008  for  hydrogen.  The  ratio  is  8.00/1.008  = 
7.94.  But  as  we  have  found  before,  this  is  just  the  ratio  of  masses 
of  oxygen  and  hydrogen  that  combine  to  produce  water.  In  general, 
when  two  elements  combine,  the  ratio  of  their  combining  masses 
is  equal  to  the  ration  of  their  values  for  Alv. 

Faraday's  second  law  of  electrolysis  has  an  important  implica- 
tion. It  shows  that  a  given  amount  of  electric  charge  is  somehow 
closely  connected  with  the  atomic  mass  and  valence  of  an  element. 
The  mass  and  valence  are  characteristic  of  the  atoms  of  the 
element.  Perhaps,  then,  a  certain  amount  of  electricity  is  somehow 
connected  with  an  atom  of  the  element.  The  implication  is  that 
electricity  may  also  be  atomic  in  character.  This  possibility  was 
considered  by  Faraday,  who  wrote  cautiously: 

...  if  we  adopt  the  atomic  theory  or  phraseology,  then  the 
atoms  of  bodies  which  are  equivalents  to  each  other  in 
their  ordinary  chemical  action  have  equal  quantities  of 
electricity  naturally  associated  with  them.  But  I  must 
confess  that  I  am  jealous  of  the  term  atom;  for  though  it 
is  very  easy  to  talk  of  atoms,  it  is  very  difficult  to  form  a 
clear  idea  of  their  nature,  especially  when  compound 
bodies  are  under  consideration. 


In  Chapter  18  you  will  read  about  the  details  of  the  research 
that  did  establish  the  fact  that  electricity  itself  is  atomic  in 
character,  and  that  the  "atoms"  of  electricity  are  part  of  the  atoms 
of  matter.  This  research,  for  which  Faraday's  work  and  his  cautious 
guess  prepared,  helped  make  possible  the  exploration  of  the  structure 
of  the  atom. 

Q15    The  amount  of  an  element  deposited  in  electrolysis 
depends  on  three  factors.  What  are  they? 

Q16    What  are  the  significances  of  the  quantity  Alv  for  an 
element? 


SG  17.17-17.20 


STUDY  GUIDE 


17.1  The  Project  Physics  learning  materials 
particularly  appropriate  for  Chapter  17  include 
the  following: 

Experiment 

Electrolysis 
Activities 

Dalton's  Puzzle 
Electrolysis  of  Water 
Periodic  Table 
Single-electrode  Plating 
Activities  from  the  Scientific  American 
Film  Loops 

Production  of  Sodium  by  Electrolysis 
Articles  of  general  interest  in  Reader  5  are: 
The  Island  of  Research 
The  Sentinel 
Although  most  of  the  articles  in  Reader  5  are 
related  to  ideas  presented  in  Chapter  20,  you 
may  prefer  to  read  some  of  them  earlier. 

17.2  The  chemical  compound  zinc  oxide  (molec- 
ular formula  ZnO)  contains  equal  numbers  of 
atoms  of  zinc  and  oxygen.  Using  values  of  atomic 
masses  from  the  modern  version  of  the  periodic 
table  (on  page  23),  find  the  percentage  by  mass  of 
zinic  in  zinic  oxide.  What  is  the  percentage  of 
oxygen  in  zinc  oxide? 

17.3  The  chemical  compound  zinc  chloride 
(molecular  formula  ZnCla)  contains  two  atoms  of 
chlorine  for  each  atom  of  zinc.  Using  values  of 
atomic  masses  from  the  modern  version  of  the 
periodic  table,  find  the  percentage  by  mass  of 
zinc  in  zinc  chloride. 

17.4  During  the  complete  decomposition  of  a 
5.00-gram  sample  of  ammonia  gas  into  its  com- 
ponent elements,  nitrogen  and  hydrogen,  4.11 
grams  of  nitrogen  were  obtained.  The  molecular 
formula  of  ammonia  is  NH3.  Find  the  mass  of  a 
nitrogen  atom  relative  to  that  of  a  hydrogen 
atom.  Compare  your  result  with  the  one  you 
would  get  by  using  the  values  of  the  atomic 
masses  in  the  modern  version  of  the  periodic 
table.  If  your  result  is  different  from  the  latter 
result,  how  do  you  account  for  the  difference? 

17.5  From  the  information  in  Problem  17.3, 
calculate  how  much  nitrogen  and  hydrogen  are 
needed  to  make  1.2  kg  of  ammonia. 

17.6  //  the  molecular  formula  of  ammonia  were 
falsely  thought  to  be  NH^,  and  you  used  the  result 
of  the  experiment  in  Problem  17.3,  what  value 
would  you  get  for  the  ratio  of  the  mass  of  a 
nitrogen  atom  relative  to  that  of  a  hydrogen 
atom? 

17.7  A  sample  of  nitric  oxide  gas,  weighing 
1.00  g,  after  separation  into  its  components,  is 
found  to  have  contained  0.47  g  of  nitrogen. 
Taking  the  atomic  mass  of  oxygen  to  be  16.00, 
find  the  corresponding  numbers  that  express  the 
atomic  mass  of  nitrogen  relative  to  oxygen  on  the 
respective  assumptions  that  the  molecular  formula 
of  nitric  oxide  is  (a)  NO;  (b)  NO^;  (c)  N.O. 


17.8  Early  data  yielded  8.2  8.0  for  the  mass  ratio 
of  nitrogen  and  oxygen  atoms,  and  17  for  the 
mass  ratio  of  hydrogen  and  oxygen  atoms.  Show 
that  these  results  lead  to  a  value  of  6  for  the 
relative  atomic  mass  of  nitrogen,  provided  that 
the  value  1  is  assigned  to  hydrogen. 

17.9  Given  the  molecular  formulae  HCl.  NaCl. 
CaCl.,,  AICI3,  SnCL,,  PCI,,  finf  possible  combining 
capacities  of  sodium,  calcium,  aluminum,  tin  and 
phosphorus. 

17.10  (a)  Examine  the  modem  periodic  table  of 

elements  and  cite  all  reversals  of  order 
of  increasing  atomic  mass. 

(b)  Restate  the  periodic  law  in  your  own 
words,  not  forgetting  about  these 
reversals. 

17.11  On  the  next  page  is  a  table  of  the  melting 
and  boiling  temperatures  of  the  elements. 

(a)  Plot  these  quantities  against  atomic 
number  in  two  separate  graphs.  Comment 
on  any  periodicity  you  observe  in  the 
plots. 

(b)  Predict  the  values  for  melting  and  boiling 
points  of  the  noble  gases,  which  were 
unknown  in  1872.  Compare  your  predic- 
tions with  the  modern  values  given  in. 
say,  the  Handbook  of  Chemistry  and 
Physics. 

17.12  In  recent  editions  of  the  Handbook  of 
Chemistry  and  Physics  there  are  printed  in  or 
below  one  of  the  periodic  tables  the  valence 
numbers  of  the  elements.  Neglect  the  negative 
valence  numbers  and  plot  (to  element  65)  a 
graph  of  maximum  valences  observed  vs.  atomic 
mass.  What  periodicity  is  found?  Is  there  any 
physical  or  chemical  significance  to  this 
periodicity? 

17.13  According  to  the  table  on  p.  29,  when  about 
96,500  coulombs  of  charge  pass  through  a  water 
solution,  how  much  of  oxygen  will  be  released 

at  the  same  time  when  (on  the  other  electrode) 
1.008  g  of  hydrogen  are  released?  How  much 
oxygen  will  be  produced  when  a  current  of 
3  amperes  is  passed  through  water  for  60  minutes 
(3600  seconds)? 

17.14  If  a  current  of  0.5  amperes  is  passed 
through  molten  zinc  chloride  in  an  electrolytic 
apparatus,  what  mass  of  zinc  will  be  deposited  in 

(a)  5  minutes  (300  seconds); 

(b)  30  minutes; 

(c)  120  minutes? 

17.15  (a)  For  20  minutes  (1200  seconds),  a  cur- 

rent of  2.0  amperes  is  passed  through 
molten  zinc  chloride  in  an  electrolytic 
apparatus.  What  mass  of  chlorine  will 
be  released  at  the  anode? 


30 


(b)  If  the  current  had  been  passed  through 
molten  zinc  iodide  rather  than  molten 
zinc  chloride  what  mass  of  iodine 
would  have  been  released  at  the  anode? 

(c)  Would  the  quantity  of  zinc  deposited  in 
part  (b)  have  been  different  from  what 
it  was  in  part  (a)?  Why? 

(d)  How  would  you  set  up  a  device  for 
plating  a  copper  spoon  with  silver? 

17.16  What  may  be  the  relation  of  Faraday's 
speculation  about  an  "atom  of  electricity"  to  the 
presumed  atomicity  in  the  composition  of  chemical 
elements? 

17.17  96,540  coulombs  in  electrolysis  frees  A 
grams  of  a  monovalent  element  of  atomic  mass 
A  such  as  hydrogen  when  hydrochloric  acid  is 
used  as  electrolyte.  How  much  chlorine  will  be 
released  on  the  other  electrode? 

17.18  If  96,540  coulombs  in  electrolysis  always 
frees  A  grams  of  a  monovalent  element,  A/2 
grams  of  a  divalent  element,  etc.,  what  relation 
does  this  suggest  between  valence  and  "atoms" 
of  electricity? 

17.19  The  idea  of  chemical  elements  composed 
of  identical  atoms  makes  it  easier  to  correlate 
the  phenomena  discussed  in  this  chapter.  Could 
the  phenomena  be  explained  without  using  the 
idea  of  atoms?  Are  chemical  phenomena,  which 
usually  involve  a  fairly  large  quantity  of  material 
(in  terms  of  the  number  of  "atoms"),  sufficient 
evidence  for  Daltons  belief  that  an  element 
consists  of  atoms,  all  of  which  are  exactly 
identical  with  each  other? 

17.20  A  sociologist  recently  wrote  a  book  about 
the  place  of  man  in  modern  society,  called 
Multivalent  Man.  In  general,  what  validity  is 
there  for  using  such  terms  for  sociological  or 
other  descriptions? 

17.21  Which  of  Dalton's  main  postulates  (pp. 
11-12)  were  similar  to  those  in  Greek  atomism 
(pp.  4-5)?  Which  are  quite  different? 

Melting  and  Boiling  Temperatures  of  the 
Elements  Known  by  1872 


Melting  and  Boiling  Temperatures  of  the 
Elements  Known  by  1872  (cont.) 


ATOMIC 

NUMBER 

NAME 

1 

hydrogen 

3 

lithium 

4 

beryllium 

5 

boron 

6 

carbon 

7 

nitrogen 

8 

oxygen 

9 

fluorine 

11 

sodium 

MELTING 

BOILING 

POINT 

POINT 

-259°C 

-253°C 

186 

1340 

1280 

2970 

2300 

2550 

>3350 

4200 

-210 

-196 

-218 

-183 

-223 

-188 

98 

880 

ATOMIC 

MELTING 

BOILING 

NUMBER 

NAME 

POINT 

POINT 

12 

magnesium 

651 

1107 

13 

aluminum 

660 

2057 

14 

silicon 

1420 

2355 

15 

phosphorus 

44 

280 

16 

sulfur 

113 

445 

17 

chlorine 

-103 

-35 

19 

potassium 

62 

760 

20 

calcium 

842 

1240 

22 

titanium 

1800 

>3000 

23 

vanadium 

1710 

3000 

24 

chromium 

1890 

2480 

25 

manganese 

1260 

1900 

26 

iron 

1535 

3000 

27 

cobalt 

1495 

2900 

28 

nickel 

1455 

2900 

29 

copper 

1083 

2336 

30 

zinc 

419 

907 

33 

arsenic 

814 

615 

34 

selenium 

217 

688 

35 

bromine 

-7 

59 

37 

rubidium 

39 

700 

38 

strontium 

774 

1150 

39 

yttrium 

1490 

2500 

40 

zirconium 

1857 

>2900 

41 

niobium 

2500 

3700 

42 

molybdenum 

2620 

4800 

44 

ruthenium 

2450 

2700 

45 

rhodium 

1966 

2500 

46 

palladium 

1549 

2200 

47 

silver 

961 

1950 

48 

cadmium 

321 

767 

49 

indium 

156 

2000 

50 

tin 

232 

2270 

51 

antimony 

631 

1380 

52 

tellurium 

452 

1390 

53 

iodine 

114 

184 

55 

cesium 

29 

670 

56 

barium 

725 

1140 

57 

lanthanum 

826 

58 

cerium 

804 

1400 

68 

erbium 

73 

tantalum 

3000 

4100 

74 

tungsten 

3370 

5900 

76 

osmium 

2700 

>5300 

77 

iridium 

2454 

>4800 

78 

platinum 

1774 

4300 

79 

gold 

1063 

2600 

80 

mercury 

-39 

357 

81 

thallium 

302 

1460 

82 

lead 

327 

1620 

83 

bismuth 

271 

1560 

90 

thorium 

1845 

4500 

92 

uranium 

1133 

ignites 

31 


18.1  The  idea  of  atomic  structure 

18.2  Cathode  rays 

18.3  The  measurement  of  the  charge  of  the  electron: 
Millikan's  experiment 

18.4  The  photoelectric  effect 

18.5  Einstein's  theory  of  the  photoelectric  effect 

18.6  X  rays 

18.7  Electrons,  quanta,  and  the  atom 


33 
34 

37 
40 
43 
48 
54 


The  tube  used  by  J.  J.  Thomson  to  determine  the  charge-to-mass  ratio  of  electrons. 


CHAPTER  EIGHTEEN 


Electrons  and  Quanta 


18.1     The  idea  of  atomic  structure 


The  successes  of  chemistry  in  the  nineteenth  century,  in  ac- 
counting for  combining  proportions  and  in  predicting  chemical 
reactions,  had  proved  to  the  satisfaction  of  most  scientists  that 
matter  is  composed  of  atoms. 

But  there  remained  a  related  question:  are  atoms  really 
indivisible,  or  do  they  consist  of  still  smaller  particles?  We  can  see 
the  way  in  which  this  question  arose  by  thinking  a  little  more  about 
the  periodic  table.  Mendeleev  had  arranged  the  elements  in  the 
order  of  increasing  atomic  mass.  But  the  atomic  masses  of  the 
elements  cannot  explain  the  periodic  features  of  Mendeleev's  table. 
Why,  for  example,  do  the  3rd,  11th,  19th,  37th,  55th,  and  87th 
elements,  with  quite  different  atomic  masses,  have  similar  chemical 
properties?  Why  are  these  properties  somewhat  different  from  those 
of  the  4th,  12th,  20th,  38th,  56th,  and  88th  elements  in  the  hst,  but 
greatly  different  from  the  properties  of  the  2nd,  10th,  18th,  36th, 
54th,  and  86th  elements? 

The  periodicity  in  the  properties  of  the  elements  led  to  specula- 
tion about  the  possibility  that  atoms  might  have  structure,  that 
they  might  be  made  up  of  smaller  pieces.  The  gradual  changes  of 
properties  from  group  to  group  might  suggest  that  some  unit  of 
atomic  structure  is  added,  in  successive  elements,  until  a  certain 
portion  of  the  structure  is  completed.  The  completed  condition 
would  occur  in  the  atom  of  a  noble  gas.  In  an  atom  of  the  next 
heavier  element,  a  new  portion  of  the  structure  may  be  started, 
and  so  on.  The  methods  and  techniques  of  classical  chemistry 
could  not  supply  experimental  evidence  for  such  structure.  In  the 
nineteenth  century,  however,  discoveries  and  new  techniques  in 
physics  opened  the  way  to  the  proof  that  atoms  do,  indeed,  consist 
of  smaller  pieces.  Evidence  piled  up  that  suggested  the  atoms  of  dif- 
ferent elements  differ  in  the  number  and  arrangement  of  these  pieces. 

In  this  chapter,  we  shall  discuss  the  discovery  of  one  structural 
element  which  all  atoms  contain:  the  electron.  Then  we  shall  see 
how  experiments  with  light  and  electrons  led  to  a  revolutionary 


SG  18.1 


These  elements  burn  when  exposed 
to  air;  they  decompose  water,  often 
explosively. 

These  elements  react  slowly  with 
air  or  water. 

These  elements  rarely  combine  with 
any  others. 


33 


34 


Electrons  and  Quanta 


idea  — that  light  energy  is  transmitted  in  discrete  amounts.  In 
Chapter  19,  we  shall  describe  the  discovery  of  another  part  of  the 
atom,  the  nucleus.  Finally  we  shall  show  how  Niels  Bohr  combined 
these  pieces  to  create  a  workable  model  of  the  atom.  The  story 
starts  with  the  discovery  of  cathode  rays. 

18.2    Cathode  rays 


■^      J-- 


Cathode  ray  apparatus 

Substances  which  glow  when 
exposed  to  light  are  called 
fluorescent.  Fluorescent  lights  are 
essentially  Geissler  tubes  with  an 
inner  coating  of  fluorescent  powder. 


cathode 


Bent  Geissler  tube.  The  most  intense 
green  glow  appeared  at  g 


A  Crookes  tube 


In  1855  the  German  physicist  Heinrich  Geissler  invented  a 
vacuum  pump  which  could  remove  enough  gas  from  a  strong  glass 
tube  to  reduce  the  pressure  to  0.01  percent  of  normal  air  pressure. 
It  was  the  first  major  improvement  in  vacuum  pumps  after 
Guericke's  invention  of  the  air  pump,  two  centuries  earlier.  It 
turned  out  to  be  a  critical  technical  innovation  that  opened  new 
fields  to  pure  scientific  research.  Geissler's  friend  Julius  Pliicker 
connected  one  of  Geissler's  evacuated  tubes  to  a  battery.  He  was 
surprised  to  find  that  at  the  very  low  pressure  that  could  be  obtained 
with  Geissler's  pump,  electricity  flowed  through  the  tube.  Pliicker 
used  apparatus  similar  to  that  sketched  in  the  margin.  He  sealed  a 
wire  into  each  end  of  a  strong  glass  tube.  Inside  the  tube,  each  wire 
ended  in  a  metal  plate,  called  an  electrode.  Outside  the  tube,  each 
wire  ran  to  a  source  of  high  voltage.  (The  negative  plate  is  called 
the  cathode,  and  the  positive  plate  is  called  anode.)  A  meter 
indicated  the  current  in  the  tube. 

Pliicker  and  his  student,  Johann  Hittorf,  noticed  that  when  an 
electric  current  passes  through  the  low-pressure  gas  in  a  tube,  the 
tube  itself  glows  with  a  pale  green  color.  Several  other  scientists 
observed  these  effects,  but  two  decades  passed  before  anyone  under- 
took a  thorough  study  of  the  glowing  tubes.  By  1875,  Sir  William 
Crookes  had  designed  new  tubes  for  studying  the  glow  produced 
when  an  electric  current  passes  through  an  evacuated  tube.  When 
he  used  a  bent  tube,  (see  figure  at  the  left)  the  most  intense  green 
glow  appeared  on  the  part  of  the  tube  which  was  directly  opposite 
the  cathode  (at  g).  This  suggested  that  the  green  glow  was  produced 
by  something  which  comes  out  of  the  cathode  and  travels  down  the 
tube  until  it  hits  the  glass.  Another  physicist,  Eugen  Goldstein,  who 
was  studying  the  effects  of  passing  an  electric  current  through  a 
gas  at  low  pressure,  named  whatever  it  was  that  appeared  to  be 
coming  from  the  cathode,  cathode  rays.  For  the  time  being,  it  was 
quite  mysterious  just  what  these  cathode  rays  were. 

To  study  the  nature  of  the  rays,  Crookes  did  some  ingenious 
experiments.  He  reasoned  that  if  cathode  rays  could  be  stopped 
before  they  reached  the  end  of  the  tube,  the  intense  green  glow 
would  disappear.  He  therefore  introduced  barriers  (for  example,  in 
the  form  of  a  Maltese  cross,  as  in  the  sketch  in  the  margin).  A 
shadow  of  the  barrier  appeared  in  the  midst  of  the  green  glow  at 
the  end  of  the  tube.  The  cathode  seemed  to  act  like  a  source  which 
radiates  a  kind  of  light;  the  cross  acted  like  a  barrier  blocking  the 
light.  Because  the  shadow,  cross,  and  cathode  appeared  along  one 
straight  line,  Crookes  concluded  that  the  cathode  rays,  like  light 
rays,  travel  in  straight  lines.  Next,  Crookes  moved  a  magnet  near 


Section  18.2 


35 


the  tube,  and  the  shadow  moved.  Thus  he  found  that  magnetic 
fields  deflected  the  paths  of  cathode  rays  (which  does  not  happen 
with  light). 

In  the  course  of  many  experiments,  Crookes  found  the  following 
properties  of  cathode  rays: 

(a)  No  matter  what  material  the  cathode  is  made  of,  it  produces 
rays  with  the  same  properties. 

(b)  In  the  absence  of  a  magnetic  field,  the  rays  travel  in  straight 
lines  perpendicular  to  the  surface  that  emits  them. 

(c)  A  magnetic  field  deflects  the  path  of  the  cathode  rays. 

(d)  The  rays  can  produce  some  chemical  reactions  similar  to  the 
reactions  produced  by  light;  for  example,  certain  silver  salts  change 
color  when  hit  by  the  rays. 

In  addition,  Crookes  suspected  (but  did  not  succeed  in  showing) 
that  (e)  charged  objects  deflect  the  path  of  cathode  rays. 

Physicists  were  fascinated  by  the  cathode  rays.  Some  thought 
that  the  rays  must  be  a  form  of  light,  because  they  have  so  many 
of  the  properties  of  light:  they  travel  in  straight  lines,  and  produce 
chemical  changes  and  fluorescent  glows  just  as  light  does.  Accord- 
ing to  Maxwell's  theory  of  electricity  and  magnetism,  light  consists 
of  electromagnetic  waves.  So  the  cathode  rays  might,  for  example, 
be  electromagnetic  waves  of  frequency  much  higher  than  that  of 
visible  light. 

However,  magnetic  fields  do  not  bend  light;  they  do  bend  the 
path  of  cathode  rays.  In  Chapter  14  we  described  how  magnetic 
fields  exert  forces  on  currents,  that  is,  on  moving  electric  charges. 
Since  a  magnetic  field  deflects  cathode  rays  in  the  same  way  that  it 
deflects  negative  charges,  some  physicists  believed  that  cathode 
rays  consisted  of  negatively  charged  particles. 

The  controversy  over  whether  cathode  rays  are  a  force  of 
electromagnetic  waves  or  a  stream  of  charged  particles  continued 
for  25  years.  Finally,  in  1897,  J.  J.  Thomson  made  a  series  of 
experiments  which  convinced  physicists  that  the  cathode  rays  are 
negatively  charged  particles.  Details  of  Thomson's  experiment  and 
calculations  are  given  on  page  36. 

It  was  then  well-known  that  the  paths  of  charged  particles  are 
affected  by  both  magnetic  and  electric  fields.  By  assuming  that 
the  cathode  rays  were  negatively  charged  particles,  Thomson  could 
predict  what  should  happen  to  the  cathode  rays  when  they  passed 
through  such  fields.  For  example,  it  should  be  possible  to  balance 
the  deflection  of  a  beam  of  cathode  rays  by  a  magnetic  field  by 
turning  on  an  electric  field  of  just  the  right  magnitude  and 
direction.  As  page  36  indicates,  the  predictions  were  verified,  and 
Thomson  could  therefore  conclude  that  the  cathode  rays  were 
indeed  made  up  of  negatively  charged  particles.  He  was  then  able 
to  calculate,  from  the  experimental  data,  the  ratio  of  the  charge  of 
a  particle  to  its  mass.  This  ratio  is  denoted  by  qlm,  where  q  is  the 
charge  and  m  is  the  mass  of  the  particle. 

Thomson  found  that  the  rays  coming  from  cathodes  made  of 
different  materials  all  had  the  same  value  of  qlm,  namely  1.76  x 
10''  coulombs  per  kilogram. 


J.  J.  Thomson  later  observed  this 
to  be  possible. 


Sir  Joseph  John  Thomson  (1856- 
1940),  one  of  the  greatest  British 
physicists,  attended  Owens  College 
in  Manchester,  England  and  then 
Cambridge  University.  He  worked 
on  the  conduction  of  electricity 
through  gases,  on  the  relation  be- 
tween electricity  and  matter  and  on 
atomic  models.  His  greatest  single 
contribution  was  the  discovery  of  the 
electron.  He  was  the  head  of  the  fa- 
mous Cavendish  Laboratory  at  Cam- 
bridge University,  where  one  of  his 
students  was  Ernest  Rutherford. 


Thomson's  q/m  Experiment 


J.  J.  Thomson  measured  the  ratio  of  charge  q  to  mass  m  for  cathode-ray  particles  by  means  of  the 
evacuated  tube  shown  in  the  photograph  on  page  32.  A  high  voltage  applied  between  two  electrodes  in  the 
left  end  of  the  tube  produced  cathode  rays.  Those  rays  that  passed  through  both  slotted  cylinders  in  the 
narrow  neck  of  the  tube  formed  a  nearly  parallel  beam.  The  beam  produced  a  spot  of  light  on  a  fluorescent 
coating  inside  the  large  end  of  the  tube  at  the  right. 


The  path  of  the  beam  was  deflected  by  an  electric  field  applied  between  two  horizontal  plates  in  the 
mid-section  of  the  tube;  (note  that  direction  of  electric  field  '^ \s  upward  along  plane  of  page): 


G^~*^    -f. 


The  beam's  path  was  also  deflected  when  there  was  no  electric  field  but  when  a  magnetic  field  was  set 
up  by  means  of  a  pair  of  current-carrying  wire  coils  placed  around  the  midsection  of  the  tube;  (the  direction 
of  the  magnetic  field  ^  is  into  the  plane  of  the  page): 


When  only  the  magnetic  field  ^  is  turned  on,  particles  in  the  beam,  having  charge  q  and  speed  v,  would 
experience  a  force  Bqv;  because  the  force  is  always  perpendicular  to  the  direction  of  the  velocity  vector, 
the  beam  would  be  deflected  in  a  nearly  circular  arc  of  radius  R  as  long  as  it  is  in  the  nearly  uniform 
magnetic  field.  If  the  particles  in  the  beam  have  mass  m,  they  must  be  experiencing  a  centripetal  force 
mv'/R  while  moving  in  a  circular  arc.  Since  the  centripetal  force  is  provided  by  the  magnetic  force  Bqv. 
we  can  write  Bqv  =  mv'-R.  Rearranging  terms:  q/m  =  v/BR- 

B  can  be  calculated  from  the  geometry  of  the  coils  and  the  electric  current  in  them.  R  can  be  found 
geometrically  from  the  displacement  of  the  beam  spot  on  the  end  of  the  tube.  To  determine  v,  Thomson 
applied  the  electric  field  and  the  magnetic  field  at  the  same  time,  and  arranged  the  directions  and  strengths 
of  the  two  fields  so  that  the  electric  field  ^exerted  a  downward  force  Eq  on  the  beam  particles  exactly  equal 
to  the  upward  force  Bqv  due  to  the  magnetic  field -as  seen  by  the  fact  that  the  beam,  acted  on  by  both 
fields  in  opposing  ways,  goes  along  a  straight  line. 


If  the  magnitudes  of  the  forces  due  to  the  electric  and  magnetic  fields  are  equal,  then  Eq  =  Bqv.  Solving 
for  V  we  have:  v  ^  E/B.  E  can  be  calculated  from  the  separation  of  the  two  plates  and  the  voltage  between 
them;  so  the  speed  of  the  particles  v  can  be  determined.  Now  all  the  terms  on  the  right  of  the  earlier  equation 

for  n/m  arp  knn\A/n    anH  n/m  ran  V\p  rnmn\iior{ 


Section  18.3  37 

Thus,  it  was  clear  that  cathode  rays  must  be  made  of  something 
all  materials  have  in  common.  Thomson's  negatively  charged 
particles  were  later  called  electrons.  The  value  of  qlm  for  the 
cathode  ray  particles  was  about  1800  times  larger  than  the  values 
of  qlm  for  hydrogen  ions,  9.6  x  10'  coulombs  per  kilogram  as 
measured  in  electrolysis  experiments  of  the  kind  we  discussed  in 
Sec.  17.8.  (See  table  on  p.  29.)  Thomson  concluded  from  these  results        SG  18.2 
that  either  the  charge  of  the  cathode  ray  particles  is  much  greater 
than  that  of  the  hydrogen  ion,  or  the  mass  of  the  cathode  ray 
particles  is  much  less  than  the  mass  of  the  hydrogen  ion. 

Thomson  also  made  measurements  of  the  charge  q  on  these 
negatively  charged  particles  with  methods  other  than  those 
involving  deflection  by  electric  and  magnetic  fields.  Although  these 
experiments  were  not  very  accurate,  they  were  good  enough  to 
indicate  that  the  charge  of  a  cathode  ray  particle  was  the  same  or 
not  much  diff"erent  from  that  of  the  hydrogen  ion  in  electrolysis.  In 
view  of  the  small  value  of  qlm,  Thomson  was  therefore  able  to 
conclude  that  the  mass  of  cathode  ray  particles  is  much  less  than 
the  mass  of  hydrogen  ions. 

In  short,  the  cathode  ray  particles,  or  electrons,  were  found  to 
have  two  important  properties:  (1)  they  were  emitted  by  a  wide 
variety  of  cathode  materials,  and  (2)  they  were  much  smaller 
in  mass  than  the  hydrogen  atom,  which  has  the  smallest  known 
mass.  Thomson  therefore  concluded  that  the  cathode  ray  particles 
form  a  part  of  all  kinds  of  matter.  He  suggested  that  the  atom  is 
not  the  ultimate  hmit  to  the  subdivision  of  matter,  and  that  the 
electron  is  part  of  an  atom,  that  it  is.  perhaps  even  a  basic  building 
block  of  atoms.  We  now  know  that  this  is  correct:  the  elctron  — 
whose  existence  Thomson  had  first  proved  by  quantitative  experi- 
ment—is one  of  the  fundamental  or  "elementary"  particles  of  which 
matter  is  made. 

In  the  article  in  which  he  published  his  discovery,  Thomson  also 
speculated  on  the  ways  in  which  such  particles  might  be  arranged 
in  atoms  of  different  elements,  in  order  to  account  for  the  periodicity 
of  the  chemical  properties  of  the  elements.  Although,  as  we  shall 
see  in  the  next  chapter,  he  did  not  say  the  last  word  about  the 
arrangement  and  number  of  electrons  in  the  atom,  he  did  say  the 
first  word  about  it. 

Q1     What  was  the  most  convincing  evidence  that  cathode  rays 
were  not  electromagnetic  radiation? 

Q2     What  was  the  reason  given  for  the  ratio  qlm  for  electrons 
being  1800  times  larger  than  qlm  for  hydrogen  ions? 

Q3     What  were  two  main  reasons  for  Thomson's  beUef  that 
electrons  may  be  "building  blocks"  from  which  all  atoms  are  made? 


18.3    The  measurement  of  the  charge  of  the  electron:  Millikan's 
experiment 

After  the  ratio  of  charge  to  the  mass  (qlm)  of  the  electron  had 
been  determined,  physicists  tried  to  measure  the  value  of  the 


38 


Electrons  and  Quanta 


From  now  on  we  denote  the  magni- 
tude of  the  charge  of  the  electron 
by  q,: 

q,  =  ^.6x  10  ''coul. 

The  sign  of  the  charge  is  negative 
for  the  electron. 


SG  18.3 


In  1964,  an  American  physicist, 
Murray  Gell-Mann,  suggested  that 
particles  with  charge  equal  to  1/3 
or  2/3  of  q   might  exist.  He  named 
these  particles  "quarks"— the  word 
comes  from  James  Joyce's  novel 
Finnegan's  Wake.  Quarks  are  now 
being  looked  for  in  cosmic-ray  and 
bubble-chamber  experiments. 


Thomson  found  that 

q,./m  =  1.76  X  10"  coul/kg. 

According  to  Millikan's  experiment 
the  magnitude  of  q,.  is   1.6  x  10  '"  coul. 

Therefore,  the  mass  of  an  electron  is: 

_     1.6  X  10'"  coul 
"^      1.76  X  10"  coul/kg 

=  0.91  X  10   '"  kg 

(Mass  of  a  hydrogen  ion  is  1.66  x 
10  -"  kg.  This  is  approximately  the 
value  of  one  "atomic  mass  unit.") 


charge  q  itself  in  a  variety  of  ways.  If  the  charge  could  be  deter- 
mined, the  mass  of  the  electron  could  be  found  from  the  known 
value  of  qlm.  In  the  years  between  1909  and  1916,  the  American 
physicist  Robert  A.  Milhkan  succeeded  in  measuring  the  charge  of 
the  electron.  This  quantity  is  one  of  the  fundamental  constants  of 
physics;  it  comes  up  again  and  again  in  atomic  and  nuclear  physics 
as  well  as  in  electricity  and  electromagnetism. 

Millikan's  "oil-drop  experiment"  is  still  one  of  the  nicest 
experiments  that  students  can  do,  and  is  described  in  general  out- 
line on  page  39.  He  found  that  the  electric  charge  that  a  small 
object  such  as  an  oil  drop  can  pick  up  is  always  a  simple  multiple 
of  a  certain  minimum  value.  For  example,  the  charge  may  have 
the  value  -4.8  x  10"'^  coulombs,  or  -1.6  x  10~'^  coulombs,  or  -6.4  x 
10"'^  coulombs,  or  -1.6  x  10"'^  coulombs.  But  it  never  has  a  charge 
of,  say,  —2.4  x  10~'^  coulombs,  and  it  never  has  a  value  smaller 
than  —1.6  x  10""*  coulombs.  In  other  words,  electric  charges  always 
come  in  multiples  (I,  2,  3  .  .  .)  of  1.6  x  10"'^  coulombs,  a  quantity 
often  symbolized  by  q^.  Milhkan  took  this  minimum  charge  to  be 
the  amount  of  charge  of  a  single  electron. 

The  magnitude  of  the  charge  of  nuclei  or  atomic  and  molecular 
ions  has  also  turned  out  always  to  come  in  multiples  of  the  electron 
charge  q^.  For  example,  when  a  chemist  refers  to  a  "doubly  charged 
oxygen  ion,"  he  means  that  the  magnitude  of  the  charge  of  the  ion 
is  2qg,  or  3.2  x  10"'*  coulombs. 

Note  that  Milhkan's  experiments  did  not  prove  that  no  charges 
smaller  than  q^  can  exist.  All  we  can  say  is  that  no  experiment  has 
yet  proved  the  existence  of  smaller  charges.  There  are  recent 
theoretical  reasons  to  expect  that  in  some  very  high-energy  experi- 
ments, another  elementary  particle  of  charge  of  j  q^  may 
eventually  be  discovered;  but  no  such  "fractional"  charge  is 
expected  to  be  found  on  nuclei,  ions,  or  droplets. 

In  everyday  life,  the  electric  charge  one  meets  is  so  large 
compared  to  that  on  one  electron  that  one  can  think  of  such  charges 
or  currents  as  being  continuous— just  as  one  usually  thinks  of  the 
flow  of  water  in  a  river  as  continuous  rather  than  as  a  flow  of 
individual  molecules.  A  current  of  one  ampere,  for  example,  is 
equivalent  to  the  flow  of  6.25  x  10"*  electrons  per  second.  The 
"static"  electric  charge  one  accumulates  by  shuffling  over  a  rug  on 
a  dry  day  consists  of  something  like  10'^  electron  charges. 

Since  the  work  of  Millikan,  a  wide  variety  of  other  experiments 
involving  many  diff'erent  fields  within  physics  have  all  pointed  to 
the  same  basic  unit  of  charge  as  being  fundamental  in  the  structure 
and  behavior  of  atoms,  nuclei,  and  particles  smaller  than  these.  For 
example,  it  has  been  shown  directly  that  cathode  ray  particles  carry 
this  basic  unit  of  charge  — that  they  are,  in  other  words,  electrons. 

By  combining  Millikan's  value  for  the  electron  charge  q^  with 
Thomson's  value  for  the  ratio  of  charge  to  mass  {qjm.),  we  can 
calculate  the  mass  of  a  single  electron  (see  margin).  The  result 
found  is  that  the  mass  of  the  electron  is  about  10"''"  kilograms. 
From  electrolysis  experiments  (see  Sec.  17.8)  we  know  that  the 


vAciftBte 


Millikan's  Oil-drop  Experiment 

R.  A.  Millikan's  own  apparatus  (about  1910) 
for  measuring  the  charge  of  the  electron  is  seen 
in  the  photograph  above.  A  student  version  of 
Millikan's  apparatus  shown  in  the  lower 
photograph  was  taken  in  a  laboratory  period 
of  the  Projects  Physics  Course. 

In  principle  Millikan's  experiment  is  simple; 
the  essential  part  of  the  apparatus  is  sketched 
above.  When  oil  is  sprayed  into  the 
chamber  containmg  two  horizontal  plates, 
the  minute  droplets  formed  are  electrically 
charged  as  they  emerge  from  the  spray  nozzle. 
The  charge  of  a  droplet  is  what  must  be 
measured.  Consider  a  small  oil  drop  of  mass  m 
carrying  an  electric  charge  Q.  It  is  situated 
between  the  two  horizontal  plates  that  are 
separated  by  a  distance  d  and  at  an  electrical 
potential  difference  V.  There  will  be  a  uniform 
electric  field  ^  between  the  plates,  of  strength 
V/6  (see  Sec.  14.8).  This  field  can  be  adjusted 
so  that  the  electrical  force  qE' exerted  upward 


on  the  drop's  charge  will  balance  the  force  maf, 
exerted  downward  by  gravity.  In  this  balanced 
situation, 


therefore 
or 


el        'grav 


qE  =  mag 
q  =  ma,j/E 


The  mass  of  the  drop  can,  in  principle,  be 
determined  from  its  radius  and  the  density  of 
the  oil  from  which  it  was  made.  Millikan  had  to 
measure  these  quantities  by  an  indirect  method, 
but  it  is  now  possible  to  do  the  experiment 
with  small  manufactured  polystyrene  spheres 
instead  of  oil  drops.  Their  mass  is  known,  so 
that  some  of  the  complications  of  the  original 
experiment  can  be  avoided.  Millikan's  remark- 
able result  was  that  the  charge  q  on  objects 
such  as  an  oil  drop  is  always  a  multiple  (1,  2, 
3  .  .  .)  times  a  smallest  charge,  which  he 
identified  with  the  charge  of  one  electron  (Qp). 


40  Electrons  and  Quanta 

charge-to-mass  ratio  of  a  hydrogen  ion  is  1836  times  smaller  than 
the  charge-to-mass  ratio  of  an  electron.  Since  an  electron  and  a 
hydrogen  ion  form  a  neutral  hydrogen  atom  when  they  combine,  it 
is  reasonable  to  expect  that  they  have  equal  and  opposite  charges. 
We  may  therefore  conclude  that  the  mass  of  the  hydrogen  ion  is 
1836  times  as  great  as  the  mass  of  the  electron:  that  is  the  mass 
of  the  hydrogen  ion  is  1836  x  0.91  x  IQ-^o  kg  =  1.66  x  IQ-'  kg.  This 
is  approximately  the  value  of  one  atomic  mass  unit. 

Q4     Oil  drops  pick  up  different  amounts  of  electric  charge.  On 
what  basis  did  Millikan  decide  that  the  lowest  charge  he  found  was 
actually  just  one  electron  charge? 


18.4    The  photoelectric  effect 

In  1887  the  German  physicist  Heinrich  Hertz  was  testing 
Maxwell's  theory  of  electromagnetic  waves.  He  noticed  that  a 
metalhc  surface  can  emit  electric  charges  when  hght  of  very  short 
wavelength  falls  on  it.  Because  light  and  electricity  are  both 
involved,  the  name  photoelectric  effect  was  given  to  this  phenome- 
non. When  the  electric  charges  so  produced  passed  through  electric 
and  magnetic  fields,  the  direction  of  their  paths  was  changed  in 
the  same  rays  as  the  path  of  cathode  rays.  It  was  therefore  deduced 
that  the  electric  charges  consist  of  negatively  charged  particles.  In 
1898,  J.  J.  Thomson  measured  the  value  of  the  ratio  qlm  for  these 
photoelectrically  emitted  particles  with  the  same  method  that  he 
used  for  the  cathode  ray  particles.  He  got  the  same  value  for  the 
particles  ejected  in  the  photoelectric  effect  as  he  had  earlier  for 
the  cathode-ray  particles.  By  means  of  these  experiments  (and 
others)  the  photoelectric  particles  were  shown  to  have  the  same 
properties  as  electrons.  In  fact,  we  must  consider  them  to  be 
ordinary  electrons,  although  they  are  often  referred  to  as  photo- 
electrons,  to  indicate  their  origin.  Later  work  showed  that  all 
substances,  sohds,  Uquids  and  gases,  exhibit  the  photoelectric  effect 
under  appropriate  conditions.  It  is,  however,  convenient  to  study  the 
effect  with  metallic  surfaces. 

The  photoelectric  effect,  which  we  shall  be  stud\ing  in  greater 
detail,  has  had  an  important  place  in  the  development  of  atomic 
physics.  The  effect  could  not  be  explained  in  terms  of  the  ideas  of 
physics  we  have  studied  so  far.  New  ideas  had  to  be  introduced  to 
account  for  the  experimental  results.  In  particular,  a  revolutionary 
concept  was  introduced  — that  of  quanta.  A  new  branch  of  physics  — 
quantum  t/ieor?y  — developed  at  least  in  part  because  of  the 
explanation  provided  for  the  photoelectric  effect. 

The  basic  inforination  for  studying  the  photoelectric  effect 
comes  from  two  kinds  of  measurements:  measurements  of  the 
photoelectric  current  (the  number  of  photoelectrons  emitted  per 
unit  time);  and  measurements  of  the  kinetic  energies  with  which 
the  photoelectrons  are  emitted. 


Section  18.4 


41 


The  photoelectric  current  can  be  studied  with  an  apparatus 
like  that  sketched  in  Fig.  (a)  in  the  margin.  Two  metal  plates,  C  and 
A,  are  sealed  inside  a  well-evacuated  quartz  tube.  (Quartz  glass  is 
transparent  to  ultraviolet  light  as  well  as  visible  light.)  The  two 
plates  are  connected  to  a  source  of  potential  difference  (for 
example,  a  battery).  In  the  circuit  is  also  an  ammeter.  As  long  as 
light  strikes  plate  C,  as  in  Fig.  (b),  electrons  are  emitted  from  it.  If 
the  potential  of  plate  A  is  positive  relative  to  plate  C,  these  emitted 
photoelectrons  will  accelerate  to  plate  A.  (Some  emitted  electrons 
will  reach  plate  A  even  if  it  is  not  positive  relative  to  C.)  The  result- 
ing "photoelectric"  cun-ent  is  indicated  by  the  ammeter.  The  result 
of  the  experiment  is  that  the  stronger  the  beam  of  light  of  a  given 
color  (frequency),  the  greater  the  photoelectric  current. 


The  best  way  to  study  this  part -as 
most  other  parts  — of  physics  is 
really  by  doing  the  experiments 
discussed! 


Schematic  diagram  of  apparatus  for 
photoelectric  experiments. 


(a)      L_3 


Any  metal  used  as  the  plate  C  shows  a  photoelectric  effect,  but 
only  if  the  light  has  a  frequency  greater  than  a  certain  value.  This 
value  of  the  frequency  is  called  the  threshold  frequency  for  that 
metal.  Different  metals  have  different  threshold  frequencies.  If  the 
incident  Ught  has  a  frequency  lower  than  the  threshold  frequency, 
no  photoelectrons  are  emitted,  no  matter  how  great  the  intensity  of 
the  light  is  or  how  long  the  light  is  left  on!  This  is  the  first  of  a  set 
of  surprising  discoveries. 

The  kinetic  energies  of  the  electrons  can  be  measured  in  a 
shghtly  modified  version  of  the  apparatus,  sketched  in  Fig.  (c) 
below.  The  battery  is  reversed  so  that  the  plate  A  now  tends  to  repel 
the  photoelectrons.  The  voltage  can  be  changed  from  zero  to  a  value 
just  large  enough  to  keep  any  electrons  from  reaching  the  plate  A, 
as  indicated  in  Fig.  (d). 


42 


Electrons  and  Quanta 


SG  18.4 


In  Sec.  14.8,  we  saw  that  the  change 
in  potential  energy  of  a  charge  is 
given  by  Vxq.  In  Unit  3  we  saw  that 
(in  the  absence  of  friction)  the 
decrease  in  kinetic  energy  in  a 
system  is  equal  to  the  increase 
in  its  potential  energy. 


^/Ce^uevry  of  it^oe*Jr  n^^r 


Photoelectric  effect:  maximum  kinetic 
energy  of  the  electrons  as  a  function 
of  the  frequency  of  the  incident  light; 
different  metals  yield  lines  that  are 
parallel,  but  have  different  threshold 
frequencies. 


When  the  voltage  across  the  plates  is  zero,  the  meter  will 
indicate  a  current,  showing  that  the  photoelectrons  emerge  from 
the  metallic  surface  with  kinetic  energy  and  so  can  reach  plate  A. 
As  the  repelling  voltage  is  increased  the  photoelectric  current 
decreases  until  a  certain  voltage  is  reached  at  which  the  current 
becomes  zero,  as  indicated  in  Fig.  (d)  above.  This  voltage,  which  is 
called  the  stopping  voltage,  is  a  measure  of  the  maximum  kinetic 
energy  of  the  emitted  photoelectrons  (KE,„qj.).  If  the  stopping  voltage 
is  denoted  by  Vgi^p,  this  maximum  kinetic  energy  is  given  by  the 
relation: 

XF       =V       a 

'■^'-'max  '  stop  rie 

The  results  may  be  stated  more  precisely.  For  this  purpose  let 
us  now  number  the  important  experimental  results  to  make  it  more 
convenient  to  discuss  their  theoretical  interpretation  later. 

(1)  A  substance  shows  a  photoelectric  effect  only  if  the  incident 
light  radiation  has  a  frequency  above  a  certain  value  called  the 
threshold  frequency  (symbol  /„). 

(2)  If  Ught  of  a  given  frequency  does  produce  a  photoelectric 
effect,  the  photoelectric  current  from  the  surface  is  proportional  to 
the  intensity  of  the  light  falling  on  it. 

(3)  If  Ught  of  a  given  frequency  liberates  photoelectrons,  the 
emission  of  these  electrons  is  immediate.  The  time  interval  between 
the  incidence  of  the  Ught  on  the  metallic  surface  and  the  appear- 
ance of  electrons  has  been  found  to  be  at  most  3  x  10~"  sec.  and  is 
probably  much  less.  In  some  experiments,  the  light  intensity  used 
was  so  low  that,  according  to  the  classical  theory,  it  should  take 
several  hundred  seconds  for  an  electron  to  accumulate  enough 
energy  from  the  Ught  to  be  emitted.  But  even  in  these  cases 
electrons  are  sometimes  emitted  about  a  bilUonth  of  a  second  after 
the  light  strikes  the  surface. 

(4)  The  maximum  kinetic  energy  of  the  photoelectrons  increases 
in  direct  proportion  to  the  frequency  of  the  Ught  which  causes 
their  emission,  and  is  independent  of  the  intensity  of  the  incident 
light.  The  way  in  which  the  maximum  kinetic  energy  of  the 
electrons  varies  with  the  frequency  of  the  incident  light  is  shown  in 
the  margin  where  the  symbols  (/o)i,  (/o)2  and  (/„)3  stand  for  the 
different  threshold  frequencies  of  three  different  substances.  For 
each  substance,  the  experimental  data  points  fall  on  a  straight  Une. 
All  the  lines  have  the  same  slope. 

What  is  most  astonishing  about  the  results  is  that  photo- 
electrons are  emitted  if  the  light  frequencies  are  a  little  above  the 
threshold  frequency,  no  matter  how  weak  the  beam  of  light  is;  but 
if  the  light  frequencies  are  just  a  bit  below  the  threshold  frequency, 
no  electrons  are  emitted  no  matter  how  great  the  intensity  of  the 
light  beam  is. 

Findings  (1),  (3)  and  (4)  could  not  be  explained  on  the  basis  of 
the  classical  electromagnetic  theory  of  light.  There  was  no  way  in 
which  a  low-intensity  train  of  light  waves  spread  out  over  a  large 
number  of  atoms  could,  in  a  very  short  time  interval,  concentrate 


Section  18.5 


43 


enough  energy  on  one  electron  to  knock  the  electron  out  of  the 
metal. 

Furthermore,  the  classical  wave  theory  was  unable  to  account 
for  the  existence  of  a  threshold  frequency.  There  seemed  to  be  no 
reason  why  a  sufficiently  intense  beam  of  low-frequency  radiation 
would  not  be  able  to  produce  photoelectricity,  if  low-intensity 
radiation  of  higher  frequency  could  produce  it.  Similarly,  the  classi- 
cal theory  was  unable  to  account  for  the  fact  that  the  maximum 
kinetic  energy  of  the  photoelectrons  increases  linearly  with  the 
frequency  of  the  light  but  is  independent  of  the  intensity.  Thus, 
the  photoelectric  effect  posed  a  challenge  which  the  classical  wave 
theory  of  light  could  not  meet. 

Q5     Light  falling  on  a  certain  metal  surface  causes  electrons  to 
be  emitted.  What  happens  to  the  photoelectric  current  as  the  in- 
tensity of  the  light  is  decreased? 

Q6     What  happens  as  the  frequency  of  the  light  is  decreased? 

Q7     Sketch  a  rough  diagram  of  the  equipment  and  circuit  used 
to  demonstrate  the  main  facts  of  photoelectricity. 


18.5    Einstein's  theory  of  the  photoelectric  effect 

The  explanation  of  the  photoelectric  effect  was  the  major  work 
cited  in  the  award  to  Albert  Einstein  of  the  Nobel  Prize  in  physics 
for  the  year  1921.  Einstein's  theory,  proposed  in  1905,  played  a 
major  role  in  the  development  of  atomic  physics.  The  theory  was 
based  on  a  daring  proposal.  Not  only  were  most  of  the  experimental 
details  still  unknown  in  1905,  but  the  key  point  of  Einstein's 
explanation  was  contrary  to  the  classical  ideas  of  the  time. 

Einstein  assumed  that  energy  of  hght  is  not  distributed  evenly 
over  the  whole  expanding  wave  front  (as  is  assumed  in  the  classical 
theory),  but  rather  remains  concentrated  in  separate  "lumps." 
Further,  the  amount  of  energy  in  each  of  these  regions  is  not  just 
any  amount,  but  a  definite  amount  of  energy  which  is  proportional 
to  the  frequency  /  of  the  wave.  The  proportionaUty  factor  is  a 
constant,  denoted  by  h,  and  is  called  Planck's  constant,  for  reasons 
which  will  be  discussed  later.  Thus,  in  this  model,  the  Hght  energy 
in  a  beam  of  frequency  /  comes  in  pieces,  each  of  amount  h  x  f. 
The  amount  of  radiant  energy  in  each  piece  is  called  a  quantum 
of  energy.  It  represents  the  smallest  quantity  of  energy  of  light  of 
that  frequency.  The  quantum  of  hght  energy  was  later  called  a 
photon. 

There  is  no  explanation  clearer  or  more  direct  than  Einstein's. 
We  quote  from  his  first  paper  (1905)  on  this  subject,  changing  only 
the  notation  used  there  to  make  it  coincide  with  usual  current 
practice  (including  our  own  notation): 

.  .  .  According  to  the  idea  that  the  incident  hght  consists 
of  quanta  with  energy  hf,  the  ejection  of  cathode  rays  by 
light  can  be  understood  in  the  following  way.  Energy 


See  the  articles  "Einstein"  and 
"Einstein  and  some  Civilized  Dis- 
contents" in  Reader  5. 


/]  =  6.6  X  10"  joule-sec 


44 


Electrons  and  Quanta 


SG  18.5 


Each  electron  must  be  given  a 
minimum  energy  to  emerge  from  the 
surface  because  it  must  do  woric 
against  the  forces  of  attraction  as  it 
leaves  the  rest  of  the  atoms. 


This  equation  is  usually  called 
Einstein's  photoelectric  equation. 


SG  18.6-18.8. 


How  Einstein's  theory  explains  the 
photoelectric  effect: 

(1)  No  photoelectric  emission  below 
threshold  frequency.  Reason:  low- 
frequency  photons  don't  have 
enough  energy  to  provide  electrons 
with  KE  sufficient  to  leave  the  metal. 

(2)  Current  ^•-  light  intensity.  Reason: 
one  photon  ejects  one  electron. 

SG  18.9,  18.10 


quanta  penetrate  the  surface  layer  of  the  body,  and  their 
energy  is  converted,  at  least  in  part,  into  kinetic  energy  of 
electrons.  The  simplest  picture  is  that  a  light  quantum 
gives  up  all  its  energy  to  a  single  electron;  we  shall 
assume  that  this  happens.  The  possibiUty  is  not  to  be  ex- 
cluded, however,  that  electrons  receive  their  energy  only 
in  part  from  the  light  quantum.  An  electron  provided  with 
kinetic  energy  inside  the  body  may  have  lost  part  of  its 
kinetic  energy  by  the  time  it  reaches  the  surface.  In  addi- 
tion, it  is  to  be  assumed  that  each  electron,  in  leaving  the 
body,  has  to  do  an  amount  of  work  W  (which  is  character- 
istic of  the  body).  The  electrons  ejected  directly  from  the 
surface  and  at  right  angles  to  it  will  have  the  greatest 
velocities  perpendicular  to  the  surface.  The  maximum  kinetic 
energy  of  such  an  electron  is 


KE, 


hf-W 


If  the  body  plate  C  is  charged  to  a  positive  potential, 
V,,„,,  just  large  enough  to  keep  the  body  from  losing 
electric  charge,  we  must  have 


KE, 


h/-W  =  V, 


where  q^  is  the  magnitude  of  the  electronic  charge  .  .  . 

If  the  derived  formula  is  correct,  then  V,,op,  when 
plotted  as  a  function  of  the  frequency  of  the  incident  light, 
should  yield  a  straight  line  whose  slope  should  be  inde- 
pendent of  the  nature  of  the  substance  illuminated. 

We  can  now  compare  Einstein's  photoelectric  equation  with  the 
experimental  results  to  test  whether  or  not  his  theory  accounts  for 
the  results.  According  to  the  equation,  the  kinetic  energy  is  greater 
than  zero  only  when  hf  is  greater  than  W.  Hence,  the  equation  says 
that  an  electron  can  be  emitted  only  when  the  frequency  of  the 
incident  light  is  greater  than  a  certain  lowest  value/,,  (where 
hf„  =  W.) 

Next,  according  to  Einstein's  photon  model,  it  is  an  individual 
photon  that  ejects  an  electron.  The  intensity  of  the  light  is  propor- 
tional to  the  number  of  the  photons  in  the  light  beam,  and  the 
number  of  photoelectrons  ejected  is  proportional  to  the  number  of 
photons  incident  on  the  surface.  Hence  the  number  of  electrons 
ejected  (and  with  it  the  photoelectric  current)  is  proportional  to 
the  intensity  of  the  incident  light. 

According  to  Einstein's  model  the  light  energy  is  concentrated 
in  the  quanta  (photons).  So,  no  time  is  needed  for  collecting  light 

Student  apparatus  for  photoelectric 
experiments  often  includes  a  vacuum 
phototube,  like  the  one  shown  at  the 
left.  The  collecting  wire  corresponds 
to  A  in  Fig.  (a)  on  p.  41.  and  is  at  the 
center  of  a  cylindrical  photosensitive 
surface  that  corresponds  to  C.  The 
frequency  of  the  light  entering  the 
tube  is  selected  by  placing  colored 
filters  between  the  tube  and  a  white 
light  source,  as  shown  at  the  right. 


detector 


Albert  Einstein  (1879-1955)  was  born  in  the  city  of 
Dim,  in  Germany.  Like  Newton  he  showed  no  particu- 
lar intellectual  promise  as  a  youngster.  He  received 
his  early  education  in  Germany,  but  at  the  age  of  17, 
dissatisfied  with  the  regimentation  in  school  and 
militarism  in  the  nation,  he  left  for  Switzerland.  After 
graduation  from  the  Polytechnic  School,  Einstein  (in 
1901)  found  work  in  the  Swiss  Patent  Office  in  Berne. 
This  job  gave  Einstein  a  salary  to  live  on  and  an  op- 
portunity to  use  his  spare  time  for  working  in  physics 
on  his  own.  In  1905  he  published  three  papers  of 
epoch-making  importance.  One  dealt  with  quantum 
theory  and  included  his  theory  of  the  photoelectric 
effect.  Another  treated  the  problem  of  molecular  mo- 
tions and  sizes,  and  worked  out  a  mathematical  anal- 
ysis of  the  phenomenon  of  "Brownian  motion." 
Einstein's  analysis  and  experimental  work  by  Jean 
Perrin,  a  French  physicist,  provided  a  strong  argu- 
ment for  the  molecular  motions  assumed  in  the  kinetic 
theory.  Einstein's  third  1905  paper  provided  the  theory 
of  special  relativity  which  revolutionized  modern 
thought  about  the  nature  of  space,  time,  and  physical 
theory. 


In  1915,  Einstein  published  a  paper  on  the  theory 
of  general  relativity  in  which  he  provided  a  new  theory 
of  gravitation  that  included  Newton's  theory  as  a 
special  case. 

When  Hitler  and  the  Nazis  came  to  power  in  Ger- 
many, in  1933,  Einstein  came  to  the  United  States  and 
became  a  member  of  the  Institute  for  Advanced  Stu- 
dies at  Princeton.  He  spent  the  rest  of  his  working 
life  seeking  a  unified  theory  which  would  include 
gravitation  and  electromagnetics.  Near  the  beginning 
of  World  War  II,  Einstein  wrote  a  letter  to  President 
Roosevelt,  warning  of  the  war  potential  of  an  "atomic 
bomb,"  for  which  the  Germans  had  all  necessary 
knowledge  and  motivation  to  work.  After  World  War 
II,  Einstein  devoted  much  of  his  time  to  promoting 
world  agreement  to  end  the  threat  of  atomic  warfare. 


46 


Electrons  and  Quanta 


(3)  Immediate  emission.  Reason: 

a  single  photon  provides  the  energy 
concentrated  in  one  place. 

(4)  KE,„„.r  increases  linearly  with 
frequency  above  f„.  Reason:  the 
work  needed  to  remove  the  electron 
is  IV  =  hf„;  any  energy  left  over 
from  the  original  photon  is  now 
available  for  kinetic  energy  of  the 
electron. 


The  equation  K£„,„,  -^  hf  -  IV  can 
be  said  to  have  led  to  two  Nobel 
prizes:  one  to  Einstein,  who  derived 
it  theoretically,  and  one  to  Millikan, 
who  verified  it  experimentally.  This 
equation  is  the  subject  of  a  Project 
Physics  laboratory  experiment. 

SG  18.11 


energy;  the  quanta  transfer  their  energy  immediately  to  the 
photoelectrons,  which  emerge  after  the  very  short  time  required  for 
them  to  escape  from  the  surface. 

Finally,  the  photoelectric  equation  predicts  that  the  greater 
the  frequency  of  the  incident  light,  the  greater  is  the  maximum 
kinetic  energy  of  the  ejected  electrons.  According  to  the  photon 
model,  the  photon's  energy  is  directly  proportional  to  the  hght 
frequency.  The  minimum  energy  needed  to  eject  an  electron  is  the 
energy  required  for  the  electron  to  escape  from  the  metal  surface  — 
which  explains  why  light  of  frequency  less  than  some  frequency 
fg  cannot  eject  any  electrons.  The  kinetic  energy  of  the  escaping 
electron  is  the  difference  between  the  energy  of  the  absorbed  photon 
and  the  energy  lost  by  the  electron  in  passing  through  the  surface. 

Thus,  Einstein's  photoelectric  equation  agreed  quahtatively  with 
the  experimental  results.  There  remained  two  quantitative  tests  to 
be  made:  (1)  does  the  maximum  energy  vary  in  direct  proportion  to 
the  light  frequency?  (2)  is  the  proportionality  factor  h  really  the 
same  for  all  substances?  For  some  10  years,  experimental  physicists 
attempted  these  quantitative  tests.  One  of  the  experimental 
difficulties  was  that  the  value  of  W  for  a  metal  is  greatly  changed 
if  there  are  impurities  (for  example,  a  layer  of  oxidized  metal)  on 
the  surface.  It  was  not  until  1916  that  it  was  estabhshed.  by  Robert  A. 
Milhkan,  that  there  is  indeed  a  straight-line  relationship  between 
the  frequency  of  the  absorbed  light  and  the  maximum  kinetic 
energy  of  the  photoelectrons  (as  in  the  graph  on  p.  42).  To  obtain 
his  data  Millikan  designed  an  apparatus  in  which  the  metal  photo- 
electric surface  was  cut  clean  while  in  a  vacuum.  A  knife  inside  the 
evacuated  volume  was  manipulated  by  an  electromagnet  outside 
the  vacuum  to  make  the  cuts.  This  rather  intricate  arrangement 
was  required  to  achieve  an  uncontaminated  metal  surface. 

Millikan  also  showed  that  the  straight  line  graphs  obtained  for 
different  metals  all  had  the  same  slope,  even  though  the  threshold 
frequencies  were  different.  The  value  of  h  could  be  obtained  from 
Milhkan's  measurements;  it  was  the  same  for  each  metal  surface, 
and,  it  agreed  very  well  with  a  value  obtained  by  means  of  other, 
independent  methods.  So  Einstein's  theory  of  the  photoelectric 
effect  was  verified  quantitatively. 

Historically,  the  first  suggestion  that  the  energy  in  electro- 
magnetic radiation  is  "quantized"  (comes  in  definite  quanta)  came 
not  from  the  photoelectric  effect,  but  from  studies  of  the  heat  and 
light  radiated  by  hot  solids.  The  concept  of  quanta  of  energy  was 
introduced  by  Max  Planck,  a  German  physicist,  in  1900.  five  years 
before  Einstein's  theory,  and  the  constant  h  is  known  as  Planck's 
constant.  Planck  was  trying  to  account  for  the  way  heat  (and  light) 
energy  radiated  by  a  hot  body  is  related  to  the  frequency  of  the 
radiation.  Classical  physics  (nineteenth-century  thermodynamics 
and  electromagnetism)  could  not  account  for  the  experimental 
facts.  Planck  found  that  the  facts  could  be  interpreted  only  by 
assuming  that  atoms,  on  radiating,  change  their  energy  discontin- 
uously,  in  quantized  amounts.  Einstein's  theory  of  the  photoelectric 
effect  was  actually  an  extension  and  application  of  Planck's  quan- 


Section  18.5 


47 


Robert  Andrews  Millikan  (1868-1953), 
an  American  physicist,  attended  Ober- 
lin  College,  where  his  interest  in  phys- 
ics was  only  mild.  After  his  graduation 
he  became  more  interested  in  physics, 
taught  at  Oberlin  while  taking  his 
master's  degree,  and  then  obtained 
his  doctor's  degree  from  Columbia 
University  in  1895.  After  post-doctoral 
work  in  Germany  he  went  to  the  Uni- 
versity of  Chicago,  where  he  became  a 
professor  of  physics  in  1910.  His  work 
on  the  determination  of  the  electronic 
charge  took  place  from  1906  to  1913. 
He  was  awarded  the  Nobel  Prize  in 
physics  in  1923  for  this  research,  and 
for  the  very  careful  experiments  which 
resulted  in  the  verification  of  the  Ein- 
stein photoelectric  equation  (Sec. 
18.4).  In  1921,  Millikan  moved  to  the 
California  Institute  of  Technology, 
eventually  to   become   its   president. 


turn  theory  of  thermal  radiation:  Einstein  postulated  that  the 
quantum  change  in  the  atom's  energy  is  carried  off  as  a  localized 
photon  rather  than  being  spread  continuously  over  the  light  wave. 

The  experiments  and  the  theory  on  radiation  are  much  more 
difficult  to  describe  than  the  experiments  and  the  theory  of  the 
photoelectric  effect.  That  is  why  we  have  chosen  to  introduce  the 
new  concept  of  quanta  of  energy  by  means  of  the  photoelectric 
effect.  By  now,  there  have  been  many  ways  of  checking  both 
Planck's  and  Einstein's  conceptions.  In  all  these  cases,  Planck's 
constant  h  has  now  the  same  basic  position  in  quantum  physics 
that  Newton's  universal  constant  G  has  in  the  physics  of 
gravitation. 

The  photoelectric  effect  presented  physicists  with  a  real 
dilemma.  According  to  the  classical  wave  theory,  light  consists  of 
electromagnetic  waves  extending  continuously  throughout  space. 
This  theory  was  highly  successful  in  explaining  optical  phenomena 
(reflection,  refraction,  polarization,  interference),  but  could  not 
account  for  the  photoelectric  effect.  Einstein's  theory,  in  which 
the  existence  of  separate  lumps  of  light  energy  was  postulated, 
accounted  for  the  photoelectric  effect;  it  could  not  account  for  the 
other  properties  of  hght.  The  result  was  that  there  were  two  models 
whose  basic  concepts  seemed  to  be  mutually  contradictory.  Each 
model  had  its  successes  and  failures.  The  problem  was:  what,  if 
anything,  could  be  done  about  the  contradictions  between  the  two 
models?  We  shall  see  later  that  the  problem  and  its  treatment  have 
a  central  position  in  modern  physics. 

Q8     Einstein's  idea  of  a  quantum  of  light  had  a  definite  relation 
to  the  wave  model  of  light.  What  was  it? 

Q9    Why  does  the  photoelectron  not  have  as  much  energy  as 
the  quantum  of  light  which  causes  it  to  be  ejected? 


Max  Planck  (1858-1947),  a  German 
physicist,  was  the  originator  of  the 
quantum  theory,  one  of  the  two  great 
revolutionary  physical  theories  of  the 
20th  century.  (The  other  is  Einstein's 
relativity  theory.)  Planck  won  the 
Nobel  Prize  in  1918  for  his  quantum 
theory.  He  tried  for  many  years  to 
show  that  this  theory  can  be  under- 
stood in  terms  of  the  classical  physics 
of  Newton  and  Maxwell,  but  this 
attempt  did  not  succeed.  Quantum 
physics  is  fundamentally  different, 
through  its  postulate  that  energy  in 
light  and  matter  is  not  continuously 
divisible  into  any  arbitrarily  small 
quantity,  but  exists  in  quanta  of  defi- 
nite amount. 


48 


Electrons  and  Quanta 


Wilhelm  Konrad  Rontgen  (1845-1923) 


The  discovery  of  x  rays  was  nar- 
rowly missed  by  several  physicists, 
including  Hertz  and  Lenard  (another 
well-known  German  physicist).  An 
English  physicist,  Frederick  Smith, 
found  that  photographic  plates 
kept  in  a  box  near  a  cathode-ray 
tube  were  liable  to  be  fogged  — so 
he  told  his  assistant  to  keep  them 
in  another  place! 


Q10    What  does  a  "stopping  voltage"  of,  say.  2.0  volts  indicate 
about  the  photoelectrons  emerging  from  a  metal  surface? 


18.6     X  rays 

In  1895.  a  surprising  discovery  was  made  which,  hke  the 
photoelectric  effect,  did  not  fit  in  with  accepted  ideas  about  electro- 
magnetic waves  and  eventually  needed  quanta  for  its  explanation. 
The  discovery  was  that  of  x  rays  by  the  German  physicist.  Wilhelm 
Rontgen;  its  consequences  for  atomic  physics  and  technology  are 
dramatic  and  important. 

On  November  8.  1895.  Rontgen  was  experimenting  with  the 
newly  found  cathode  rays,  as  were  many  physicists  all  over  the 
world.  According  to  a  biographer. 

...  he  had  covered  the  all-glass  pear-shaped  tube  [Crookes 
tube  — see  Sec.  18.2]  with  pieces  of  black  cardboard,  and 
had  darkened  the  room  in  order  to  test  the  opacity  of  the 
black  paper  cover.  Suddenly,  about  a  yard  from  the  tube, 
he  saw  a  weak  light  that  shimmered  on  a  little  bench  he 
knew  was  nearby.  Highly  excited,  Rontgen  Ut  a  match 
and,  to  his  great  surprise,  discovered  that  the  source  of 
the  mysterious  light  was  a  httle  barium  platinocyanide 
screen  lying  on  the  bench. 

Barium  platinocyanide,  a  mineral,  is  one  of  the  many  chemicals 
known  to  fluoresce,  that  is,  to  emit  visible  light  when  illuminated 
with  ultraviolet  hght.  But  no  source  of  ultraviolet  hght  was  present 
in  Rontgen's  experiment.  Cathode  rays  had  not  been  observed  to 
travel  more  than  a  few  centimeters  in  air.  So,  neither  ultraviolet 
light  nor  the  cathode  rays  themselves  could  have  caused  the 
fluorescence.  Rontgen  therefore  deduced  that  the  fluorescence  he 
had  observed  was  due  to  rays  of  a  new  kind,  which  he  named 
X  rays,  that  is,  rays  of  an  unknown  nature.  During  the  next  seven 
weeks  he  made  a  series  of  experiments  to  determine  the  properties 
of  this  new  radiation.  He  reported  his  results  on  December  28.  1895 
to  a  scientific  society  in  a  paper  whose  title  (translated)  is  "On  a 
New  Kind  of  Rays." 

Rontgen's  paper  described  nearly  all  of  the  properties  of  x  rays 
that  are  known  even  now.  It  included  an  account  of  the  method  of 
producing  the  rays,  and  proof  that  they  originated  in  the  glass  wall 
of  the  tube,  where  the  cathode  rays  struck  it.  Rontgen  showed  that 
the  X  rays  travel  in  straight  lines  from  their  place  of  origin  and 
that  they  darken  a  photographic  plate.  He  reported  in  detail  the 
ability  of  x  rays  to  penetrate  various  substances  — paper,  wood, 
aluminum,  platinum  and  lead.  Their  penetrating  power  was  greater 
through  light  materials  (paper,  wood,  flesh)  than  through  dense 
materials  (platinum,  lead,  bone).  He  described  photographs  showing 
"the  shadows  of  bones  of  the  hand,  of  a  set  of  weights  inside  a 
small  box,  and  of  a  piece  of  metal  whose  inhomogeneity  becomes 
apparent  with  x  rays."  He  gave  a  clear  description  of  the  shadows 

Opposite:  One  of  the  earliest  x-ray  photographs  made  in  the  United 
States  (1896).  The  man  x-rayed  had  been  hit  by  a  shotgun  blast. 


/ 


^ 


50 


Electrons  and  Quanta 


X  rays  were  often  referred  to  as 
Rontgen  rays,  after  their  discoverer. 


It  is  easy  to  see  why  a  charged 
electroscope  will  be  discharged 
when  the  air  around  it  is  ionized: 
It  attracts  the  ions  of  the  opposite 
charge  from  the  air. 


Such  a  particle -the  neutron— was 
discovered  in  1932.  You  will  see  in 
Chapter  23  (Unit  6)  how  hard  it  was 
to  identify.  But  the  neutron  has 
nothing  to  do  with  x  rays. 


SG  18.12 


cast  by  the  bones  of  the  hand  on  the  fluorescent  screen.  Rontgen 
also  reported  that  the  x  rays  were  not  deflected  by  a  magnetic  field, 
and  showed  no  reflection,  refraction  or  interference  effects  in 
ordinary  optical  apparatus. 

One  of  the  most  important  properties  of  x  rays  was  discovered 
by  J.  J.  Thomson  a  month  or  two  after  the  rays  themselves  had 
become  known.  He  found  that  when  the  rays  pass  through  a  gas 
they  make  it  a  conductor  of  electricity.  He  attributed  this  effect  to 
"a  kind  of  electrolysis,  the  molecule  being  spHt  up,  or  nearly  spHt 
up  by  the  Rontgen  rays."  The  x  rays,  in  passing  through  the  gas. 
knock  electrons  loose  from  some  of  the  atoms  or  molecules  of  the 
gas.  The  atoms  or  molecules  that  lose  these  electrons  become 
positively  charged.  They  are  called  ions  because  they  resemble  the 
positive  ions  in  electrolysis,  and  the  gas  is  said  to  be  ionized.  The 
freed  electrons  may  also  attach  themselves  to  previously  neutral 
atoms  or  molecules,  thereby  leaving  them  negatively  charged. 

Rontgen  and  Thomson  found,  independently,  that  electrified 
bodies  are  discharged  when  the  air  around  them  is  ionized  by 
X  rays.  The  rate  of  discharge  was  shown  to  depend  on  the  intensity 
of  the  rays.  This  property  was  therefore  used  as  a  convenient 
quantitative  means  of  measuring  the  intensity  of  an  x-ray  beam. 
As  a  result,  careful  quantitative  measurements  of  the  properties 
and  effects  of  x  rays  could  be  made. 

One  of  the  problems  that  aroused  keen  interest  during  the  years 
following  the  discovery  of  x  rays  was  that  of  the  nature  of  the 
mysterious  rays.  They  did  not  act  like  charged  particles  — electrons 
for  example  — because  they  were  not  deflected  by  magnetic  or 
electric  fields.  Therefore  it  seemed  that  they  had  to  be  either  neutral 
particles  or  electromagnetic  waves.  It  was  difficult  to  choose 
between  these  two  possibilities.  On  the  one  hand,  no  neutral 
particles  of  atomic  size  (or  smaller)  were  then  known  which  had 
the  penetrating  power  of  x  rays.  The  existence  of  neutral  particles 
with  high  penetrating  power  would  be  extremely  hard  to  prove  in 
any  case,  because  there  was  no  way  of  getting  at  them.  On  the 
other  hand,  if  the  x  rays  were  electromagnetic  waves,  they  would 
have  to  have  extremely  short  wavelengths:  only  in  this  case, 
according  to  theory,  could  they  have  high  penetrating  power  and 
show  no  refraction  or  interference  effects  with  ordinary  optical 
apparatus. 

As  we  have  already  discussed  in  Chapters  12  and  13,  distinctly 
wavelike  properties  become  apparent  only  when  waves  interact 
with  objects  (like  slits  in  a  barrier)  that  are  smaller  than  several 
wavelengths  across.  The  wavelength  hypothesized  for  x  rays  would 
be  on  the  order  of  10~'"  meter.  So  to  demonstrate  their  wave 
behavior,  it  would  be  necessary  to  see,  say,  a  diff'raction  grating 
with  slits  spaced  about  10"'"  meter  apart.  Several  lines  of  evidence, 
from  kinetic  theory  and  from  chemistry,  indicated  that  atoms  were 
about  10~'°  meter  in  diameter.  It  was  suggested,  therefore,  that 
X  rays  might  be  diff'racted  noticeably  by  crystals,  in  which  the 
atoms  are  arranged  in  orderly  layers  about  10~'°  meter  apart.  In 
1912,  such  experiments  succeeded;  the  layers  of  atoms  do  act  like 


Section  18.6 


51 


X-ray  diffraction  patterns  from  a  metal 
crystal.  The  black  spots  are  produced 
by  constructive  interference  of  x  rays. 


diffraction  gratings,  and  x  rays  do,  indeed,  act  like  electromagnetic 
radiations  of  very  short  wavelength  — like  ultra  ultraviolet  light. 
These  experiments  are  more  complicated  to  interpret  than  diffraction 
of  a  beam  of  light  by  a  single,  two-dimensional  optical  grating.  Now 
the  diffraction  effect  occurs  in  three  dimensions  instead  of  two. 
Hence  the  diffraction  patterns  are  far  more  elaborate  (see  the 
illustration  above). 

In  addition  to  wave  properties,  x  rays  were  also  found  to  have 
quantum  properties:  they  can,  for  example,  cause  the  emission  of 
electrons  from  metals.  These  electrons  have  greater  kinetic  energies 
than  those  produced  by  ultraviolet  hght.  (The  ionization  of  gases  by 
X  rays  is  also  an  example  of  the  photoelectric  effect;  in  this  case 
the  electrons  are  freed  from  the  atoms  and  molecules  of  the  gas.) 
Thus,  X  rays  also  require  quantum  theory  for  the  explanation  of 
some  of  their  behavior.  So,  like  Hght,  x  rays  were  shown  to  have 
both  wave  and  particle  properties. 

Rontgen's  initial  discovery  of  x  rays  excited  intense  interest 
throughout  the  entire  scientific  world.  His  experiments  were 
immediately  repeated  — and  extended  in  many  laboratories  in  both 
Europe  and  America.  The  scientific  journals  during  the  year  1896 
were  filled  with  letters  and  articles  describing  new  experiments  or 
confirming  the  results  of  earUer  experiments.  (This  widespread 
experimentation  was  made  possible  by  the  fact  that,  during  the 
years  before  Rontgen's  discovery,  the  passage  of  electricity  through 
gases  had  been  a  popular  topic  for  study  by  physicists  — many 
physics  laboratories  had  cathode-ray  tubes,  and  could  produce 
X  rays  easily.) 

Intense  interest  in  x  rays  was  generated  by  the  spectacular  use 
of  these  rays  in  medicine.  Within  three  months  of  Rontgen's 


SG  18.13 


SG  18.14-18.16 


Originally,  x  rays  were  produced  in  Rbntgen's 
laboratory  when  cathode  rays  (electrons)  struck 
a  target  (the  glass  wall  of  the  tube).  Nowadays 
X  rays  are  commonly  produced  by  directing  a  beam 
of  high  energy  electrons  onto  a  metal  target.  As 
the  electrons  are  deflected  and  stopped,  x  rays  of 
various  energies  are  produced.  The  maximum 
energy  a  single  ray  can  have  is  the  total  kinetic 
energy  the  incident  electron  is  giving  up  on  being 
stopped.  So  the  greater  the  voltage  across  which 
the  electron  beam  is  accelerated,  the  more  ener- 
getic-and  penetrating -are  the  x  rays.  One  type 
of  X  ray  tube  is  shown  in  the  sketch  below,  where 
a  stream  of  electrons  is  emitted  from  a  cathode  C 
and  accelerated  to  a  tungsten  target  T  by  a  strong 
electric  field  (high  potential  difference). 


In  the  photograph  at  the  right  is  the  inner  part  of 
a  high  voltage  generator  which  can  be  used  to 
provide  the  large  potential  differences  required 
for  making  energetic  x  rays.  This  Van  de  Graaf 
type  generator  (named  after  the  American  physi- 
cist who  invented  it),  although  not  very  different 
in  principle  from  the  electrostatic  generators  of 
the  18th  century,  can  produce  an  electric  potential 
difference  of  4,000,000  volts  between  the  top  and 
ground. 


Above  left  is  a  rose,  photographed 
with  X  rays  produced  when  the  po- 
tential difference  between  the  elec- 
tron-emitting cathode  and  the  target 
in  the  x-ray  tube  is  30,000  volts. 

Below  the  rose  is  the  head  of  a 
dogfish  shark;  its  blood  vessels  have 
been  injected  with  a  fluid  that  absorbs 
X  rays  in  order  to  study  the  blood 
vessels. 

In  the  photograph  at  the  bottom  of 
the  page,  x  rays  are  being  used  to 
inspect  the  welds  of  a  400-ton  tank 
for  a  nuclear  reactor. 

Immediately  above  is  illustrated  the 
familiar  use  of  x  rays  in  dentistry  and 
the  resulting  records.  Because  x  rays 
are  injurious  to  tissues,  a  great  deal 
of  caution  is  required  in  using  them. 
For  example,  the  shortest  possible 
pulse  of  X  rays  is  used,  lead  shielding 
is  provided  for  the  body,  and  the  tech- 
nician stands  behind  a  wall  of  lead  and 
lead  glass. 


54  Electrons  and  Quanta 

discovery,  x  rays  were  being  put  to  practical  use  in  a  hospital  in 
Vienna  in  connection  with  surgical  operations.  The  use  of  this  new 
aid  to  surgery  spread  rapidly.  Since  Rontgen's  time,  x  rays  have 
revolutionized  some  phases  of  medical  practice,  especially  the 
diagnosis  of  some  diseases,  and  the  treatment  of  some  forms  of 
cancer.  In  other  fields  of  applied  science,  both  physical  and 
biological,  uses  have  been  found  for  x  rays  which  are  nearly  as 
important  as  their  use  in  medicine.  Among  these  are  the  study  of 
the  crystal  structure  of  materials;  "industrial  diagnosis,"  such  as 
the  search  for  possible  defects  in  materials  and  engineering 
structures;  the  study  of  old  paintings  and  sculptures;  and  many 
others. 

Q11     X  rays  were  the  first  "ionizing"  radiation  discovered. 
What  does  "ionizing"  mean? 

Q12     What  were  three  properties  of  x  rays  that  led  to  the 
conclusion  that  x  rays  were  electromagnetic  waves? 

Q13     What  was  the  evidence  that  x  rays  had  a  very  short 
wavelength? 

18.7    Electrons,  quanta  and  the  atom 

By  the  beginning  of  the  twentieth  century  enough  chemical 
and  physical  information  was  available  so  that  many  physicists 
devised  models  of  atoms.  It  was  known  that  negative  particles 
with  identical  properties  — electrons  could  be  obtained  from  many 
different  substances  and  in  different  ways.  This  suggested  the 
notion  that  electrons  are  constituents  of  all  atoms.  But  electrons 
are  negatively  charged,  while  samples  of  an  element  are  ordinarily 
electrically  neutral  and  the  atoms  making  up  such  samples  are 
also  presumably  neutral.  Hence  the  presence  of  negative  electrons 
in  an  atom  would  seem  to  require  the  presence  also  of  an  equal 
amount  of  positive  charge. 

Comparison  of  the  values  of  qlm  for  the  electron  and  for 
charged  hydrogen  atoms  indicated,  as  mentioned  in  Sec.  18.2,  that 
hydrogen  atoms  are  nearly  two  thousand  times  more  massive  than 
electrons.  Experiments  (which  will  be  discussed  in  some  detail  in 
Chapter  22)  showed  that  electrons  constitute  only  a  very  small  part 
of  the  atomic  mass  in  any  atom.  Consequently  any  model  of  an 
atom  must  take  into  account  the  following  information:  (a)  an 
electrically  neutral  atom  contains  equal  amounts  of  positive  and 
negative  charge;  (b)  the  negative  charge  is  associated  with  only  a 
small  part  of  the  mass  of  the  atom.  Accordingly,  any  atomic  model 
should  answer  at  least  two  questions:  (1)  how  many  electrons  are 
there  in  an  atom,  and  (2)  how  are  the  electrons  and  the  positive 
charge  arranged  in  an  atom? 

During  the  first  ten  years  of  the  twentieth  century,  several 
atomic  models  were  proposed,  but  none  was  satisfactory.  The 
early  models  were  all  based  entirely  upon  classical  physics,  that  is, 
upon  the  physics  of  Newton  and  Maxwell.  No  one  knew  how  to 
invent  a  model  that  also  took  account  of  the  theory  of  Planck  which 


Section  18.7 


55 


incorporated  the  quantization  of  energy.  There  was  also  need  for 
more  detailed  experimental  facts  — for  example,  this  was  the  period 
during  which  the  charge  on  the  electron  and  the  main  facts  of 
photoelectricity  were  still  being  found.  Nevertheless  physicists 
cannot  and  should  not  wait  until  every  last  fact  is  in  — that  will 
never  happen,  and  you  can't  even  know  what  the  missing  facts  are 
unless  you  have  some  sort  of  model.  Even  an  incomplete  or  a  partly 
wrong  model  will  provide  clues  on  which  to  build  a  better  one. 

Until  1911  the  most  popular  model  for  the  atom  was  one 
proposed  by  J.  J.  Thomson  in  1904.  Thomson  suggested  that  an 
atom  consisted  of  a  sphere  of  positive  electricity  in  which  was 
distributed  an  equal  amount  of  negative  charge  in  the  form  of 
electrons.  Under  this  assumption,  the  atom  was  like  a  pudding 
of  positive  electricity  with  the  negative  electricity  scattered  in  it 
like  raisins.  The  positive  "fluid"  was  assumed  to  act  on  the  negative 
charges,  holding  them  in  the  atom  by  means  of  electric  forces  only. 
Thomson  did  not  specify  how  the  positive  "fluid"  was  held  together. 
The  radius  of  the  atom  was  taken  to  be  of  the  order  of  10"'°  m,  on 
the  basis  of  information  from  the  kinetic  theory  of  gases  and  other 
considerations  (see  SG  18.13).  With  this  model  Thomson  was  able 
to  calculate  that  certain  arrangements  of  electrons  would  be  stable, 
the  first  requirements  for  explaining  the  existence  of  stable  atoms. 
Thomson's  theory  also  suggested  that  chemical  properties  might  be 
associated  with  particular  groupings  of  electrons.  A  systematic 
repetition  of  chemical  properties  might  then  occur  among  groups 
of  elements.  But  it  was  not  possible  to  deduce  the  detailed  structure 
of  the  atoms  of  particular  elements,  and  no  detailed  comparison 
with  the  actual  periodic  table  could  be  made. 


I 


£»/  Z'2.  Z-3  l-A 

In  Chapter  19  we  shall  discuss  some  additional  experimental 
information  that  provided  valuable  clues  to  improved  models  of  the 
structure  of  atoms.  We  shall  also  see  how  one  of  the  greatest 
physicists  of  our  time,  Niels  Bohr,  was  able  to  combine  the  experi- 
mental evidence  then  available  with  the  new  concept  of  quanta 
into  a  successful  theory  of  atomic  structure.  Although  Bohr's  model 
was  eventually  replaced  by  more  sophisticated  ones,  it  provided  the 
clues  that  led  to  the  presently  accepted  theory  of  the  atom,  and  to 
this  day  is  in  fact  quite  adequate  for  explaining  most  of  the  main 
facts  with  which  we  shall  be  concerned  in  this  course. 

Q14    Why  was  most  of  the  mass  of  an  atom  beheved  to  be 
associated  with  positive  electric  charge? 

Q15    Why  don't  physicists  wait  until  "all  the  facts  are  in"  before 
they  begin  to  theorize  or  make  models? 


See  the  Project  Physics  film  loop 
Thomson  Model  of  the  Atom. 


2-S- 


Z^4, 


Some  stable  (hypothetical)  arrange- 
ments of  electrons  in  Thomson  atoms. 
The  atomic  number  Z  is  interpreted 
as  equal  to  the  number  of  electrons. 


STUDY  GUIDE 


18.1  The  Project  Physics  learning  materials 
particularly  appropriate  for  Chapter  18  include 
the  following: 

Experiments 

The  charge-to-mass  ratio  for  an  electron 
The  measurement  of  elementary  charge 
The  photoelectric  effect 

Activities 

Writings  by  and  about  Einstein 

Measuring  qlm  for  the  electron 

Cathode  rays  in  a  Crookes  tube 

X  rays  from  a  Crookes  tube 

Lighting  a  bulb  photoelectrically  with  a 

match 

Reader  Articles 

Failure  and  Success 
Einstein 

Transparencies 

Photoeler  trie  experiment 
Photoelectric  equation 

18.2  In  Thomson's  experiment  on  the  ratio  of 
charge  to  mass  of  cathode  ray  particles  (p.  36), 
the  following  might  have  been  typical  values  for 
B,  V  and  d:  with  a  magnetic  field  B  alone,  the 
deflection  of  the  beam  indicated  a  radius  of 
curvature  of  the  beam  within  the  field  of  0.114 
meters  for  B  =  1.0  x  10"'  tesla.*  With  the  same 
magnetic  field,  the  addition  of  an  electric  field  in 
the  same  region  (V  =  200  volts,  plate  separation 
d  =  0.01  meter)  made  the  beam  go  on  straight 
through. 

(a)  Find  the  speed  of  the  cathode  ray  particles 
in  the  beam. 

(b)  Find  qlm  for  the  cathode  ray  particles. 

18.3  Given  the  value  for  the  charge  on  the 
electron,  show  that  a  current  of  one  ampere  is 
equivalent  to  the  movement  of  6.25  x  10"* 
electrons  per  second  past  a  given  point. 

18.4  In  the  apparatus  of  Fig.  18.7,  an  electron  is 
turned  back  before  reaching  plate  A  and 
eventually  arrives  at  electrode  C  from  which  it 
was  ejected.  It  arrives  with  some  kinetic  energy. 
How  does  this  final  energy  of  the  electron  compare 
with  the  energy  it  had  as  it  left  the  electrode  C? 

18.5  It  is  found  that  at  light  frequencies  below 
the  threshold  frequency  no  photoelectrons  are 
emitted.  What  happens  to  light  energy? 

18.6  For  most  metals,  the  work  function  W  is 
about  10"'"  joules.  Light  of  what  frequency  will 
cause  photoelectrons  to  leave  the  metal  with 
virtually  no  kinetic  energy?  In  what  region  of 
the  spectrum  is  this  frequency? 

18.7  What  is  the  energy  of  a  Ught  photon  which 

*The  MKSA  unit  lor  B  is  N/ampm  and  is  now 
called  the  tesla.  (after  the  electrical  engineer 
Nikola  Tesla). 


corresponds  to  a  wavelength  of  5  x  10  '  m? 
5  X  10""  m? 

18.8  The  minimum  or  threshold  frequency  of 
light  from  emission  of  photoelectrons  for  copper 
is  1.1  X  10'^  cycles/sec.  When  ultraviolet  Ught  of 
frequency  1.5  x  10'-^  cycles/sec  shines  on  a  copper 
surface,  what  is  the  maximum  energy  of  the 
photoelectrons  emitted,  in  joules?  In  electron 
volts? 

18.9  What  is  the  lowest-frequency  bght  that  will 
cause  the  emission  of  photoelectrons  from  a 
surface  whose  work  function  is  2.0  eV  (that  is, 
an  energy  of  at  least  2.0  eV  is  needed  to  eject  an 
electron)? 

18.10  Monochromatic  light  of  wavelength  5000 
A  falls  on  a  metal  cathode  to  produce  photo- 
electrons. (lA  =  10"'"  meter)  The  Ught  intensity 
at  the  surface  of  the  metal  is  10- joules/m^ 

per  sec. 

(a)  What  is  the  frequency  of  the  Ught? 

(b)  What  is  the  energy  (in  joules)  of  a  single 
proton  of  the  light? 

(c)  How  many  photons  fall  on  1  m-  in  one  sec? 

(d)  If  the  diameter  of  an  atom  is  about  1  A. 
how  many  photons  fall  on  one  atom  in  one 
second,  on  the  average? 

(e)  How  often  would  one  photon  fall  on  one 
atom,  on  the  average? 

(f )  How  many  photons  fall  on  one  atom  in 
10"'"  sec,  on  the  average? 

(g)  Suppose  the  cathode  is  a  square  0.05  m  on 
a  side.  How  many  electrons  are  released 
per  second,  assuming  every  photon  releases 
a  photoelectron?  How  big  a  current  would 
this  be  in  amperes? 

18.11  Roughly  how  many  photons  of  visible  Ught 
are  given  off  per  second  by  a  1-watt  flashlight? 
(Only  a  bout  5  percent  of  the  electric  energy  input 
to  a  tungsten-filament  bulb  is  given  off"  as  \  isible 
Ught.) 

Hint:  first  find  the  energy,  in  joules,  of  an  average 
photon  of  visible  Ught. 

18.12  Recall  from  Sec.  17.8  that  96.540  coulombs 
of  charge  will  deposit  31.77  grams  of  copper  in 
the  electrolysis  of  copper  sulfate.  In  Sec.  18.3.  the 
charge  of  a  single  electron  was  reported  to  be  1.6 
X  10"'-' coulomb. 

(a)  How  many  electrons  must  be  transferred 
to  deposit  31.77  grams  of  copper? 

(b)  The  density  of  copper  is  8.92  grams  per 
cm'.  How  many  copper  atoms  would 
there  be  in  the  1  cm^?  (Actually  copper 
has  a  coiTibining  number  of  2.  which 
suggests  that  2  electrons  are  required  to 
deposit  a  single  copper  atom.) 

(c)  What  is  the  approximate  volume  of  each 
copper  atom? 

(d)  What  is  the  approximate  diameter  of  a 
copper  atom?  (For  this  rough  approxima- 
tion, assume  that  the  atoms  are  cubes.) 


56 


18.13     The  approximate  size  of  atoms  can  be 
calculated  in  a  simple  way  from  x-ray  scattering 
experiments.  The  diagram  below  represents  the 
paths  of  two  portions  of  an  x-ray  wavefront,  part 
of  which  is  scattered  from  the  first  layer  of  atoms 
in  a  CFN'stal,  and  part  of  which  is  scattered  from 
the  second  layer.  The  part  reflected  from  the 
second  layer  travels  a  distance  2x  further  before 
it  emerges  from  the  crystal. 


IqOO  a^<i^  O  O 

(a)  Under  what  conditions  will  the  scattered 
wavefronts  reinforce  one  another  (that 
is,  be  in  phase)? 

(b)  Under  the  conditions,  will  the  scattered 
wavefronts  cancel  one  another? 

(c)  Use  trigonometr^'  to  express  the  relation- 
ship among  wavelength  K  the  distance  d 
between  layers,  and  the  angle  of  reflection 
6„„j.  that  will  have  maximum  intensity. 

18.14     The  highest  frequency, /^^j,  of  the  x  rays 
produced  by  an  x  ray  tube  is  given  by  the  relation 

where  h  is  Planck's  constant,  q^  is  the  charge  of 
an  electron,  and  V  is  the  potential  diff'erence  at 
which  the  tube  operates.  If  V  is  50,000  volts, 
what  is/„a_r? 


18.15  The  equation  giving  the  maximum  energy 
of  the  X  rays  in  the  preceding  problem  looks  hke 
one  of  the  equations  in  Einstein's  theory  of  the 
photoelectric  effect.  How  would  you  account  for 
this  similarity?  For  the  difference? 

18.16  What  potential  difference  must  be  applied 
across  an  x-ray  tube  for  it  to  emit  x  rays  with 

a  minimum  wavelength  of  10""  m?  What  is  the 
energy  of  these  x  rays  in  joules?  In  electron  volts? 

18.17  A  glossary  is  a  collection  of  terms  Umited 
to  a  special  field  of  knowledge.  Make  a  glossary  of 
terms  that  appeared  for  the  first  time  in  this 
course  in  Chapter  18.  Make  an  informative 
statement  or  definition  for  each  term. 

18.18  In  his  Opticks,  Newton  proposed  a  set  of 
hypotheses  about  light  which,  taken  together, 
constituted  a  fairly  successful  model  of  hght. 
The  hypotheses  were  stated  as  questions.  Three  of 
the  hypotheses  are  given  below: 

Are  not  all  hypotheses  erroneous,  in  which 
light  is  supposed  to  consist  in  pression  or 
motion  waves  .  .  .  ?  [Quest.  28] 

Are  not  the  rays  of  light  very  small  bodies 
emitted  from  shining  substances?  [Quest.  29] 

Are  not  gross  bodies  and  light  convertible 
into  one  another,  and  may  not  bodies  receive 
much  of  their  activity  from  the  particles  of 
hght  which  enter  their  composition? 
[Quest.  30] 

(a)  In  what  respect  is  Newton's  model  similar 
to  and  different  from  the  photon  model  of 
hght? 

(b)  Why  would  Newton's  model  be  insufficient 
to  explain  the  photoelectric  effect?  What 
predictions  can  we  make  with  the  photon 
model  that  we  cannot  with  Newton's? 


57 


19.1  Spectra  of  gases  59 

19.2  Regularities  in  the  hydrogen  spectrum  63 

19.3  Rutherford's  nuclear  model  of  the  atom  66 

19.4  Nuclear  charge  and  size  69 

19.5  The  Bohr  theory:  the  postulates  71 

19.6  The  size  of  the  hydrogen  atom  72 

19.7  Other  consequences  of  the  Bohr  model  74 

19.8  The  Bohr  theory:  the  spectral  series  of  hydrogen  75 

19.9  Stationary  states  of  atoms:  the  Franck-Hertz  experiment  79 

19.10  The  periodic  table  of  the  elements  82 

19.11  The  inadequacy  of  the  Bohr  theory,  and  the  state  of 

atomic  theory  in  the  early  1920's  86 


Sculpture  representing  the  Bohr 
model  of  a  sodium  atom. 


CHAPTER  NINETEEN 


The  Rutherford-Bohr  Model  of  the  Atom 


19.1  Spectra  of  gases 

One  of  the  first  real  clues  to  our  understanding  of  atomic 
structure  was  provided  by  the  study  of  the  emission  and  absorption 
of  light  by  samples  of  the  elements.  The  results  of  this  study  are  so  SG  19.1 

important  to  our  story  that  we  shall  review  the  history  of  their 
development  in  some  detail. 

It  had  long  been  known  that  light  is  emitted  by  gases  or  vapors 
when  they  are  excited  in  any  one  of  several  ways:  by  heating  the 
gas  to  a  high  temperature,  as  when  a  volatile  substance  is  put  into  a 
flame;  by  an  electric  discharge  through  gas  in  the  space  between 
the  terminals  of  an  electric  arc;  by  a  continuous  electric  current 
in  a  gas  at  low  pressure  (as  in  the  now  familiar  "neon  sign"). 

The  pioneer  experiments  on  light  emitted  by  various  excited 
gases  were  made  in  1752  by  the  Scottish  physicist  Thomas  Melvill. 
He  put  one  substance  after  another  in  a  flame;  and  "having  placed 
a  pasteboard  with  a  circular  hole  in  it  between  my  eye  and  the 
flame  .  .  .  ,  I  examined  the  constitution  of  these  different  lights  with 
a  prism."  Melvill  found  the  spectrum  of  light  from  a  hot  gas  to  be 
different  from  the  well-known  continuum  of  rainbow  colors  found 
in  the  spectrum  of  a  glowing  solid  or  liquid.  Melvill's  spectrum 
consisted,  not  of  an  unbroken  stretch  of  color  continuously  graded 
from  violet  to  red,  but  of  individual  patches,  each  having  the  color 
of  that  part  of  the  spectrum  in  which  it  was  located,  and  with  dark 
gaps  (missing  colors)  between  the  patches.  Later,  when  more 
general  use  was  made  of  a  narrow  slit  through  which  to  pass  the 
light,  the  emission  spectrum  of  a  gas  was  seen  as  a  set  of  bright 
lines  (see  the  figure  in  the  margin  on  p.  61);  the  bright  lines  are  in 
fact  colored  images  of  the  slit.  The  existence  of  such  spectra  shows 
that  light  from  a  gas  is  a  mixture  of  only  a  few  definite  colors  or 
narrow  wavelength  regions  of  light. 

Melvill  also  noted  that  the  colors  and  locations  of  the  bright 
spots  were  different  when  different  substances  were  put  in  the 
flame.  For  example,  with  ordinary  table  salt  in  the  flame,  the 

59 


Hot  solids  emit  all  wavelengths  of  light,  producing  a  continu- 
ous spectrum  on  the  screen  at  right.  The  shorter-wavelength 
portions  of  light  are  refracted  more  by  the  prism  than  are  long 
wavelengths. 


Hot  gases  emit  only  certain  wavelengths  of  light,  producing  a 
"bright  line"  spectrum.  If  the  slit  had  a  different  shape,  so 
would  the  bright  lines  on  the  screen. 


1 


Cool  gases  absorb  only  certain  wavelengths  of  light,  produc- 
ing a  "dark  line"  spectrum  when  "white"  light  from  a  hot 
solid  is  passed  through  the  cool  gas. 


Section  19.1 


61 


predominant  color  was  "bright  yellow"  (now  known  to  be  character- 
isitic  of  the  element  sodium).  In  fact,  the  line  emission  spectrum  is 
markedly  different  for  each  chemically  different  gas  because  each 
chemical  element  emits  its  own  characteristic  set  of  wavelengths 
(see  the  figure  in  the  margin).  In  looking  at  a  gaseous  source  with- 
out the  aid  of  a  prism  or  a  grating,  the  eye  combines  the  separate 
colors  and  perceives  the  mixture  as  reddish  for  glowing  neon,  pale 
blue  for  nitrogen,  yellow  for  sodium  vapor,  and  so  on. 

Some  gases  have  relatively  simple  spectra.  Thus  the  most 
prominent  part  of  the  visible  spectrum  of  sodium  vapor  is  a  pair  of 
bright  yellow  lines.  Some  gases  or  vapors  have  exceedingly  complex 
spectra.  Iron  vapor,  for  example,  has  some  6000  bright  lines  in  the 
visible  range  alone. 

In  1823  the  British  astronomer  John  Herschel  suggested  that 
each  gas  could  be  identified  from  its  unique  line  spectrum.  By  the 
early  1860's  the  physicist  Gustave  R.  Kirchhoff  and  the  chemist 
Robert  W.  Bunsen,  in  Germany,  had  jointly  discovered  two  new 
elements  (rubidium  and  cesium)  by  noting  previously  unreported 
emission  lines  in  the  spectrum  of  the  vapor  of  a  mineral  water.  This 
was  the  first  of  a  series  of  such  discoveries;  it  started  the  develop- 
ment of  a  technique  making  possible  the  speedy  chemical  analysis 
of  small  amounts  of  materials  by  spectrum  analysis. 

In  1802  the  English  scientist  William  Wollaston  saw  in  the 
spectrum  of  sunlight  something  that  had  been  overlooked  before. 
Wollaston  noticed  a  set  of  seven  sharp,  irregularly  spaced  dark  lines 
across  the  continuous  solar  spectrum.  He  did  not  understand  why 
they  were  there,  and  did  not  carry  the  investigation  further.  A  dozen 
years  later,  the  German  physicist,  Joseph  von  Fraunhofer,  used 
better  instruments  and  detected  many  hundreds  of  such  dark  lines 
To  the  most  prominent  dark  lines,  Fraunhofer  assigned  the  letters 
A,  B,  C,  etc.  These  dark  lines  can  be  easily  seen  in  the  sun's 
spectrum  with  even  quite  simple  modem  spectroscopes,  and  his 
letters  A.  B,  C  .  .  .  are  still  used  to  identifv  them. 


^o! 


Parts  of  the  line  emission  spectra 
of  mercury  (Hg)  and  helium  (He), 
redrawn  from  photographic  records. 


Spectroscope:  A  device  for 
examining  the  spectrum  by  eye. 

Spectrometer  or  spectrograph: 
A  device  for  measuring  the  wave 
length  of  the  spectrum  and  for 
recording  the  spectra  (for  example 
on  film). 


The  Fraunhofer  dark  lines  in  the 
visible  part  of  the  solar  spectrum: 
only  a  few  of  the  most  prominent 
lines  are  represented. 


In  the  spectra  of  several  other  bright  stars,  Fraunhofer  found 
similar  dark  lines;  many  of  them,  although  not  all,  were  in  the  same 
positions  as  those  in  the  solar  spectrum. 

The  key  observations  toward  a  better  understanding  of  both 
the  dark-line  and  the  bright-line  spectra  of  gases  were  made  by 
Kirchhoff  in  1859.  By  that  time  it  was  known  that  the  two  promi- 
nent yellow  lines  in  the  emission  spectrum  of  heated  sodium  vapor 
in  the  laboratory  had  the  same  wavelengths  as  two  neighboring 
prominent  dark  lines  in  the  solar  spectrum  to  which  Fraunhofer  had 


62 


The  Rutherford-Bohr  Model  of  the  Atom 


absorption 
spectrum 


emission 
spectrum 


assigned  the  letter  D.  It  was  also  known  that  the  light  emitted  by  a 
glowing  solid  forms  a  perfectly  continuous  spectrum  that  shows  no 
dark  hnes.  Kirchhoff  now  demonstrated  that  if  the  light  from  a 
glowing  solid,  as  on  page  60.  is  allowed  first  to  pass  through  cooler 
sodium  vapor  and  is  then  dispersed  by  a  prism,  the  spectrum 
exhibits  two  prominent  dark  lines  at  the  same  place  in  the  spectrum 
as  the  D-lines  of  the  sun's  spectrum.  It  was  therefore  reasonable 
to  conclude  that  the  light  from  the  sun,  too,  was  passing  through  a 
mass  of  sodium  gas.  This  was  the  first  evidence  of  the  chemical 
composition  of  the  gas  envelope  around  the  sun. 

■MM  ■■■—■■ 


ultraviolet 


visible 


>4 infrared 


Comparison  of  the  line  absorption 
spectrum  and  line  emission  spectrum 
of  sodium  vapor. 


SG  19.2 


When  Kirchhoff 's  experiment  was  repeated  with  other  relatively 
cool  gases  placed  between  a  glowing  solid  and  the  prism,  each  gas 
was  found  to  produce  its  own  characteristic  set  of  dark  lines. 
Evidently  each  gas  in  some  way  absorbs  light  of  certain  wave- 
lengths from  the  passing  "white"  light.  More  interesting  still, 
Kirchhoff  showed  that  the  wavelength  corresponding  to  each 
absorption  line  is  equal  to  the  wavelength  of  a  bright  line  in  the 
emission  spectrum  of  the  same  gas.  The  conclusion  is  that  a  gas 
can  absorb  only  light  of  these  wavelengths  which,  when  excited,  it 
can  emit.  But  note  that  not  every  emission  line  is  represented  in 
the  absorption  spectrum.  (Soon  you  will  see  why.) 

Each  of  the  various  Fraunhofer  lines  across  the  spectrum  of  the 
sun  and  also  of  far  more  distant  stars  have  now  been  identified  with 
the  action  of  some  gas  as  tested  in  the  laboratory,  and  thereby  the 
whole  chemical  composition  of  the  outer  region  of  the  sun  and  other 
stars  has  been  determined.  This  is  really  quite  breathtaking  from 
several  points  of  view:  (a)  that  it  could  be  possible  to  find  the 
chemical  composition  of  immensely  distant  objects;  (b)  that  the 
chemical  materials  there  are  the  same  as  those  in  our  own  sur- 
roundings on  earth,  as  shown  by  the  fact  that  even  the  most 
complex  absorption  spectra  are  faithfully  reproduced  in  the  star 
spectra;  and  (c)  that  therefore  the  physical  processes  in  the  atom 
that  are  responsible  for  absorption  must  be  the  same  here  and 
there.  In  these  facts  we  have  a  hint  of  how  universal  physical  law 
really  is:  even  at  the  outermost  edges  of  the  cosmos  from  which  we 
get  any  light  with  absorbed  wavelengths,  the  laws  of  physics  appear 
to  be  the  same  as  for  common  materials  close  at  hand  in  our 
laboratory!  This  is  just  what  GaUleo  and  Newton  had  intuited  when 


Section  19.2 


63 


they  proposed  that  there  is  no  difference  between  terrestrial  and 
celestial  physics. 

Q1     What  can  you  infer  about  the  source  if  its  light  gives  a 
bright  line  spectrum? 

Q2    What  can  you  infer  about  the  source  if  its  light  gives  a  dark 
line  spectrum? 

Q3    What  evidence  is  there  that  the  physics  and  chemistry  of 
materials  at  great  distances  from  us  is  the  same  as  of  matter  close 
at  hand? 


19.2     Regularities  in  the  hydrogen  spectrum 


Of  all  the  spectra,  the  line  emission  spectrum  of  hydrogen  is 
especially  interesting  for  both  historical  and  theoretical  reasons.  In 
the  visible  and  near  ultraviolet  regions,  the  emission  spectrum 
consists  of  an  apparently  systematic  series  of  Hnes  whose  positions 
are  indicated  at  the  right.  In  1885,  a  Swiss  school  teacher,  Johann 
Jakob  Balmer,  found  a  simple  formula- an  empirical  relation- 
which  gave  the  wavelengths  of  the  lines  known  at  the  time.  The 
formula  is: 


Johann  Jakob  Balmer  (1825-1898), 
a  teacher  at  a  girls'  school  in 
Switzerland,  came  to  study  wave- 
lengths of  spectra  listed  in  tables 
through  his  interest  in  mathematical 
puzzles  and  numerology. 


\  =  b 


n' 


n^-2^ 


Where  b  is  a  constant  which  Balmer  determined  empirically  and 
found  to  be  equal  to  3645.6  A,  and  n  is  a  whole  number,  different  for 
each  line.  Specifically,  to  give  the  observed  value  for  the  wave- 
length, n  must  be  3  for  the  first  (red)  line  of  the  hydrogen  emission 
spectrum  (named  HJ;  n  =  4  for  the  second  (green)  line  (H^);  n  ==  5 
for  the  third  (blue)  line  (H;,);  and  n  =  6  for  the  fourth  (violet)  line 
(Hg).  The  table  below  shows  excellent  agreement  (within  0.02%) 
between  the  values  Balmer  computed  from  his  empirical  formula 
and  previously  measured  values. 


NAME 
OF  LINE 


Wavelength  A  (in  A) 

FROM  BALMER'S  BY  ANGSTROM'S 

FORMULA  MEASUREMENT 


DIFFERENCE 


H„ 

3 

6562.08 

6562.10 

+0.02 

H, 

4 

4860.8 

4860.74 

-0.06 

H. 

5 

4340 

4340.1 

+  0.1 

H« 

6 

4101.3 

4101.2 

-0.1 

The  Balmer  lines  of  hydrogen;  re- 
drawn from  a  photograph  made  with 
a  film  sensitive  to  ultraviolet  light  as 
well  as  visible.  The  lines  get  more 
crowded  as  they  approach  the  series 
limit  in  the  ultraviolet. 


Data  on  hydrogen  spectrum  (as  given 
in  Balmer's  paper  of  1885). 


It  took  nearly  30  years  before  anyone  understood  why  Balmer's 
empirical  formula  worked  so  well -why  the  hydrogen  atom  emitted 
light  whose  wavelength  made  such  a  simple  sequence.  But  this  did 
not  keep  Balmer  from  speculating,  that  there  might  be  other  series  of 


■H.o 


64  The  Rutherford-Bohr  Model  of  the  Atom 

hither-to  unsuspected  lines  in  the  hydrogen  spectrum,  and  that  their 
wavelengths  could  be  found  by  replacing  the  2'^  in  the  denominator 
of  his  equation  by  other  numbers  such  as  P,  3^,  4-,  and  so  on.  This 
suggestion,  which  stimulated  many  workers  to  search  for  such 
additional  spectral  series,  turned  out  to  be  fruitful,  as  we  shall 
discuss  shortly. 

To  use  modern  notation,  we  first  rewrite  B aimer's  formula  in  a 
form  that  will  be  more  useful. 


1-^ 


In  this  equation,  which  can  be  derived  from  the  first  one,  R^  is 
a  constant,  equal  to  4/b.  (It  is  called  the  Rydberg  constant  for 
hydrogen,  in  honor  of  the  Swedish  spectroscopist  J.  R.  Rydberg 
who,  following  B aimer,  made  great  progress  in  the  search  for 
various  spectral  series.)  The  series  of  Unes  described  by  B aimer's 
formula  are  called  the  Balmer  series.  While  Balmer  constructed  his 
formula  from  known  X  of  only  four  lines,  his  formula  predicted  that 
there  should  be  many  more  lines  in  the  same  series  (indeed, 
infinitely  many  such  lines  as  n  takes  on  values  such  as  n  =  3,  4,  5, 
6,  7,  8,  .  .  .  oo).  The  figure  in  the  margin  indicates  that  this  has 
indeed  been  observed  — and  every  one  of  the  lines  is  correctly  pre- 
dicted by  Balmer's  formula  with  considerable  accuracy. 

If  we  follow  Balmer's  speculative  suggestion  of  replacing  2^ 
by  other  numbers,  we  obtain  the  possibilities: 


k~^"{v~n^ 


X'^'^fe     1? 


1- 


"W      nV 


and  so  on.  Each  of  these  equations  describes  a  possible  series.  All 
these  hypothetical  series  of  lines  can  then  be  summarized  in  one 
overall  formula: 


k      ^"[n/      n,V 


Part  of  the  absorption  spectrum 
of  the  star  Rigel  ()3  Orion).  The 
dark  lines  are  at  the  same  loca- 
tion as  lines  due  to  absorption 
by  hydrogen  gas  in  the  ultra- 
violet region;  they  match  the 
lines  of  the  Balmer  series  as 
indicated  by  the  H  numbers 
(where  H,  would  be  H„,  H^  would 
be  H;,  etc.).  This  indicates  the 
presence  of  hydrogen  in  the 
star. 


where  n^  is  a  whole  number  that  is  fixed  for  any  one  series  for 
which  wavelengths  are  to  be  found  (for  example,  it  is  2  for  all  lines 
in  the  Balmer  series).  The  letter  n,  stands  for  integers  that  take  on 
the  values  n^^  +  1,  w^^  +  2,  n^^  +  3,  .  .  .  for  the  successive  individual 
lines  in  a  given  series  (thus,  for  the  first  two  lines  of  the  Balmer 
series,  n,  is  3  and  4.)  The  constant  R„  should  have  the  same  value 
for  all  of  these  hydrogen  series. 

So  far,  our  discussion  has  been  merely  speculation.  No  series, 
no  single  line  fitting  the  formula  in  the  general  formula,  need  exist 
(  —  except  for  the  observed  Bahner  series,  where  nf  =  2).  But  when 
physicists  began  to  look  for  these  hypothetical  lines  with  good 
spectrometers  — they  found  that  they  do  exist! 

In  1908.  F.  Paschen  in  Germany  found  two  hydrogen  lines  in 
the  infrared  whose  wavelengths  were  correctly  given  by  setting 
71/  =  3  and  n,  =  4  and  5  in  the  general  formula;  many  other  lines 


Section  19.2  65 

in  this  "Paschen  series"  have  since  been  identified.  With  improve- 
ments of  experimental  apparatus  and  techniques,  new  regions  of 
the  spectrum  could  be  explored,  and  thus  other  series  gradually 
were  added  to  the  Balmer  and  Paschen  series.  In  the  table  below 
the  name  of  each  series  listed  is  that  of  the  discoverer. 

Series  of  lines  in  the  hydrogen  spectrum 


NAME  OF 

DATE  OF 

REGION  OF 

VALUES  IN 

SERIES 

DISCOVERY 

SPECTRUM 

BALMER  EQUATION 

Lyman 

1906-1914 

ultraviolet 

n^  =  1 ,  n,  =  2,  3,  4, 

Balmer 

1885 

ultraviolet-visible 

rif  =  2,  n,.  =  3,  4,  5, 

Paschen 

1908 

infrared 

rif  =  3,  n,.  =  4,  5,  6 

Brackett 

1922 

infrared 

Pf  =  4,  n,.  =  5,  6,  7 

Pfund 

1924 

infrared 

n^=  5,  H;  =  6,  7,  8 

Balmer  had  also  hoped  that  his  formula  for  hydrogen  spectra 
might  be  a  pattern  for  finding  series  relationships  in  the  observed 
spectra  of  other  gases.  This  suggestion  bore  fruit  also.  While  his 
formula  itself  did  not  work  directly  in  describing  spectra  of  gases 
other  than  hydrogen,  it  inspired  formulas  of  similar  mathematical 
form  that  were  useful  in  expressing  order  in  portions  of  a  good 
many  complex  spectra.  The  Rydberg  constant  Rh  also  reappeared 
in  such  empirical  formulas. 

For  three  decades  after  Balmer's  success,  physicists  tried  to 
account  for  spectra  by  constructing  models  of  the  atom  that  would 
radiate  light  of  the  right  wavelengths.  But  the  great  number  and 
variety  of  spectral  lines,  emitted  even  by  the  simplest  atom, 
hydrogen,  made  it  difiicult  to  find  a  successful  model.  Eventually 
models  were  made  that  succeeded  in  revealing  the  origin  of 
spectra  and  in  this  chapter  and  the  next  one,  you  will  see  how  it 
was  done. 

What  you  have  already  learned  in  Chapter  18  about  quantum  SG  19.3-19.5 

theory  suggests  one  line  of  attack:  the  emission  and  absorption  of 
light  from  an  atom  must  correspond  to  a  decrease  and  an  increase 
of  the  amount  of  energy  the  atom  has.  If  atoms  of  an  element  emit 
light  of  only  certain  frequencies,  then  the  energy  of  the  atoms  must 
be  able  to  change  only  by  certain  amounts.  These  changes  of  energy 
must  belong  to  some  rearrangement  of  the  parts  of  the  atom. 

Q4  What  evidence  did  Balmer  have  that  there  were  other 
series  of  lines  in  the  hydrogen  spectrum,  with  terms  3^,  4^,  etc. 
instead  of  2^? 

Q5    Often  discoveries  result  from  grand  theories  (like 
Newton's)  or  from  a  good  intuitive  grasp  of  phenomena  (like 
Faraday's).  What  led  Balmer  to  his  relation  for  spectra? 

Q6    What  accounts  for  the  success  of  Balmer's  overall  formula 
in  predicting  new  series  of  the  emission  spectrum  of  hydrogen? 


66 


The  Rutherford-Bohr  Model  of  the  Atom 


SG  19.6 


HCTAu 


In  somewhat  the  same  way,  you 
could,  in  principle,  use  a  scattering 
experiment  to  discover  the  size  and 
shape  of  an  object  hidden  from  view 
in  a  cloud  or  fog  — by  directing  a 
series  of  projectiles  at  the  unseen 
object  and  tracing  their  paths  back 
after  deflection. 


19.3    Rutherford's  nuclear  model  of  the  atom 

A  new  basis  for  atomic  models  was  provided  during  the  period 
1909  to  1911  by  Ernest  Rutherford,  a  New  Zealander  who  had 
already  shown  ability  as  an  experimentalist  at  McGill  University  in 
Montreal,  Canada.  He  had  been  invited  in  1907  to  Manchester 
University  in  England  where  he  headed  a  productive  research 
laboratory.  Rutherford  was  specially  interested  in  the  rays  emitted 
by  radioactive  substances,  in  particular  in  a  (alpha)  rays.  As  we 
shall  see  in  Chapter  21,  a  rays  consist  of  positively  charged  particles. 
These  particles  are  positively  charged  helium  atoms  with  masses 
about  7500  times  greater  than  the  electron  mass.  Some  radioactive 
substances  emit  a  particles  at  rates  and  energies  great  enough  for 
the  particles  to  be  used  as  projectiles  to  bombard  samples  of  ele- 
ments. The  experiments  that  Rutherford  and  his  colleagues  did 
with  a  particles  are  examples  of  a  highly  important  kind  of 
experiment  in  atomic  and  nuclear  physics  — the  scattering 
experiment. 

In  a  scattering  experiment,  a  narrow,  parallel  beam  of  projec- 
tiles (for  example,  a  particles,  electrons,  x  rays)  is  aimed  at  a  target 
that  is  usually  a  thin  foil  or  film  of  some  material.  As  the  beam 
strikes  the  target,  some  of  the  projectiles  are  deflected,  or  scattered 
from  their  original  direction.  The  scattering  is  the  result  of  the 
interaction  between  the  particles  in  the  beam  and  the  atoms  of  the 
material.  A  careful  study  of  the  projectiles  after  they  have  been 
scattered  can  yield  information  about  the  projectiles,  the  atoms, 
or  both  — or  the  interaction  between  them.  Thus  if  we  know  the 
mass,  energy  and  direction  of  the  projectiles,  and  see  what  happens 
to  them  in  a  scattering  experiment,  we  can  deduce  properties  of 
the  atoms  that  scattered  the  projectiles. 

Rutherford  noticed  that  when  a  beam  of  a  particles  passed 
through  a  thin  metal  foil,  the  beam  spread  out.  The  scattering  of  a 
particles  can  be  imagined  to  be  caused  by  the  electrostatic  forces 
between  the  positively  charged  a  particles  and  the  charges  that 
make  up  atoms.  Since  atoms  contain  both  positive  and  negative 
charges,  an  a  particle  is  subjected  to  both  repulsive  and  attractive 
forces  as  it  passes  through  matter.  The  magnitude  and  direction  of 
these  forces  depend  on  how  near  the  particle  happens  to  approach 
to  the  centers  of  the  atoms  among  which  it  moves.  When  a  particu- 
lar atomic  model  is  proposed,  the  extent  of  the  expected  scattering 
can  be  calculated  and  compared  with  experiment.  In  the  case  of  the 
Thomson  model  of  the  atom,  calculation  showed  that  the  probability 
is  so  negligibly  small  that  an  a  particle  would  be  scattered  through 
an  angle  of  more  than  a  few  degrees. 

The  breakthrough  to  the  modern  model  of  the  atom  came  when 
one  of  Rutherford's  assistants,  Hans  Geiger,  found  that  the  number 
of  particles  scattered  through  angles  of  10°  or  more  was  much 
greater  than  the  number  predicted  on  the  basis  of  the  Thomson 
model.  In  fact,  one  out  of  about  every  8000  a  particles  was  scattered 
through  an  angle  greater  than  90°.  Thus  a  significant  number  of  a 
particles  virtually  bounced  right  back  from  the  foil.  This  result  was 
entirely  unexpected  on  the  basis  of  Thomson's  model  of  the  atom. 


Section  19.3 


67 


Ernest  Rutherford  (1871-1937)  was  born,  grew  up.  and 
received  most  of  his  education  in  New  Zealand.  At 
age  24  he  went  to  Cambridge,  England  to  work  at 
the  Cavendish  Laboratory  under  J.  J.  Thomson.  From 
there  he  went  to  McGill  University  in  Canada,  then 
home  to  be  married  and  back  to  England  again,  now 
to  Manchester  University.  At  these  universities,  and 
later  at  the  Cavendish  Laboratory  where  he  succeeded 
J.  J.  Thomson  as  director,  Rutherford  performed 
important  experiments  on  radioactivity,  the  nuclear 
nature  of  the  atom,  and  the  structure  of  the  nucleus. 
Rutherford  introduced  the  concepts  alpha,"  "beta" 
and  gamma"  rays,  "protons,"  and  "half-life."  His 
contributions  will  be  further  discussed  in  Unit  6.  For 
his  scientific  work,  Rutherford  was  knighted  and 
received  a  Nobel  Prize. 


by  which  the  atom  should  have  acted  on  the  projectile  more  like  a 
cloud  in  which  fine  dust  is  suspended.  Some  years  later,  Rutherford 
wrote: 


...  I  had  observed  the  scattering  of  a-particles.  and  Dr. 
Geiger  in  my  laboratory  had  examined  it  in  detail.  He 
found,  in  thin  pieces  of  heavy  metal,  that  the  scattering 
was  usually  small,  of  the  order  of  one  degree.  One  day 
Geiger  came  to  me  and  said.  "Don't  you  think  that  young 
Marsden,  whom  I  am  training  in  radioactive  methods, 
ought  to  begin  a  small  research?"  Now  I  had  thought 
that,  too,  so  I  said,  "Why  not  let  him  see  if  any  a-particles 
can  be  scattered  through  a  large  angle?""  I  may  tell  you  in 
confidence  that  I  did  not  believe  that  they  would  be,  since 
we  knew  that  the  a-particle  was  a  very^  fast,  massive 
particle,  with  a  great  deal  of  [kinetic]  energy,  and  you 
could  show  that  if  the  scattering  was  due  to  the  accumu- 
lated effect  of  a  number  of  small  scatterings,  the  chance 
of  an  a-particle"s  being  scattered  backward  was  very 
small.  Then  I  remember  two  or  three  days  later  Geiger 
coming  to  me  in  great  excitement  and  saying,  "We  have 


68 


The  Rutherford-Bohr  Model  of  the  Atom 


been  able  to  get  some  of  the  a-particles  coming  back- 
ward .  .  ."  It  was  quite  the  most  incredible  event  that 
has  ever  happened  to  me  in  my  life.  It  was  almost  as 
incredible  as  if  you  fired  a  15-inch  shell  at  a  piece  of 
tissue  paper  and  it  came  back  and  hit  you.  On  considera- 
tion, I  realized  that  this  scattering  backward  must  be  the 
result  of  a  single  collision,  and  when  I  made  calculations 
I  saw  that  it  was  impossible  to  get  anything  of  that  order 
of  magnitude  unless  you  took  a  system  in  which  the 
greater  part  of  the  mass  of  the  atom  was  concentrated  in 
a  minute  nucleus.  It  was  then  that  I  had  the  idea  of  an 
atom  with  a  minute  massive  centre,  carrying  a  charge. 


SG  19.6,  19.7 


Paths  of  two  a  particles  A  and  A'  ap- 
proaching a  nucleus  N.  (Based  on 
Rutherford,  Philosophical  Magazine. 
vol.  21  (1911),  p.  669.) 


oc 


¥ 


Rutherford's  scintillation  apparatus 
was  placed  in  an  evacuated  chamber 
so  that  tne  a  particles  would  not  be 
slowed  down  by  collisions  with  air 
molecules. 


These  experiments  and  Rutherford's  interpretation  marked  the 
origin  of  the  modern  concept  of  the  nuclear  atom.  Let  us  look  at 
the  experiments  more  closely  to  see  why  Rutherford  concluded  that 
the  atom  must  have  its  mass  and  positive  charge  concentrated  in  a 
tiny  space  at  the  center,  thus  forming  a  nucleus  about  which  the 
electrons  are  clustered. 

A  possible  explanation  of  the  observed  scattering  is  that  there 
exist  in  the  foil  concentrations  of  mass  and  charge  — positively 
charged  nuclei  — much  more  dense  than  in  Thomson's  atoms.  An  a. 
particle  heading  directly  toward  one  of  them  is  stopped  and  turned 
back,  as  a  ball  would  bounce  back  from  a  rock  but  not  from  a  cloud 
of  dust  particles.  The  figure  in  the  margin  is  based  on  one  of 
Rutherford's  diagrams  in  his  paper  of  1911,  which  may  be  said 
to  have  laid  the  foundation  for  the  modern  theory  of  atomic 
structure.  It  shows  two  positively  charged  a  particles,  A  and  A'. 
The  a  particle  A  is  heading  directly  toward  a  massive  nucleus  N. 
If  the  nucleus  has  a  positive  electric  charge,  it  will  repel  the 
positive  oc  particle.  Because  of  the  electrical  repulsive  force 
between  the  two,  A  is  slowed  to  a  stop  at  some  distance  r  from  N. 
and  then  moves  directly  back.  A'  is  another  a  particle  that  is  not 
headed  directly  toward  the  nucleus  N;  it  is  repelled  by  N  along  a 
path  which  calculation  showed  must  be  an  hyperbola.  The  deflection 
of  A'  from  its  original  path  is  indicated  by  the  angle  (/>. 

Rutherford  considered  the  effects  on  the  path  of  the  a.  particle 
due  to  the  important  variables  — the  a  particle's  speed,  the  foil 
thickness,  and  the  quantity  of  charge  Q  on  each  nucleus.  According 
to  the  model  most  of  the  a  particles  should  be  scattered  through 
small  angles,  because  the  chance  of  approaching  a  very  small 
nucleus  nearly  head-on  is  so  small;  but  a  noticeable  number  of  a 
particles  should  be  scattered  through  large  angles. 

Geiger  and  Marsden  undertook  tests  of  these  predictions  with 
the  apparatus  shown  schematically  in  the  margin.  The  lead  box  B 
contains  a  radioactive  substance  (radon)  which  emits  a  particles. 
The  particles  emerging  from  the  small  hole  in  the  box  are  deflected 
through  various  angles  4>  in  passing  through  the  thin  metal  foil  F. 
The  number  of  particles  deflected  through  each  angle  <i>  is  found 
by  letting  the  particles  strike  a  zinc  sulftide  screen  S.  Each  a 
particle  that  strikes  the  screen  produces  a  scintillation  (a  momen- 


Section  19.4 


69 


tary  pinpoint  of  fluorescence).  These  scintillations  can  be  observed 
and  counted  by  looking  through  the  microscope  M;  S  and  M  can  be 
moved  together  along  the  arc  of  a  circle.  In  later  experiments,  the 
number  of  a  particles  at  any  angle  </>  was  counted  more  conven- 
iently by  replacing  S  and  M  by  a  counter  invented  by  Geiger  (see 
sketch  in  the  margin).  The  Geiger  counter,  in  its  more  recent 
versions,  is  now  a  standard  laboratory  item. 

Geiger  and  Marsden  found  that  the  number  of  a  particles 
counted  depended  on  the  scattering  angle,  the  speed  of  the  particles, 
and  on  the  thickness  of  the  foil  of  scattering  material,  just  as 
Rutherford  had  predicted.  This  bore  out  the  model  of  the  atom  in 
which  most  of  the  mass  and  all  positive  charge  are  concentrated  in 
a  very  small  region  at  the  center  of  the  atom. 


Q7    Why  are  a  particles  scattered  by  atoms?  Why  is  the  angle 
of  scattering  mostly  small  but  sometimes  large? 

Q8     What  was  the  basic  diff'erence  between  the  Rutherford 
and  the  Thomson  models  of  the  atom? 


19.4     Nuclear  charge  and  size 

At  the  time  Rutherford  made  his  predictions  about  the  effect  of 
the  speed  of  the  a  particle  and  the  thickness  of  foil  on  the  angle  of 
scattering,  there  was  no  way  to  measure  independently  the 
nucleus  charge  Q  which  he  had  to  assume.  However,  some  of 
Rutherford's  predictions  were  confirmed  by  scattering  experiments 
and,  as  often  happens  when  part  of  a  theory  is  confirmed,  it  is 
reasonable  to  proceed  temporarily  as  if  the  whole  of  that  theory  were 
justified.  That  is,  pending  further  proof,  one  could  assume  that  the 
value  of  Q  needed  to  explain  the  observed  scattering  data  was  the 
correct  value  of  Q  for  the  actual  nucleus.  On  this  basis,  from  the 
scattering  by  different  elements  — among  them  carbon,  aluminum 
and  gold  — the  following  nuclear  charges  were  obtained:  for  carbon, 
Q  =  6qp,  for  aluminum,  Q  =  13  or  Mg^,  and  for  gold,  Q  =  78  or  IGq^. 
Similarly,  tentative  values  were  found  for  other  elements. 

The  magnitude  of  the  positive  charge  of  the  nucleus  was  an 
important  and  welcome  piece  of  information  about  the  atom.  If  the 
nucleus  has  a  positive  charge  of  6  q^,  13  to  14  q^,  etc.,  the  number 
of  electrons  surrounding  the  nucleus  must  be  6  for  carbon,  13  or  14 
for  aluminum,  etc.,  since  the  atom  as  a  whole  is  electrically  neutral. 
This  gave  for  the  first  time  a  good  idea  of  just  how  many  electrons 
an  atom  may  have.  But  even  more  important,  it  was  soon  noticed 
that  for  each  element  the  value  found  for  the  nuclear  charge  — in 
multiples  of  q^- was  close  to  the  atomic  number  Z,  the  place 
number  of  that  element  in  the  periodic  table!  While  the  results  of 
experiments  on  the  scattering  of  a  particles  were  not  yet  precise 
enough  to  permit  this  conclusion  to  be  made  with  certainty,  the 
data  indicated  that  each  nucleus  has  a  positive  charge  Q  numer- 
ically equal  to  Zq^. 

The  suggestion  that  the  number  of  positive  charges  on  the 


SG  19.8 


A  Geiger  counter  (1928).  It  consists 
of  a  metal  cylinder  C  containing  a  gas 
and  a  thin  axial  wire  A  that  is  insulated 
from  the  cylinder.  A  potential  differ- 
ence slightly  less  than  that  needed 
to  produce  a  discharge  through  the 
gas  is  maintained  between  the  wire 
(anode  A)  and  cylinder  (cathode  C). 
When  an  a  particle  enters  through  the 
thin  mica  window  (W),  it  frees  a  few 
electrons  from  the  gas  molecules. 
The  electrons  are  accelerated  toward 
the  anode,  freeing  more  electrons 
along  the  way  by  collisions  with  gas 
molecules.  The  avalanche  of  electrons 
constitutes  a  sudden  surge  of  current 
which  may  be  amplified  to  produce  a 
click  in  the  loudspeaker  (L)  or  to  oper- 
ate a  register  (as  in  the  Project  Physics 
scaler,  used  in  experiments  in  Unit  6). 


q,  =  numerical  value  of  charge  of 
one  electron. 


70 


The  Rutherford-Bohr  Model  of  the  Atom 


The  central  dot  representing  the 
nucleus  in  relation  to  the  size  of  the 
atom  as  a  whole  is  about  100  times 
too  large.  Popular  diagrams  of  atoms 
often  greatly  exaggerate  the  relative 
size  of  the  nucleus,  (perhaps  in  order 
to  suggest  the  greater  mass). 


nucleus  and  also  the  number  of  electrons  around  the  nucleus  are 
equal  to  the  atomic  number  Z  made  the  picture  of  the  nuclear  atom 
at  once  much  clearer  and  simpler.  On  this  basis,  the  hydrogen 
atom  (Z  =  1)  has  one  electron  outside  the  nucleus;  a  helium  atom 
(Z  =  2)  has  in  its  neutral  state  two  electrons  outside  the  nucleus; 
a  uranium  atom  (Z  =  92)  has  92  electrons.  This  simple  scheme  was 
made  more  plausible  when  additional  experiments  showed  that  it 
was  possible  to  produce  singly  ionized  hydrogen  atoms,  H^,  and 
doubly  ionized  helium  atoms,  He^^,  but  not  H^^  or  He^^^  — evidently 
because  a  hydrogen  atom  has  only  one  electron  to  lose,  and  a 
helium  atom  only  two.  Unexpectedly,  the  concept  of  the  nuclear 
atom  thus  provided  new  insight  into  the  periodic  table  of  the 
elements:  it  suggested  that  the  periodic  table  is  really  a  listing  of 
the  elements  according  to  the  number  of  electrons  around  the 
nucleus,  or  according  to  the  number  of  positive  units  of  charge  on 
the  nucleus. 

These  results  made  it  possible  to  understand  some  of  the  dis- 
crepancies in  Mendeleev's  periodic  table.  For  example,  the  elements 
tellurium  and  iodine  had  been  put  into  positions  Z  =  52  and  Z  =  53 
on  the  basis  of  their  chemical  properties,  contrary  to  the  order  of 
their  atomic  weights.  Now  that  Z  was  seen  to  correspond  to  a 
fundamental  fact  about  the  nucleus,  the  reversed  order  of  their 
atomic  weights  was  understood  to  be  a  curious  accident  rather 
than  a  basic  fault  in  the  scheme. 

As  an  important  additional  result  of  these  scattering  experi- 
ments the  size  of  the  nucleus  may  be  estimated.  Suppose  an  a 
particle  is  moving  directly  toward  a  nucleus.  Its  kinetic  energy  on 
approach  is  transformed  into  electrical  potential  energy.  It  slows 
down  and  eventually  stops.  The  distance  of  closest  approach  may 
be  computed  from  the  original  kinetic  energy  of  the  a  particle  and 
the  charges  of  a  particle  and  nucleus.  (See  SG  19.8.)  The  value 
calculated  for  the  closest  approach  is  approximately  3  x  10 "'^m.  If 
the  a  particle  is  not  penetrating  the  nucleus,  this  distance  must  be 
at  least  as  great  as  the  sum  of  the  radii  of  oc  particle  and  nucleus; 
so  the  radius  of  the  nucleus  could  not  be  larger  than  about  10"'*m, 
only  about  1/1000  of  the  known  radius  of  an  atom.  Thus  if  one 
considers  its  volume,  which  is  proportional  to  the  cube  of  the  radius, 
it  is  clear  that  the  atom  is  mostly  empty,  with  the  nucleus  occupying 
only  one  billionth  of  the  space!  This  in  turn  explains  the  ease  with 
which  a  particles  or  electrons  penetrate  thousands  of  layers  of 
atoms  in  metal  foils  or  in  gases,  with  only  occasional  large 
deflection  backward. 

Successful  as  this  model  of  the  nuclear  atom  was  in  explaining 
scattering  phenomena,  it  raised  many  new  questions:  What  is  the 
arrangement  of  electrons  about  the  nucleus?  What  keeps  the 
negative  electron  from  falling  into  a  positive  nucleus  by  electrical 
attraction?  Of  what  is  the  nucleus  composed?  What  keeps  it  from 
exploding  on  account  of  the  repulsion  of  its  positive  charges? 
Rutherford  openly  realized  the  problems  raised  by  these  questions, 
and  the  failure  of  his  model  to  answer  them.  But  he  rightly  said 
that  one  should  not  expect  one  model,  made  on  the  basis  of  one 


Section  19.5  71 

set  of  puzzling  results  which  it  handled  well,  also  to  handle  all 

other  puzzles.  Additional  assumptions  were  needed  to  complete  the  SG  19.9 

model  — to  find  answers  to  the  additional  questions  posed  about  the 

details  of  atomic  structure.  The  remainder  of  this  chapter  will  deal 

with  the  theory  proposed  by  Niels  Bohr,  a  young  Danish  physicist 

who  joined  Rutherford's  group  just  as  the  nuclear  model  was  being 

announced. 

Q9    What  does  the  "atomic  number"  of  an  element  refer  to, 
according  to  the  Rutherford  model  of  the  atom? 

Q10       What  is  the  greatest  positive  charge  that  an  ion  of 
lithium  (the  next  heaviest  element  after  helium)  could  have? 


19.5    The  Bohr  theory:  the  postulates 

If  an  atom  consists  of  a  positively  charged  nucleus  surrounded 
by  a  number  of  negatively  charged  electrons,  what  keeps  the 
electrons  from  falling  into  the  nucleus  — from  being  pulled  in  by  the 
electric  force  of  attraction?  One  possible  answer  to  this  question 
is  that  an  atom  may  be  like  a  planetary  system  with  the  electrons 
revolving  in  orbits  around  the  nucleus.  Instead  of  the  gravitational 
force,  the  electric  attractive  force  between  the  nucleus  and  an 
electron  would  supply  a  centripetal  force  that  would  tend  to  keep 
the  moving  electron  in  orbit. 

Although  this  idea  seems  to  start  us  on  the  road  to  a  theory  of 
atomic  structure,  a  serious  problem  arises  concerning  the  stability 
of  a  planetary  atom.  According  to  Maxwell's  theory  of  electro- 
magnetism,  a  charged  particle  radiates  energy  when  it  is 
accelerated.  Now.  an  electron  moving  in  an  orbit  around  a  nucleus 
continually  changes  its  velocity  vector,  always  being  accelerated  by 
the  centripetal  electric  force.  The  electron,  therefore,  should  lose 
energy  by  emitting  radiation.  A  detailed  analysis  of  the  motion  of 
the  electron  shows  that  the  electron  should  be  drawn  closer  to  the 
nucleus,  somewhat  as  an  artificial  satellite  that  loses  energy  due 
to  friction  in  the  upper  atmosphere  spirals  toward  the  earth.  Within 
a  very  short  time,  the  energy-radiating  electron  should  actually 
be  pulled  into  the  nucleus.  According  to  classical  physics  — 
mechanics  and  electromagnetism  — a  planetary  atom  would  not  be 
stable  for  more  than  a  very  small  fraction  of  a  second. 

The  idea  of  a  planetary  atom  was  nevertheless  sufficiently 
appealing  that  physicists  continued  to  look  for  a  theoi^y  that  would 
include  a  stable  planetary  structure  and  predict  discrete  line  spectra 
for  the  elements.  Niels  Bohr,  an  unknown  young  Danish  physicist 
who  had  just  received  his  PhD  degree,  succeeded  in  constructing 
such  a  theory  in  1912-1913.  This  theory,  although  it  had  to  be 
modified  later  to  make  it  applicable  to  many  more  phenomena,  was 
widely  recognized  as  a  major  victory,  showing  how  to  attack 
atomic  problems  by  using  quantum  theory.  In  fact,  even  though  it 
is  now  a  comparatively  naive  way  of  thinking  about  the  atom 
compared  to  the  view  given  by  more  recent  quantum-mechanical 


72 


The  Rutherford-Bohr  Model  of  the  Atom 


Since  Bohr  incorporated  Ruther- 
ford's idea  of  the  nucleus,  the  model 
which  Bohr's  theory  discusses  is 
often  called  the  Rutherford-Bohr 

model. 


E.  state-. 


\ 


\        -  - '-        / 

\ 


eiriiSSiOn; 


/ 


\^3^^ 


]\^ 


-£  sf<j+e; 


y 


theories,  Bohr's  theory  is  a  beautiful  example  of  a  successful 
physical  model,  measured  by  what  it  was  designed  to  do. 

Bohr  introduced  two  novel  postulates  designed  specifically  to 
account  for  the  existence  of  stable  electron  orbits  and  of  the  discrete 
emission  spectra.  These  postulates  may  be  stated  as  follows. 

(1)  Contrary  to  the  expectations  based  on  classical  mechanics 
and  electromagnetism,  an  atomic  system  can  exist  in  any  one  of  a 
number  of  states  in  which  no  emission  of  radiation  takes  place, 
even  if  the  particles  (electrons  and  nucleus)  are  in  motion  relative 
to  each  other.  These  states  are  called  stationary  states  of  the  atom. 

(2)  Any  emission  or  absorption  of  radiation,  either  as  visible 
light  or  other  electromagnetic  radiation,  will  correspond  to  a  sudden, 
discontinuous  transition  between  two  such  stationary  states.  The 
radiation  emitted  or  absorbed  in  a  transition  has  a  frequency  / 
determined  by  the  relation  hf—  E,  —  E/,  where  h  is  Planck's  constant 
and  Ei  and  Ef  are  the  energies  of  the  atom  in  the  initial  and  final 
stationary  states,  respectively. 

The  quantum  theory  had  begun  with  Planck's  idea  that  atoms 
emit  light  only  in  definite  amounts  of  energy;  it  was  extended  by 
Einstein's  idea  that  light  travels  only  as  definite  parcels  of  energy; 
and  now  it  was  extended  further  by  Bohr's  idea  that  atoms  exist 
only  in  definite  energy  states.  But  Bohr  also  used  the  quantum 
concept  in  deciding  which  of  all  the  conceivable  stationary  states 
of  the  atom  were  actually  possible.  An  example  of  how  Bohr  did 
this  is  given  in  the  next  section. 

For  simplicity  we  consider  the  hydrogen  atom,  with  a  single 
electron  revolving  around  the  nucleus.  Following  Bohr,  we  assume 
that  the  possible  electron  orbits  are  simply  circular.  The  details  of 
some  additional  assumptions  and  the  calculation  are  worked  out 
on  page  73.  Bohr's  result  for  the  possible  orbit  radii  r„  was  r„  =  an^ 
where  a  is  a  constant  {h^l4'jT'^mkqe)  that  can  be  calculated  from 
known  physical  values,  and  n  stands  for  any  whole  number,  1, 
2,  3 

Q1 1     What  was  the  main  evidence  that  an  atom  could  exist 
only  in  certain  energy  states? 

Q12     What  reason  did  Bohr  give  for  the  atom  existing  only  in 
certain  energy  states? 


19.6    The  size  of  the  hydrogen  atom 

This  is  a  remarkable  result:  in  the  hydrogen  atom,  the  allowed 
orbital  radii  of  the  electrons  are  whole  multiples  of  a  constant  that 
we  can  at  once  evaluate.  That  is  n^  takes  on  values  of  V,  2^  3^,  .  .  .  , 
and  all  factors  to  the  right  of  n'^  are  quantities  known  previously  by 
independent  measurement!  Calculating  the  value  (h'^l4Tr-mkq^)  gives 
us  5.3  X  10~"m.  Hence  we  now  know  that  according  to  Bohr's 
model  the  radii  of  stable  electron  orbits  should  be  r„  =  5.3  x  10~"m 
X  n\  That  is,  5.3  x  10-"m  when  n  =  1  (first  allowed  orbit),  4  x  5.3  = 
10~"m  when  n  =  2  (second  allowed  orbit),  9  x  5.3  x  10~"m  when 
n  —  3,  etc.  In  between  these  values,  there  are  no  allowed  radii.  In 


Bohr's  Quantization  Rule  and  the  Size  of  Orbits 


The  magnitude  of  the  charge  on  the  electron 
is  Qf.;  the  charge  on  a  nucleus  is  Zq,.,  and  for 
hydrogen  (Z  =  1)  is  just  q^.  The  electric  force 
with  which  the  hydrogen  nucleus  attracts  its 
electron  is  therefore 


Fel 


QeQe 


where  k  is  the  coulomb  constant,  and  r  is  the 
center-to-center  distance.  If  the  electron  is  in  a 
stable  circular  orbit  of  radius  r  around  the 
nucleus,  moving  at  a  constant  speed  v,  then 
the  centripetal  force  is  equal  to  mv^/r.  Since 
the  centripetal  force  is  the  electric  attraction, 
we  can  write 


mv'  _     q' 


In  the  last  equation,  m,  q^  and  k  are 
constants;  r  and  v  are  variables,  whose  values 
are  related  by  the  equation.  What  are  the 
possible  values  of  \/  and  r  for  stationary  states 
of  the  atom? 

We  can  begin  to  get  an  answer  if  we  write 
the  last  equation  in  slightly  different  form,  by 
multiplying  both  sides  by  r^  and  dividing  both 
sides  by  v\  the  result  is 

mvr  =  — ^ 

V 

The  quantity  on  the  left  side  of  this  equa- 
tion, which  is  the  product  of  the  momentum  of 
the  electron  and  the  radius  of  the  orbit,  can  be 
used  to  characterize  the  stable  orbits.  According 
to  classical  mechanics,  the  radius  of  the  orbit 
could  have  any  value,  so  the  quantity  mvr  could 
also  have  any  value.  But  we  have  seen  that 
classical  physics  seemed  to  deny  that  there 
could  be  any  stable  orbits  in  the  hydrogen 
atom.  Since  Bohr's  first  postulate  implies  that 
certain  stable  orbits  (and  only  those)  are 
permitted,  Bohr  needed  to  find  the  rule  that 
decides  which  stable  orbits  were  possible.  Here 
Bohr  appears  to  have  been  largely  guided  by 
his  intuition.  He  found  that  what  was  needed 
was  the  recognition  that  the  quantity  m\//'  does 
not  take  on  any  arbitrary  value,  but  only  certain 


discrete  values.  These  values  are  defined  by  the 
relation 


mvr 


2v 


where  h  is  Planck's  constant,  and  n  is  a  posi- 
tive integer;  that  is,  n  =  1,  2,  3,  4,  .  .  .  (but  not 
zero).  When  the  possible  values  of  the  mvr  are 
restricted  in  this  way,  the  quantity  mvr  is  said 
to  be  quantized.  The  integer  n  which  appears 
in  the  formula,  is  called  the  quantum  number. 
The  main  point  is  that  each  quantum  number 
(n  =  1  or  2  or  3  .  .  .)  corresponds  to  one 
allowed,  stable  orbit  of  the  electron. 

If  we  accept  this  rule,  we  can  at  once 
describe  the  "allowed"  states  of  the  atom,  say 
in  terms  of  the  radii  r  of  the  possible  orbits. 
We  can  combine  the  last  expression  above 
with  the  classical  centripetal  force  relation  as 
follows:  the  quantization  rule  is 

nh 


mi/r  =  — - 
2v 

nh 

so 

Inmv 

and 

,,_     n'h^ 

4v'-m-v^ 
From  classical  mechanics,  we  had 


mv  _  1^  q'e 


so 


mr 


Substituting  this  "classical"  value  for  v-  into  the 
quantization  expression  for  r-  gives 


r- 


n'h' 


47^/77-1 

mr 


which  simplifies  to  the  expression  for  the 
allowed  radii,  r„: 

n'h' 


r  = 


A-TT-kmq^ 


74  The  Rutherford-Bohr  Model  of  the  Atom 

SG  19.10  short,  we  have  found  that  the  separate  allowed  electron  orbits  are 

spaced  around  the  nucleus  in  a  regular  way,  with  the  allowed  radii 
quantized  in  a  regular  manner,  as  indicated  in  the  marginal 
drawing.  Emission  and  absorption  of  light  should  then  be  accom- 
panied by  the  transition  of  the  electron  from  one  allowed  orbit  to 
another. 

This  is  just  the  kind  of  result  we  had  hoped  for;  it  tells  us 
which  radii  are  possible,  and  where  they  lie.  But  so  far,  it  has  all 
been  model  building.  Do  the  orbits  in  a  real  hydrogen  atom  actually 
correspond  to  this  model?  In  his  first  paper  of  1913,  Bohr  could 
give  at  least  a  partial  yes  as  answer:  It  was  long  known  that  the 
normal  "unexcited"  hydrogen  atom  has  a  radius  of  about  5  x  10~"  m. 
(That  is,  for  example,  the  size  of  the  atom  obtained  by  interpreting 
measured  characteristics  of  gases  in  the  light  of  the  kinetic  theory.) 
This  known  value  of  5  x  10~"  m  corresponds  excellently  to  the 
prediction  from  the  equation  for  the  orbital  radius  r  if  n  has  the 
lower  value,  namely  1.  For  the  first  time  there  was  now  a  way  to 
understand  the  size  of  the  neutral,  unexcited  hydrogen  atom:  for 
every  atom  the  size  corresponds  to  the  size  of  the  innermost  allowed 
electron  orbit,  and  that  is  fixed  by  nature  as  described  by  the 
quantization  rule. 


Q13    Why  do  all  unexcited  hydrogen  atoms  have  the  same  size? 
Q14    Why  does  the  hydrogen  atom  have  just  the  size  it  has? 


19.7    Other  consequences  of  the  Bohr  model 

With  his  two  postulates  and  his  choise  of  the  permitted 
stationary  states,  Bohr  could  calculate  not  only  the  radius  of  each 
permitted  orbit,  but  also  the  total  energy  of  the  electron  in  each 
orbit;  this  energy  is  the  energy  of  the  stationary  state. 

The  results  that  Bohr  obtained  may  be  summarized  in  two 
simple  formulas.  As  we  saw,  the  radius  of  an  orbit  with  quantum 
number  n  is  given  by  the  expression 


where  r,  is  the  radius  of  the  first  orbit  (the  orbit  for  n  =  1)  and 
has  the  value  5.3  x  IQ-**  cm  or  5.3  x  lO""  m. 

The  energy  (including  both  kinetic  and  electric  potential 
energy)  of  the  electron  in  the  orbit  with  quantum  number  n  can  be 
computed  from  Bohr's  postulate  also  (see  SG  19.11).  As  we  pointed 
out  in  Chapter  10,  it  makes  no  sense  to  assign  an  absolute  value  to 
potential  energy  — since  only  changes  in  energy  have  physical 
meaning  we  can  pick  any  convenient  zero  level.  For  an  electron 
orbiting  in  an  electric  field,  the  mathematics  is  particularly  simple 

Note:  Do  not  confuse  this  use  of  £  '^  ^^  ^^°°^^  ^^  ^  ^^^°  ^^^^^  ^°^  ^"^^^8^  ^^^  ^*^^^  ^  =  "'  ^^^^  ^^' 

for  energy  with  earlier  use  of  Efor  when  the  electron  is  infinitely  far  from  the  nucleus  (and  therefore 

electric  field.  free  of  it).  If  we  consider  the  energy  for  any  other  state  E„  to  be 


Section  19.8  75 

the  difference  from  this  free  state,  we  can  write  the  possible 
energy  states  for  the  hydrogen  atom  as 

where  Ej  is  the  total  energy  of  the  atom  when  the  electron  is  in  the 
first  orbit;  Ei,  the  lowest  energy  possible  for  an  electron  in  a 
hydrogen  atom,  is  —13.6  eV  (the  negative  value  means  only  that  the 
energy  is  13.6  eV  less  than  the  free  state  value  Ex).  This  is  called 
the  "ground"  state.  In  that  state,  the  electron  is  most  tightly 
"bound"  to  the  nucleus.  The  value  of  E,,  the  first  "excited"  state 
above  the  ground  state,  is  1/2^  x  -13.6  eV  =  -3.4  eV,  that  is,  only 
3.4  eV  less  than  in  the  free  state. 

According  to  the  formula  for  r„,  the  first  Bohr  orbit  has  the 
smallest  radius,  with  n  =  1.  Higher  values  of  n  correspond  to 
orbits  that  have  larger  radii.  Although  the  higher  orbits  are  spaced 
further  and  further  apart,  the  force  field  of  the  nucleus  falls  off 
rapidly,  so  the  work  required  to  move  out  to  the  next  larger  orbit 
actually  becomes  smaller  and  smaller;  therefore  also  the  jumps  in 
energy  from  one  level  of  allowed  energy  E  to  the  next  become  small 
and  smaller. 


19.8    The  Bohr  theory:  the  spectral  series  of  hydrogen 

It  is  commonly  agreed  that  the  most  spectacular  success  of  «       .         ..  . 

See  the  radius  and  energy  diagrams 
Bohr's  model  was  that  it  could  be  used  to  explain  all  emission  (and  ^^  -g^g  jq 

absorption  lines  in  the  hydrogen  spectrum.  That  is,  Bohr  could  use 

his  model  to  derive,  and  so  to  explain,  the  B aimer  formula!  By 

applying  his  second  postulate,  we  know  that  the  radiation  emitted 

or  absorbed  in  a  transition  in  Bohr's  atom  should  have  a  frequency 

/  determined  by  the  relation 

hf  =  E,.  -  E, 

If  Uf  is  the  quantum  number  of  the  final  state,  and  n,  is  the 
quantum  number  of  the  initial  state,  then  according  to  the  result 
for  E„  we  know  that 

E^  =  ^Ei        and        Ei^^—^E, 

The  frequency  of  radiation  emitted  or  absorbed  when  the  atom  goes 
from  the  initial  state  to  the  final  state  is  therefore  determined  by 
the  equation 

hf-^.-h        or        hf=E,{-\-\ 
n,-      Uf-  \7Vi     n^f. 

To  deal  with  wavelength  A.  (as  in  Balmer's  original  formula,  p.  63) 
rather  than  frequency  /,  we  use  now  the  relation  between  fre- 
quency and  wavelength  given  in  Unit  3:  the  frequency  is  equal  to 


Niels  Bohr  (1885-1962)  was  born  in  Copenhagen, 
Denmark  and  was  educated  there,  receiving  his 
doctor's  degree  in  physics  in  191 1.  In  1912  he  was 
at  work  in  Rutherford's  laboratory  in  Manchester, 
England,  which  was  a  center  of  research  on  radio- 
activity and  atomic  structure.  There  he  developed 
his  theory  of  atomic  structure  to  explain  chemical 
properties  and  atomic  spectra.  Bohr  later  played  an 
important  part  in  the  development  of  quantum 
mechanics,  in  the  advancement  of  nuclear  physics, 
and  in  the  study  of  the  philosophical  aspects  of 
modern  physics.  In  his  later  years  he  devoted  much 
time  to  promoting  plans  for  international  coopera- 
tion and  the  peaceful  uses  of  nuclear  physics. 


Section  19.8  77 

the  speed  of  the  hght  wave  divided  by  its  wavelength:  /=  c/X.  If 
we  substitute  c/X  for /in  this  equation,  and  then  divide  both  sides 
by  the  constant  he  (Planck's  constant  times  the  speed  of  light),  we 
obtain  the  equation 


i^£i/J L 

X      he  \n^i     n^f. 


According  to  Bohr's  model,  then,  this  equation  gives  the  wave- 
length X  of  the  radiation  that  will  be  emitted  or  absorbed  when  the 
state  of  a  hydrogen  atom  changes  from  one  stationary  state  with 
quantum  number  n,  to  another  with  Uf. 

How  does  this  prediction  from  Bohr's  model  compare  with  the 
empirical  Balmer  formula  for  the  Balmer  series?  The  Balmer 
formula  was  given  on  page  64: 

i=R      (1-1 

We  see  at  once  that  the  equation  for  X  of  emitted  (or  absorbed) 
light  derived  from  the  Bohr  model  is  exactly  the  same  as  B aimer's 
formula,  if  Rff  =  —EJhc  and  nf=  2. 

The  Rydberg  constant  R^,  long  known  from  spectroscopic 
measurements  to  have  the  value  of  1.097  x  10^m~S  now  could 
be  compared  with  the  value  for  —(EJhc).  Remarkably,  there  SG  19.11 

was  fine  agreement.  R^,  which  had  previously  been  regarded  as 
just  an  experimentally  determined  constant,  was  now  shown 
not  to  be  arbitrary  or  accidental,  but  to  depend  on  the  mass  and 
charge  of  the  electron,  on  Planck's  constant,  and  on  the  speed 
of  hght. 

More  important,  one  now  saw  the  meaning,  in  physical  terms, 
of  the  old  empirical  formula  for  the  Balmer  series.  All  the  lines  in 
the  Balmer  series  simply  correspond  to  transitions  from  various 
initial  states  (various  values  of  n,)  to  the  same  final  state,  the  state 
for  which  nf  =  2. 

When  the  Bohr  theory  was  proposed,  in  1913,  emission  lines  in 
only  the  Balmer  and  Paschen  series  for  hydrogen  were  known 
definitely.  Balmer  had  suggested,  and  the  Bohr  model  agreed,  that 
additional  series  should  exist.  The  experimental  search  for  these 
series  yielded  the  discovery  of  the  Lyman  series  in  the  ultraviolet 
portion  of  the  spectrum  (1916),  the  Brackett  series  (1922),  and  the 
Pfund  series  (1924).  In  each  series  the  measured  frequencies  of  the 
lines  were  found  to  be  those  predicted  by  Bohr's  theory.  Similarly, 
the  general  formula  that  Balmer  guessed  might  apply  for  all  spec- 
tral lines  of  hydrogen  is  explained;  lines  of  the  Lyman  series 
correspond  to  transitions  from  various  initial  states  to  the  final 
state  n^==  1,  the  lines  of  the  Paschen  series  correspond  to  transitions 
from  various  initial  states  to  the  final  state  Uf  =  3,  etc.  (see  table  SG  19.12,  19.13 

on  page  65).  The  general  scheme  of  possible  transitions  among  the 


78 


The  Rutherford-Bohr  Model  of  the  Atom 


n  =  6 


Above:  A  schematic  diagram  of  the 
possible  transitions  of  an  electron 
in  the  Bohr  model  of  the  hydrogen 
atom  (first  six  orbits). 
At  the  right:  Energy-level  diagram  for 
the  hydrogen  atom.  Possible  transi- 
tions between  energy  states  are  shown 
for  the  first  six  levels.  The  dotted  arrow 
for  each  series  indicates  the  series 
limit,  a  transition  from  the  state  where 
the  electron  is  completely  free  (in- 
finitely far)  from  the  nucleus. 


first  six  stable  orbits  is  shown  in  the  figure  at  the  left.  Thus 
the  theory  not  only  correlated  currently  known  information 
about  the  spectrum  of  hydrogen,  but  also  predicted 
correctly    the    wavelength   of  hitherto   unknown 
series  of  lines  in  the  spectrum.  Moreover,  it  did 
so  on  a  physically  plausible  model  rather  than, 
as  Balmer's  general  formula  had  done,  with 
out  any  physical  reason.  All  in  all,  these 
were  indeed  triumphs  that  are  worth  cele- 
brating! 
The  schematic  diagram  shown  at  the  left 
is  useful  as  an  aid  for  the  imagination, 
but  it  also  has  the  danger  of  being  too 
specific.  For  instance,  it  leads  us  to  visual- 
ize the  emission  of  radiation  in  terms  of 
"jumps"  of  electrons  between  orbits.  These 
are  useful  ideas  to  aid  our  thinking,  but  one 
must  not  forget  that  we  cannot  actually  de- 
tect an  electron  moving  in  an  orbit,  nor  can  we 
watch  an  electron  "jump"  from  one  orbit  to  an- 
other. Hence  a  second  way  of  presenting  the  results 
'"'         of  Bohr's  theory  is  used  which  yields  the  same  facts 
,.--^''      but  does  not  commit  us  too  closely  to  a  picture  of  orbits. 
This  scheme  is  shown  in  the  figure  below.  It  focuses  attention 
not  on  orbits  but  on  the  corresponding  possible  energy  states,  which 

ENERGY 


Lyman 
series 


Balmer 
series 


n  =  5 
n  =  4 

n  =  3 


n  =  2 


t      t 


I 


Paschen 

series 

1 

It 


Brackett 
series 


'ii 


Pfund 
series 

...       X   - 


\i 


iiit  i 


0.0 

-0.87  X  10- 

-1.36 

-2.42 


-5.43 


n=i  JjiiH 


■21.76 


Section  19.9 


79 


are  all  given  by  the  formula  E„  =  1/n^  x  £,.  In  terms  of  this 
mathematical  m.odel,  the  atom  is  normally  unexcited,  with  an 
energy  £,  about  -22  x  10~'^  joules  (-13.6  eV).  Absorption  of  energy 
can  place  the  atoms  in  an  excited  state,  with  a  correspondingly 
higher  energy.  The  excited  atom  is  then  ready  to  emit  light,  with 
a  consequent  reduction  in  energy.  The  energy  absorbed  or  emitted 
always  shifts  the  total  energy  of  the  atom  to  one  of  the  values 
specified  by  the  formula  for  E„.  We  may  thus,  if  we  prefer,  represent 
the  hydrogen  atom  by  means  of  the  energy-level  diagram. 

Q15     Balmer  had  predicted  accurately  the  other  spectral  series 
of  hydrogen  thirty  years  before  Bohr  did.  Why  is  Bohr's  prediction 
considered  more  significant? 

Q16    How  does  Bohr's  model  explain  line  absorption  spectra? 


19.9    Stationary  states  of  atoms:  the  Franck-Hertz  experiment 

The  success  of  the  Bohr  theory  in  accounting  for  the  spectrum 
of  hydrogen  leaves  this  question:  can  experiments  show  directly 
that  atoms  have  only  certain  discrete  energy  states?  In  other  words, 
apart  from  the  success  of  the  idea  in  explaining  spectra,  are  there 
really  gaps  between  the  energies  that  an  atom  can  have?  A  famous 
experiment  in  1914,  by  the  German  physicists  James  Franck  and 
Gustav  Hertz,  showed  the  existence  of  these  discrete  energy  states. 

Franck  and  Hertz  bombarded  atoms  with  electrons  (from  an 
electron  gun)  and  were  able  to  measure  the  energy  lost  by  electrons 
in  collisions  with  atoms.  They  could  also  determine  the  energy 
gained  by  atoms  in  these  collisions.  In  their  first  experiment,  Franck 
and  Hertz  bombarded  mercury  vapor  contained  in  a  chamber  at 
very  low  pressure.  Their  experimental  procedure  was  equivalent  to 
measuring  the  kinetic  energy  of  electrons  leaving  the  electron  gun 
and  the  kinetic  energy  of  electrons  after  they  had  passed  through 
the  mercury  vapor.  The  only  way  electrons  could  lose  energy  was  in 
collisions  with  mercury  atoms.  Franck  and  Hertz  found  that  when 
the  kinetic  energy  of  the  electrons  leaving  the  electron  gun  was 
small,  for  example,  up  to  several  eV,  the  electrons  after  passage 
through  the  mercury  vapor  still  had  almost  exactly  the  same  energy 
as  they  had  on  leaving  the  gun.  This  result  could  be  explained  in 
the  following  way.  A  mercury  atom  is  several  hundred  thousand 
times  more  massive  than  an  electron.  When  it  has  low  kinetic 
energy  the  electron  just  bounces  off  a  mercury  atom,  much  as  a 
golf  ball  thrown  at  a  bowling  ball  would  bounce  off.  A  collision  of 
this  kind  is  called  an  "elastic"  collision.  In  an  elastic  collision, 
the  mercury  atom  (bowling  ball)  takes  up  only  an  extremely  small 
part  of  the  kinetic  energy  of  the  electron  (golf  ball).  The  electron 
loses  practically  none  of  its  kinetic  energy. 

But  when  the  kinetic  energy  of  the  bombarding  electrons  was 
raised  to  5  electron-volts,  there  was  a  dramatic  change  in  the 
experimental  results.  When  an  electron  collided  with  a  mercury 


'4.oeV 


4.o&f  O 

BcSCTROfJ 


M6RCUR.y  AfVH 


0.1  eV  t^^j£f^i 


'BlecrOm 


HeUCOBJATDH 


/^eeoisy  Aran 


The  Nobel  Prize 


Alfred  Bernhard  Nobel  (1833-1896),  a  Swed- 
ish chemist,  was  the  inventor  of  dynamite. 
As  a  result  of  his  studies  of  explosives, 
Nobel  found  that  when  nitroglycerine  (an 
extremely  unstable  chemical)  was  absorbed 
in  an  inert  substance  it  could  be  used 
safely  as  an  explosive.  This  combination 
is  dynamite.  He  also  invented  other  ex- 
plosives (blasting  gelatin  and  ballistite) 
and  detonators.  Nobel  was  primarily  inter- 
ested in  the  peaceful  uses  of  explosives, 
such  as  mining,  road  building  and  tunnel 
blasting,  and  he  amassed  a  large  fortune 
from  the  manufacture  of  explosives  for 
these  applications.  Nobel  abhorred  war  and 
was  conscience-stricken  by  the  military 
uses  to  which  his  explosives  were  put.  At 


his  death,  he  left  a  fund  of  some  $315  mil- 
lion to  honor  important  accomplishments  in 
science,  literature  and  international  under- 
standing. Prizes  were  established  to  be 
awarded  each  year  to  persons  who  have 
made  notable  contributions  in  the  fields  of 
physics,  chemistry,  medicine  or  physiology, 
literature  or  peace.  (Since  1969  there  has 
been  a  Nobel  Memorial  Prize  in  economics 
as  well.)  The  first  Nobel  Prizes  were  awarded 
in  1901.  Since  then,  men  and  women  from 
about  30  countries  have  received  prizes. 
At  the  award  ceremonies  the  recipient  re- 
ceives a  medal  and  the  prize  money  from  the 
king  of  Sweden,  and  is  expected  to  deliver 
a  lecture  on  his  work.  The  Nobel  Prize  is 
generally  considered  the  most  prestigious 
prize  in  science. 


I 


Nobel  Prize  winners  in  Physics. 


1901  Wilhelm  Rontgen  (Ger)  — discovery  of  x-rays.  1938 

1902  H.  A.   Lorentz  and  P.  Zeeman  (Neth)-influence  of 
magnetism  on  radiation. 

1903  A.  H.  Becquerel  (Fr)- discovery  of  spontaneous  radio-  1939 
activity.  Pierre  and   Marie  Curie  (Fr)  — work  on  rays 

first  discovered  by  Becquerel.  1940 

1904  Lord  Rayieigh  (Gr  Brit)-density  of  gases  and  dis-  1941 
covery  of  argon.  1942 

1905  Philipp  Lenard  (Ger) -work  on  cathode  rays.  1943 

1906  J.   J.   Thomson   (Gr  Brit)-conduction  of  electricity 

by  gases.  1944 

1907  Albert  A.  Michelson  (US)  — optical  precision   instru- 
ments and  spectroscopic  and  metrological  investi-         1945 
gations.  1946 

1908  Gabriel  Lippmann  (Fr)-color  photography  by  1947 
interference. 

1909  Guglielmo    Marconi    (Ital)-and    Ferdinand    Braum 

(Ger)  — development  of  wireless  telegraphy.  1948 

1910  Johannes  van  der  Waals  (Neth)-equation  of  state 
for  gases  and  liquids. 

1911  Wilhelm  Wien  (Ger)-laws  governing  the  radiation  1949 
of  heat. 

1912  Nils  Gustaf  Dalen  (Swed) -automatic  gas  regulators  1950 
for  lighthouses  and  buoys. 

1913  Kamerlingh    Onnes    (Neth)-low    temperature    and 
production  of  liquid  helium.  1951 

1914  Max   von    Laue   (Ger) -diffraction   of   Rontgen    rays 
by  crystals. 

1915  W.  H.  and  W.  L.  Bragg  (Gr  Brit)-analysis  of  crystal  1952 
structure  by  Rontgen  rays. 

1916  No  award.  1953 

1917  Charles  Glover  Barkla(GrBrit)-discovery  of  Rontgen  1954 
radiation  of  the  elements. 

1918  Max  Planck  (Ger)  — discovery  of  energy  quanta. 

1919  Johannes  Stark  (Ger) -discovery  of  the  Doppler  1955 
effect  in  canal  rays  and  the  splitting  of  spectral  lines 

in  electric  fields. 

1920  Charles-Edouard  Guillaume  (Switz)- discovery  of  1956 
anomalies  in  nickel  steel  alloys. 

1921  Albert  Einstein  (Ger)-for  contributions  to  theoretical 
physics  and  especially  for  his  discovery  of  the  law         1957 
of  the  photoelectric  effect. 

1922  Niels   Bohr   (Den)  — atomic   structure   and   radiation. 

1923  Robert  Andrews  Millikan  (US)-elementary  charge  1958 
of  electricity  and  photoelectric  effect. 

1924  Karl  Siegbahn  (Swed) -field  of  x-ray  spectroscopy. 

1925  James  Franck  and  Gustav  Hertz  (Ger)  — laws  govern-  1959 
ing  the  impact  of  an  electron  upon  an  atom. 

1926  Jean  Baptiste  Perrin  (Fr)-discontinuous  structure  1960 
of  matter  and  especially  for  his  discovery  of  sedi-  1961 
mentation  equilibrium. 

1927  Arthur   Compton    (US) -discovery    of   effect   named 
after  him.  C.  T.  R.  Wilson  (Gr  Brit)  -  method  of  making 

paths  of  electrically  charged  particles  visible  by  con-  1962 

densation  of  vapor. 

1928  Owen  Williams  Richardson  (Gr  Brit)-thermionic  1963 
phenomena  and  discovery  of  effect  named  after  him. 

1929  Louis-Victor    de    Broglie    (Fr)- discovery    of    wave 

nature  of  electrons.  1964 

1930  Sir   Chandrasehara   V.    Raman    (Ind)-scattering    of 

light  and  effect  named  after  him.  1965 

1931  No  award. 

1932  Werner  Heisenberg  (Ger)  — quantum  mechanics  lead- 
ing to  discovery  of  allotropic  forms  of  hydrogen.  1966 

1933  Erwin  Schrodinger  (Ger)  and  P.  A.  M.  Dirac  (Gr  Brit)- 

new  productive  forms  of  atomic  theory.  1967 

1934  No  award. 

1935  James  Chadwick  (Gr  Brit)  — discovery  of  the  neutron.  1968 

1936  Victor  Franz  Hess  (Aus.  — cosmic  radiation.  Carl  David 
Anderson  (US)  — discovery  of  the  positron. 

1937  Clinton  J.  Davisson  (US) -and  George  P.  Thomson  1969 
(Gr   Brit)  — experimental   diffraction   of  electrons   by 
crystals. 


Enrico  Fermi  (Ital)  — new  radioactive  elements  pro- 
duced by  neutron  irradiation  and  nuclear  reactions 
by  slow  neutrons. 

Ernest  O.   Lawrence  (US) -cyclotron  and  its  use  in 
regard  to  artificial  radioactive  elements. 
No  award 
No  award 
No  award 

Otto  Stern  (Ger) -molecular  ray  method  and  magnetic 
moment  of  the  proton. 

Isidor  Isaac  Rabi  (US) -resonance  method  for  mag- 
netic properties  of  atomic  nuclei. 
Wolfgang  Pauli  (Aus)  — exclusion  or  Pauli  principle. 
P.  W.  Bridgman  (US)  — high  pressure  physics. 
Sir  Edward  V.  Appleton  (Gr  Brit)  — physics  of  the  upper 
atmosphere  and  discovery  of  so-called  Appleton 
layers. 

Patrick  M.  S.  Blackett,  (Gr  Brit) -development  of 
Wilson  cloud  chamber  and  discoveries  in  nuclear 
physics  and  cosmic  rays. 

Hideki  Yukawa  (Japan)  — prediction  of  mesons  and 
theory  of  nuclear  forces. 

Cecil  Frank  Powell  (Gr  Brit)  — Photographic  method  of 
studying  nuclear  processes  and  discoveries  regarding 
mesons. 

Sir  John  D.  Cockcroft  and  Ernest  T.  S.  Walton  (Gr 
Brit) -transmutation  of  atomic  nuclei  by  artificially 
accelerated  atomic  particles. 

Felix   Bloch   (Switz)   and   Edward   M.  Purcell   (US)- 
nuclear  magnetic  precision  measurements. 
Frits  Zernike  (Neth)- phase-contrast  microscope. 
Max   Born    (Ger) -statistical    interpretation   of  wave 
functions,    and    Walter    Bothe    (Ger)-coincidence 
method  for  nuclear  reactions  and  cosmic  rays. 
Willis  E.  Lamb  (US) -fine  structure  of  hydrogen  spec- 
trum and  Polykarp  Kusch  (US) -precision  determina- 
tions of  magnetic  moment  of  electron. 
William  Shockley,  John  Bardeen  and  Walter  Houser 
Brattain   (US) -researches   on   semiconductors   and 
their  discovery  of  the  transistor  effects. 
Chen  Ning  Yang  and  Tsung  Dao  Lee  (Chin)-investi- 
gation  of  laws  of  parity,  leading  to  discoveries  regard- 
ing the  elementary  particles. 

Pavel  A.  Cerenkov,  H'ya  M.  Frank  and  Igor  E.  Tamm 
(USSR) -discovery  and  interpretation  of  the  Cerenkov 
effect. 

Emilio  G.  Segre  and  Owen  Chamberlain  (US)-dis- 
covery  of  the  antiproton. 

Donald  A.  Giaser  (US)-invention  of  bubble  chamber. 
Robert  Hofstadter  (US) -electron  scattering  in  atomic 
nuclei.  Rudolf  Ludwig  Mossbauer  (Ger) -resonance 
absorption  of  -y-radiation  and  discovery  of  effect 
which  bears  his  name. 

Lev  D.  Landau  (USSR)-theories  for  condensed  mat- 
ter, especially  liquid  helium. 

Eugene  P.  Wigner  (US)-theory  of  the  atomic  nucleus 
and  elementary  particles.  Marie  Goeppert-Mayer  (US) 
and  J.  Hans  D.  Jensen  (Ger) -nuclear  shell  structure. 
Charles  Townes  (US),  Alexander  Prokhorov  and 
Nikolay  Basov  (USSR) -development  of  maser. 
S.  Tomonaga  (Japan),  Julian  Schwinger  and  Richard 
Feynman  (US)-quantum  electrodynamics  and  ele- 
mentary particles. 

Alfred  Kastler  (Fr)-new  optical  methods  for  studying 
properties  of  atom. 

Hans  Bethe  (US) -nuclear  physics  and  theory  of 
energy  production  in  the  sun. 

Louis  W.  Alvarez  (American)  for  research  in  physics 
of  sub  atomic  particles  and  techniques  for  detection 
of  these  particles. 

Murray  Gell-Mann  (American)  for  contributions  and 
discoveries  concerning  the  classification  of  elemen- 
tary particles  and  their  interactions. 


82 


The  Rutherford-Bohr  Model  of  the  Atom 


We  now  know  two  ways  of  "exciting" 
an  atom:  by  absorption  of  a  photon 
with  just  the  right  energy  to  make 
a  transition  from  the  lowest  energy 
level  to  a  higher  one,  or  by  doing 
the  same  thing  by  collision— with  an 
electron  from  an  electron  gun,  or  by 
collision  among  agitated  atoms  (as 
in  a  heated  enclosure  or  a  discharge 
tube). 


SG  19.14,  19.15 


SG  19.16 


atom  it  lost  almost  exactly  4.9  eV  of  energy.  And  when  the  electron 
energy  was  increased  to  6  eV,  the  electron  still  lost  just  4.9  eV  of 
energy  in  a  collision  with  a  mercury  atom,  being  left  with  1.1  eV  of 
energy.  These  results  indicated  that  a  mercury  atom  cannot  accept 
less  than  4.9  eV  of  energy;  and  that  when  it  is  offered  somewhat 
more,  for  example,  5  or  6  eV,  it  still  can  accept  only  4.9  eV.  The 
accepted  amount  of  energy  cannot  go  into  kinetic  energy  of  the 
mercury  because  of  the  relatively  enormous  mass  of  the  atom  as 
compared  with  that  of  an  electron.  Hence,  Franck  and  Hertz  con- 
cluded that  the  4.9  eV  of  energy  is  added  to  the  internal  energy 
of  the  mercury  atom  — that  the  mercury  atom  has  a  stationary  state 
with  energy  4.9  eV  greater  than  that  of  the  lowest  energy  state, 
with  no  allowed  energy  level  in  between. 

What  happens  to  this  extra  4.9  eV  of  internal  energy?  According 
to  the  Bohr  model  of  atoms,  this  amount  of  energy  should  be 
emitted  in  the  form  of  electromagnetic  radiation  when  the  atom 
returns  to  its  lowest  state.  Franck  and  Hertz  looked  for  this  radia- 
tion, and  found  it.  They  observed  that  the  mercury  vapor  emitted 
light  at  a  wavelength  of  2535  A,  a  line  known  previously  to  exist 
in  the  emission  spectrum  of  hot  mercury  vapor.  The  wavelength 
corresponds  to  a  frequency  /  for  which  the  photon's  energy,  hf, 
is  just  4.9  eV  (as  you  can  calculate).  This  result  showed  that 
mercury  atoms  had  indeed  gained  (and  then  radiated)  4.9  eV  of 
energy  in  collisions  with  electrons. 

Later  experiments  showed  that  mercury  atoms  bombarded  by 
electrons  could  also  gain  other,  sharply  defined  amounts  of  energy, 
for  example,  6.7  eV  and  10.4  eV.  In  each  case  radiation  was  emitted 
that  corresponded  to  known  lines  in  the  emission  spectrum  of 
mercury;  in  each  case  analogous  results  were  obtained.  The  elec- 
trons always  lost  energy,  and  the  atoms  always  gained  energy,  both 
in  sharply  defined  amounts.  Each  type  of  atom  studied  was  found  to 
have  discrete  energy  states.  The  amounts  of  energy  gained  by  the 
atoms  in  collisions  with  electrons  could  always  be  correlated  with 
known  spectrum  lines.  The  existence  of  discrete  or  stationary 
states  of  atoms  predicted  by  the  Bohr  theory  of  atomic  spectra  was 
thus  verified  by  direct  experiment.  This  verification  was  considered 
to  provide  strong  confirmation  of  the  validity  of  the  Bohr  theory. 


Q17     How  much  kinetic  energy  will  an  electron  have  after  a 
collision  with  a  mercury  atom  if  its  kinetic  energy  before  collision 
is  (a)  4.0  eV?  (b)  5.0  eV?  (c)  7.0  eV? 


19.10    The  periodic  table  of  the  elements 


In  the  Rutherford-Bohr  model,  atoms  of  the  different  elements 
differ  in  the  charge  and  mass  of  their  nuclei,  and  in  the  number 
and  arrangement  of  the  electrons  about  each  nucleus.  Bohr  came 
to  picture  the  electronic  orbits  as  shown  on  the  next  page,  though 
not  as  a  series  of  concentric  rings  in  one  plane  but  as  tracing  out 


Section  19.10 


83 


patterns  in  three  dimensions.  For  example,  the  orbits  of  the  two 
electrons  of  helium  in  the  normal  state  are  indicated  as  circles  in 
planes  inclined  at  about  60°  with  respect  to  each  other.  For  each 
circular  orbit,  elliptical  ones  with  the  nucleus  at  one  focus  are  also 
possible,  and  with  the  same  (or  nearly  the  same)  total  energy  as  in 
the  circular  orbit. 

Bohr  found  a  way  of  using  his  model  to  understand  better  the 
periodic  table  of  the  elements.  In  fact,  it  was  the  periodic  table 

rather  than  the  explanation  of  B aimer  spectra  that  was  Bohr's 
primary  concern  when  he  began  his  study.  He  suggested  that  the 
chemical  and  physical  properties  of  an  element  depend  on  how  the 
electrons  are  arranged  around  the  nucleus.  He  also  indicated  how 
this  might  come  about.  He  regarded  the  electrons  in  an  atom  as 
grouped  together  in  layers  or  shells  around  the  nucleus.  Each  shell 
can  contain  not  more  than  a  certain  number  of  electrons.  The 
chemical  properties  are  related  to  how  nearly  full  or  empty  a  shell 
is.  For  example,  full  shells  are  associated  with  chemical  stabiUty, 
and  in  the  inert  gases  the  electron  shells  are  completely  filled. 

To  see  how  the  Bohr  model  of  atoms  helps  to  understand 
chemical  properties  we  may  begin  with  the  observation  that  the 
elements  hydrogen  (Z  =  1)  and  hthium  (Z  =  3)  are  somewhat  alike 
chemically.  Both  have  valences  of  1.  Both  enter  into  compounds  of 
analogous  types,  for  example  hydrogen  chloride,  HCl,  and  hthium 
chloride.  LiCl.  Furthermore  there  are  some  similarities  in  their 
spectra.  All  this  suggests  that  the  lithium  atom  resembles  the 
hydrogen  atom  in  some  important  respects.  Bohr  conjectured  that 
two  of  the  three  electrons  of  the  lithium  atom  are  relatively  close 
to  the  nucleus,  in  orbits  resembling  those  of  the  helium  atom,  while 
the  third  is  in  a  circular  or  elliptical  orbit  outside  the  inner  system. 
Since  this  inner  system  consists  of  a  nucleus  of  charge  (+)  Sq^  and 
two  electrons  each  of  the  charge  (— )  <?«,,  its  net  charge  is  (+)  Qg.  Thus 
the  lithium  atom  may  be  roughly  pictured  as  having  a  central  core 
of  charge  (+)  gp,  around  which  one  electron  revolves,  somewhat  as 
for  a  hydrogen  atom.  The  analogous  physical  structure,  then,  is  the 
reason  for  the  analogous  chemical  behavior. 

Helium  (Z  =  2)  is  a  chemically  inert  element,  belonging  to  the 
family  of  noble  gases.  So  far  no  one  has  been  able  to  form  com- 
pounds from  it.  These  properties  indicated  that  the  helium  atom  is 
highly  stable,  having  both  of  its  electrons  closely  bound  to  the 
nucleus.  It  seemed  sensible  to  regard  both  electrons  as  moving  in 
the  same  innermost  shell  around  the  nucleus  when  the  atom  is 
unexcited.  Moreover,  because  the  helium  atom  is  so  stable  and 
chemically  inert,  we  may  reasonably  assume  that  this  shell  cannot 
accommodate  more  than  two  electrons.  This  shell  is  called  the 
K-shell.  The  single  electron  of  hydrogen  is  also  said  to  be  in  the 
K-shell  when  the  atom  is  unexcited.  For  lithium  two  electrons  are 
in  the  K-shell.  filling  it  to  capacity,  and  the  third  electron  starts  a 
new  one,  called  the  L-shell.  This  single  outlying  and  loosely  bound 
electron  is  the  reason  for  the  strong  chemical  affinity  of  lithium  for 
oxygen,  chlorine,  and  many  other  elements. 


The  sketches  below  are  based  on  dia- 
grams Bohr  used  in  his  university 
lectures. 


i^^KO^BU  Cfl^ 


ye.uioMCi=2>^ 


L/ry-ioM  C'i^S'y 


/VfipVc'H'/C) 


5op\UH  (2  -  ll') 


At^$e*/(^Z'^lg) 


84 


The  Rutherford-Bohr  Model  of  the  Atom 


These  two  pages  will  be  easier  to 
follow  if  you  refer  to  the  table  of  the 
elements  and  the  periodic  table  in 
Chapter  17  page  23. 


Shell  Number  of  electrons  in 

name  filled  shell 


2 

8 

18 


Sodium  (Z  =  11)  is  the  next  element  in  the  periodic  table  that 
has  chemical  properties  similar  to  those  of  hydrogen  and  lithium, 
and  this  suggests  that  the  sodium  atom  also  is  hydrogen-like  in 
having  a  central  core  about  which  one  electron  revolves.  More- 
over, just  as  lithium  follows  helium  in  the  periodic  table,  so  does 
sodium  follow  another  noble  gas,  neon  (Z  =  10).  For  the  neon  atom, 
we  may  assume  that  two  of  its  10  electrons  are  in  the  first  (K)  shell, 
and  that  the  remaining  8  electrons  are  in  the  second  (L)  shell. 
Because  of  the  chemical  inertness  and  stability  of  neon,  these  8 
electrons  may  be  expected  to  fill  the  L-shell  to  capacity.  For  sodium, 
then,  the  eleventh  electron  must  be  in  a  third  shell,  which  is  called 
the  M-shell.  Passing  on  to  potassium  (Z  =  19),  the  next  alkah  metal 
in  the  periodic  table,  we  again  have  the  picture  of  an  inner  core 
and  a  single  electron  outside  it.  The  core  consists  of  a  nucleus  with 
charge  (+)  19q^  and  with  2,  8,  and  8  electrons  occupying  the  K-.  L-. 
and  M-shells,  respectively.  The  19th  electron  revolves  around  the 
core  in  a  fourth  shell,  called  the  N-shell.  The  atom  of  the  noble 
gas  argon,  with  Z  =  18,  just  before  potassium  in  the  periodic  table, 
again  represents  a  distribution  of  electrons  in  a  tight  and  stable 
pattern,  with  2  in  the  K-,  8  in  the  L-,  and  8  in  the  M-shell. 

These  qualitative  considerations  have  led  us  to  a  consistent 
picture  of  electrons  distributed  in  groups,  or  shells,  around  the 
nucleus.  The  arrangement  of  electrons  in  the  noble  gases  can  be 
taken  to  be  particularly  stable,  and  each  time  we  encounter  a  new 
alkali  metal  in  Group  I  of  the  periodic  table,  a  new  shell  is  started; 
there  is  a  single  electron  around  a  core  which  resembles  the  pattern 
for  the  preceding  noble  gas.  We  may  expect  that  this  outlying 
electron  will  easily  come  loose  by  the  action  of  neighboring  atoms, 
and  this  corresponds  with  the  facts.  The  elements  lithium,  sodium 
and  potassium  belong  to  the  group  of  alkali  metals.  In  compounds 
or  in  solution  (as  in  electrolysis)  they  may  be  considered  to  be  in  the 
form  of  ions  such  as  Li+,  Na*  and  K^,  each  lacking  one  electron  and 
hence  having  one  positive  net  charge  (+)  q^.  In  the  neutral  atoms 
of  these  elements,  the  outer  electron  is  relatively  free  to  move  about. 
This  property  has  been  used  as  the  basis  of  a  theory  of  electrical 
conductivity.  According  to  this  theory,  a  good  conductor  has  many 
"free"  electrons  which  can  form  a  current  under  appropriate 
conditions.  A  poor  conductor  has  relatively  few  "free"  electrons. 
The  alkali  metals  are  all  good  conductors.  Elements  whose  electron 
shells  are  filled  are  very  poor  conductors;  they  have  no  "free" 
electrons. 

Turning  now  to  Group  II  of  the  periodic  table,  we  would  expect 
those  elements  that  follow  immediately  after  the  alkali  metals  to 
have  atoms  with  two  outlying  electrons.  For  example,  beryllium 
(Z  =  4)  should  have  2  electrons  in  the  K-shell.  thus  filhng  it.  and  2 
in  the  L-shell.  If  the  atoms  of  all  these  elements  have  two  outlying 
electrons,  they  should  be  chemically  similar,  as  indeed  they  are. 
Thus,  calcium  and  magnesium,  which  belong  to  this  group,  should 
easily  form  ions  such  as  Ca^^  and  Mg^^,  each  with  a  positive  net 
charge  of  (+)  2(j2.  and  this  is  also  found  to  be  tioie. 


Section  19.10  85 

As  a  final  example,  consider  those  elements  that  immediately 
precede  the  noble  gases  in  the  periodic  table.  For  example,  fluorine 
atoms  (Z  =  9)  should  have  2  electrons  filHng  the  K-shell  but  only  7 
electrons  in  the  L-shell,  which  is  one  less  than  enough  to  fill  it.  If 
a  fluorine  atom  should  capture  an  additional  electron,  it  should 
become  an  ion  F"  with  one  negative  net  charge.  The  L-shell  would 
then  be  filled,  as  it  is  for  neutral  neon  (Z  =  10),  and  thus  we  would 
expect  the  F"  ion  to  be  relatively  stable.  This  prediction  is  in  accord 
with  observation.  Indeed,  all  the  elements  immediately  preceding 
the  inert  gases  in  the  periodic  table  tend  to  form  stable  singly- 
charged  negative  ions  in  solution.  In  the  solid  state,  we  would 
expect  these  elements  to  be  lacking  in  free  electrons,  and  all  of 
them  are  in  fact  poor  conductors  of  electricity. 

Altogether  there  are  seven  main  shells,  K,  L,  M,  .  .  .  Q,  and 
further  analysis  shows  that  all  but  the  first  are  divided  into  sub- 
shells.  The  second  (L)  shell  consists  of  two  subshells,  the  third  (M) 
shell  consists  of  three  subshells,  and  so  on.  The  first  subshell  in  any 
shell  can  always  hold  up  to  2  electrons,  the  second  up  to  6,  the  third 
up  to  10,  the  fourth  up  to  14,  and  so  on.  For  all  the  elements  up  to 
and  including  argon  (Z  =  18),  the  buildup  of  electrons  proceeds 
quite  simply.  Thus  the  argon  atom  has  2  electrons  in  the  K-shell, 
8  in  the  L-shell,  then  2  in  the  first  M-subshell  and  6  in  the  second 
M-subshell.  But  the  first  subshell  of  the  N-shell  is  lower  in  energy 
than  the  third  subshell  of  the  M-shell.  Since  atoms  are  most  likely 
to  be  in  the  lowest  energy  state  available,  the  N-shell  will  begin  to 
fill  before  the  M-shell  is  completed.  Therefore,  after  argon,  there 
may  be  electrons  in  an  outer  shell  before  an  inner  one  is  filled. 


SG  19.17, 19.18 


Relative  energy  levels  of  electron 
states  in  atoms.  Each  circle  represents 
a  state  which  can  be  occupied  by  2 
electrons. 


86 


The  Rutherford-Bohr  Model  of  the  Atom 


Period 
II 


This  complicates  the  scheme  somewhat  but  still  allows  it  to  be  con- 
sistent. The  arrangement  of  the  electrons  in  any  unexcited  atom  is 
always  the  one  that  provides  greatest  stability  for  the  whole  atom. 
And  according  to  this  model,  chemical  phenomena  generally  involve 
only  the  outermost  electrons  of  the  atoms. 

Bohr  carried  through  a  complete  analysis  along  these  lines  and, 
in  1921,  proposed  the  form  of  the  periodic  table  shown  below.  The 
periodicity  results  from  the  completion  of  subshells,  which  is 
complicated  even  beyond  the  shell  overlap  in  the  figure  on  page  85 
by  the  interaction  of  electrons  in  the  same  subshell.  This  still 
useful  table  was  the  result  of  physical  theory  and  offered  a  funda- 
mental physical  basis  for  understanding  chemistry  — for  example, 
how  the  structure  of  the  periodic  table  follows  from  the  shell 
structure  of  atoms.  This  was  another  triumph  of  the  Bohr  theory. 

Period 
VII 

87  -- 

88  Ra 

89  Ac 

90  Th 
Period  Period    ///      59  Pr  91  Pa 

IV  V     ///         60  Nd  92  U 


Bohr's  periodic  table  of  the  elements  (1921).  For  example  some  of  the  names 
and  symbols  have  been  changed.  Masurium  (43)  is  now  called  Technetium 
(43),  and  Niton  (86)  is  Radon  (86).  The  rectangles  indicate  the  filling  of  sub- 
shells  of  a  higher  shell. 


Q18    Why  do  the  next  heavier  elements  after  the  noble  gases 
easily  become  positively  charged? 

Q19    Why  are  there  only  2  elements  in  Period  I.  8  in  Period  II, 
8  in  Period  III,  etc? 


19.11     The  inadequacy  of  the  Bohr  theory,  and  the  state  of  atomic 
theory  in  the  early  1920's 


As  we  are  quite  prepared  to  find,  every  model,  every  theory  has 
limits.  In  spite  of  the  successes  achieved  with  the  Bohr  theory 
in  the  years  between  1913  and  1924,  problems  arose  for  which  the 


1800 


1850 


1900 


1950 


Louisiana  Purchase 
Napoleonic  Empire 

Battle  of  Waterloo 

Monroe  Doctrine 

Discovery  of  Electro-magnetic  Induction 

(V 

1 

.2 
1 

-4-> 
M 

C 

3 

s 
a 

1 

< 

""elephone 

*-> 
C 

! 

BOHR 

I 

03 

> 
LU 

"(5 
o 

o 

0) 

:S 
<t-i 
o 

•1 

s 

1 

i 

11 
si 

.a  « 

66 

Hg 

1 

c              ^ 
o  g        S 

m 

2        CO 
c       ^ 

•£     ,5 

o        «« 

ia  M  o 

■ 

I 

X 

9 

1                ^^K^S? 

^^Hp 

1 

r- 

n 

0) 

E 

VICTORIA 

J  JOHN  F.  KENNEDY 

^K-  -X' 

n 

V^H 

^H!:' 

a; 

ABRAHAM  LINCOLN 

NIKOLAI  LENIN                                                        ^^^^^H 

r^ 

1 

> 
o 

FRANKLIN  D.  ROOSEVELT 

■^1 

O 

THOMAS  YOUNG 

L 

MARIE  CURIE                                                                                 ^H 

JOHN  DALTON 

ERNEST  RUTHERFORD                                                             | 

HANS  CHRISTIAN  OERSTED 

I 

MICHAEL  FARADAY 

CHARLES  DARWIN 

GREGOR  MENDEL 

DMITRI  MENDELEEV 


88 


The  Rutherford-Bohr  Model  of  the  Atom 


In  March  1913,  Bohr  wrote  to  Ruther- 
ford enclosing  a  draft  of  his  first 
paper  on  the  quantum  theory  of 
atomic  constitution.  On  March  20, 
1913,  Rutherford  replied  in  a  letter, 
the  first  part  of  which  we  quote, 
"Dear  Dr.  Bohr: 

I  have  received  your  paper  and 
read  it  with  great  interest,  but  I  want 
to  look  it  over  again  carefully  when 
I  have  more  leisure.  Your  ideas  as 
to  the  mode  of  origin  of  spectra  in 
hydrogen  are  very  ingenious  and 
seem  to  work  out  well;  but  the  mix- 
ture of  Planck's  ideas  with  the  old 
mechanics  makes  it  very  difficult 
to  form  a  physical  idea  of  what  is 
the  basis  of  it.  There  appears  to  me 
one  grave  difficulty  in  your  hypoth- 
esis, which  I  have  no  doubt  you  fully 
realize,  namely,  how  does  an  elec- 
tron decide  what  frequency  it  is 
going  to  vibrate  at  when  it  passes 
from  one  stationary  state  to  the 
other.  It  seems  to  me  that  you  would 
have  to  assume  that  the  electron 
knows  before  hand  where  it  is  going 
to  stop.  .  . 


theory  proved  inadequate.  Bohr's  theory  accounted  excellently  for 
the  spectra  of  atoms  with  a  single  electron  in  the  outermost  shell, 
but  serious  discrepancies  between  theory  and  experiment  appeared 
in  the  spectra  of  atoms  with  two  electrons  or  more  in  the  outermost 
shell.  It  was  also  found  experimentally  that  when  a  sample  of  an 
element  is  placed  in  an  electric  or  magnetic  field,  its  emission 
spectrum  shows  additional  lines.  For  example,  in  a  magnetic  field 
each  line  is  split  into  several  lines.  The  Bohr  theory  could  not 
account  in  a  quantitative  way  for  the  observed  splitting.  Further, 
the  theory  supplied  no  method  for  predicting  the  relative  brightness 
of  spectral  lines.  These  relative  intensities  depend  on  the  probabili- 
ties with  which  atoms  in  a  sample  undergo  transitions  among  the 
stationary  states.  Physicists  wanted  to  be  able  to  calculate  the 
probability  of  a  transition  from  one  stationary  state  to  another.  They 
could  not  make  such  calculations  with  the  Bohr  theory. 

By  the  early  1920's  it  had  become  clear  that  the  Bohr  theory, 
despite  its  great  successes,  was  deficient  beyond  certain  limits. 
It  was  understood  that  to  get  a  theory  that  would  be  successful  in 
solving  more  problems,  Bohr's  theory  would  have  to  be  revised,  or 
replaced  by  a  new  one.  But  the  successes  of  Bohr's  theory  showed 
that  a  better  theory  of  atomic  structure  would  still  have  to  account 
also  for  the  existence  of  stationary  states  — discrete  atomic  energy 
levels  — and  would,  therefore,  have  to  be  based  on  quantum 
concepts. 

Besides  the  inability  to  predict  certain  properties  of  atoms  at  all, 
the  Bohr  theory  had  two  additional  shortcomings:  it  predicted  some 
results  that  were  not  in  accord  with  experiment  (such  as  the 
spectra  of  elements  with  two  or  three  electrons  in  the  outermost 
electron  shells);  and  it  predicted  others  that  could  not  be  tested  in 
any  known  way  (such  as  the  details  of  electron  orbits).  Although 
orbits  were  easy  to  draw  on  paper,  they  could  not  be  observed 
directly,  nor  could  they  be  related  to  any  observable  properties  of 
atoms.  Planetary  theory  has  very  different  imphcations  when 
applied  to  a  real  planet  moving  in  an  orbit  around  the  sun.  and 
when  applied  to  an  electron  in  an  atom.  The  precise  position  of  a 
planet  is  important,  especially  if  we  want  to  do  experiments  such 
as  photographing  an  eclipse,  or  a  portion  of  the  surface  of  Mars 
from  a  satellite.  But  the  moment-to-moment  position  of  an  electron 
in  an  orbit  has  no  such  meaning  because  it  has  no  relation  to  any 
experiment  physicists  have  been  able  to  devise.  It  thus  became 
evident  that,  in  using  the  Bohr  theory,  physicists  could  be  led  to  ask 
some  questions  which  could  not  be  answered  experimentally. 

In  the  early  1920's,  physicists  — above  all.  Bohr  himself— began 
to  work  seriously  on  the  revision  of  the  basic  ideas  of  the  theory. 
One  fact  that  stood  out  was  that  the  theory  started  with  a  mixture 
of  classical  and  quantum  ideas.  An  atom  was  assumed  to  act  in 
accordance  with  the  laws  of  classical  physics  up  to  the  point  where 
these  laws  did  not  work;  then  the  quantum  ideas  were  introduced. 
The  picture  of  the  atom  that  emerged  from  this  inconsistent  mixture 
was  a  combination  of  ideas  from  classical  physics  and  concepts  for 


Section  19.11 


89 


which  there  was  no  place  in  classical  physics.  The  orbits  of  the 
electrons  were  determined  by  the  classical,  Newtonian  laws  of 
motion.  But  of  the  many  possible  orbits,  only  a  small  portion  were 
regarded  as  possible,  and  these  were  selected  by  rules  that  contra- 
dicted classical  mechanics.  Or  again,  the  frequency  calculated  for 
the  orbital  revolution  of  electrons  was  quite  different  from  the 
frequency  of  light  emitted  or  absorbed  when  the  electron  moved 
from  or  to  this  orbit.  Or  again,  the  decision  that  n  could  never  be 
zero  was  purely  arbitrary,  just  to  prevent  the  model  from  collapsing 
by  letting  the  electron  fall  on  the  nucleus.  It  became  evident  that 
a  better  theory  of  atomic  structure  would  have  to  be  built  on  a  more 
consistent  foundation  in  quantum  concepts. 

In  retrospect,  the  contribution  of  the  Bohr  theory  may  be  sum- 
marized as  follows.  It  provided  some  excellent  answers  to  earlier 
questions  raised  about  atomic  structure  in  Chapters  17  and  18. 
Although  the  theory  turned  out  to  be  inadequate  it  drew  attention  to 
how  quantum  concepts  can  be  used.  It  indicated  the  path  that  a 
new  theory  would  have  to  take.  A  new  theory  would  have  to  supply 
the  right  answers  that  the  Bohr  theory  gave,  and  would  also  have  to 
supply  the  right  answers  for  the  problems  the  Bohr  theory  could 
not  solve.  And  without  doubt  one  of  the  most  intriguing  aspects  of 
Bohr's  work  was  the  proof  that  physical  and  chemical  properties  of 
matter  can  be  traced  back  to  the  fundamental  role  of  integers  — 
(quantum  numbers  such  as  n  =  1,  2,  3  .  .  .).  As  Bohr  said,  "The 
solution  of  one  of  the  boldest  dreams  of  natural  science  is  to  build 
up  an  understanding  of  the  regularities  of  nature  upon  the  con- 
sideration of  pure  number."  We  catch  here  an  echo  of  the  hope  of 
Pythagoras  and  Plato,  of  Kepler  and  GaUleo. 

Since  the  1920's,  a  successful  theory  of  atomic  structure  has 
been  developed  and  has  been  generally  accepted  by  physicists.  It  is 
part  of  "quantum  mechanics,"  so  called  because  it  is  built  directly 
on  quantum  concepts;  it  goes  now  far  beyond  understanding  atomic 
structure,  and  in  fact  is  the  basis  of  our  modern  conception  of 
events  on  a  submicroscopic  scale.  Some  aspects  will  be  discussed  in 
the  next  chapter.  Significantly,  Bohr  himself  was  again  a  leading 
contributor. 


Remember,  for  example,  (in  Unit  1) 
how  proudly  Galileo  pointed  out, 
when  announcing  that  all  falling 
bodies  are  equally  and  constantly 
accelerated:  "So  far  as  I  know, 
no  one  has  yet  pointed  out  that 
the  distances  traversed,  during 
equal  intervals  of  time,  by  a  body 
falling  from  rest,  stand  to  one 
another  in  the  same  ratio  as  the 
odd  numbers  beginning  with  unity 
[namely  1:3:5:7:  . .  .]." 


SG  19.19-19.23 


Q20    The  Bohr  model  of  atoms  is  widely  given  in  science 
books.  What  is  wrong  with  it?  What  is  good  about  it? 


STUDY  GUIDE 


19.1  The  Project  Physics  materials  particularly 
appropriate  for  Chapter  19  include: 

Experiment 

Spectroscopy 

Activities 

Scientists  on  stamps 

Measuring  ionization,  a  quantum  effect 

"Black  box"  atoms 

Reader  Article 

The  Teacher  and  the  Bohr  Theory  of  the  Atom 

Film  Loop 

Rutherford  Scattering 

Transparencies 

Alpha  Scattering 

Energy  Levels  — Bohr  Theory 

19.2  (a)  Suggest  experiments  to  show  which  of 

the  Fraunhofer  lines  in  the  spectrum  of 
sunlight  are  due  to  absorption  in  the 
sun's  atmosphere  rather  than  to  absorp- 
tion by  gases  in  the  earth's  atmosphere. 

(b)  How  might  one  decide  from  spectro- 
scopic observations  whether  the  moon 
and  the  planets  shine  by  their  own  light 
or  by  reflected  light  from  the  sun? 

19.3  Theoretically,  how  many  series  of  lines  are 
there  in  the  emission  spectrum  of  hydrogen?  In 
all  these  series,  how  many  lines  are  in  the  visible 
region? 

19.4  The  Rydberg  constant  for  hydrogen,  JR„,  has 
the  value  1.097  x  lOVm.  Calculate  the  wave- 
lengths of  the  lines  in  the  Balmer  series 
corresponding  to  n  =  8,  n  =  10,  n  =  12.  Compare 
the  values  you  get  with  the  wavelengths  listed 

in  the  table  on  p.  63.  Do  you  see  any  trend  in  the 
values? 


19.6  In  what  ways  do  Thomson's  and  Ruther- 
ford's atomic  models  agree?  In  what  ways  do 
they  disagree? 

19.7  In  1903,  the  German  physicist  Philipp 
Lenard  (1864-1947)  proposed  an  atomic  model 
diff'erent  from  those  of  Thomson  and  Rutherford. 
He  observed  that,  since  cathode-ray  particles 
can  penetrate  matter,  most  of  the  atomic  volume 
must  off'er  no  obstacle  to  their  penetration.  In 
Lenard's  model  there  were  no  electrons  and  no 
positive  charges  separate  from  the  electrons.  His 
atom  was  made  up  of  particles  called  dynamides, 
each  of  which  was  an  electric  doublet  possessing 
mass.  (An  electric  doublet  is  a  combination  of  a 
positive  charge  and  a  negative  charge  very  close 
together.)  All  the  dynamides  were  supposed  to  be 
identical,  and  an  atom  contained  as  many  of 
them  as  were  needed  to  make  up  its  mass.  They 
were  distributed  throughout  the  volume  of  the 
atom,  but  their  radius  was  so  small  compared 
with  that  of  the  atom  that  most  of  the  atom  was 
empty. 

(a)  In  what  ways  does  Lenard's  model  agree 
with  those  of  Thomson  and  Rutherford?  In 
what  ways  does  it  disagree  with  those 
models? 

(b)  Why  would  you  not  expect  a  particles  to  be 
scattered  through  large  angles  if  Lenard's 
model  were  valid? 

(c)  In  view  of  the  scattering  of  a  particles  that 
is  observed,  is  Lenard's  model  valid? 

19.8  Determine  a  plausible  upper  limit  for  the 
eff'ective  size  of  a  gold  atom  from  the  following 
facts  and  hypotheses: 

i.  A  beam  of  a-particles  of  known  velocity  v  = 
2  X  10'  m/sec  is  scattered  from  a  gold  foil  in  a 
manner  explicable  only  if  the  a  particles  were 
repelled  by  nuclear  charges  that  exert  a  Coulomb's 
law  repulsion  on  the  a  particles. 


19.5     (a)  As  indicated  in  the  figure  on  p.  63  the 

lines  in  one  of  hydrogen's  spectral  series 
are  bunched  very  closely  at  one  end. 
Does  the  formula 


A  in 


suggest  that  such  bunching  will  occur? 

(b)  The  "series  limit"  corresponds  to  the 
last  possible  line(s)  of  the  series.  What 
value  should  be  taken  for  n,  in  the 
above  equation  to  compute  the  wave- 
length of  the  series  limit? 

(c)  Compute  the  series  limit  for  the  Lyman, 
Balmer,  or  Paschen  series  of  hydrogen. 

(d)  Consider  a  photon  with  a  wavelength 
corresponding  to  the  series  limit  of  the 
Lyman  series.  What  energy  could  it 
carry?  Express  the  answer  in  joules  and 
in  electron  volts  (1  eV  =  1.6  x  lO"'*  J). 


ii.  Some  of  these  a  particles  come  straight  back 
after  scattering.  They  therefore  approached  the 
nuclei  up  to  a  distance  r  from  the  nucleus'  center, 
where  the  initicd  kinetic  energy  jTn„vJ  has  been 
completely  changed  to  the  potential  energy  of 
the  system. 

iii.  The  potential  energy  of  a  system  made  up 
of  an  a  particle  of  charge  2q^  at  a  distance  r 
from  a  nucleus  of  charge  Zq^  is  given  by  the 
product  of  the  "potential"  (Zqjr)  set  up  by  the 
nucleus  at  distance  r,  and  the  charge  (2^^)  of 
the  a  particle. 

iv.  The  distance  r  can  now  be  computed,  since 
we  know  Va,  rUa  (7  x  10~"  kg.  from  other  evidences 
to  be  discussed  in  Unit  6).  Z  for  gold  atoms  (see 
periodic  table),  q^  (see  Section  14.5). 

V.  The  nuclear  radius  must  be  equal  to  or  less 
than  r.  Thus  we  have  a  plausible  upper  limit 
for  the  size  of  this  nucleus. 


90 


19.9    We  generally  suppose  that  the  atom  and  the 
nucleus  are  each  spherical,  that^the  diameter  of 
the  atom  is  of  the  order  of  1  A  (Angstrom  unit  = 
10"'"  m)  and  that  the  diameter  of  the  nucleus  is 
of  the  order  of  10"'-  cm. 


the  hydrogen  atom  for  each  of  the  first  4  allowed 
orbits  (n  =  1,  2,  3,  4). 

vi.  As  a  final  point,  show  that  the  quantity  —EJhc 
has  the  same  value  as  the  constant  R^,  as 
claimed  in  Sec.  19.8. 


(a)  What  are  the  evidences  that  these  are 
reasonable  suppositions? 

(b)  What  is  the  ratio  of  the  diameter  of  the 
nucleus  to  that  of  the  atom? 

19.10  The  nucleus  of  the  hydrogen  atom  is 
thought  to  have  a  radius  of  about  1.5  x  10"'^  cm. 
If  the  atom  were  magnified  so  that  the  nucleus  is 
0.1  mm  across  (the  size  of  a  grain  of  dust),  how 
far  away  from  it  would  the  electron  be  in  the 
Bohr  orbit  closest  to  it? 

19.11  Show  that  the  total  energy  of  a  neutral 
hydrogen  atom  made  up  of  a  positively  charged 
nucleus  and  an  electron  is  given  by 

n^     ' 

where  E,  is  the  energy  when  the  electron  is  in  the 
first  orbit  (n  =  1),  and  where  the  value  of 
E,  =  —13.6  electron-volts.  (You  may  consult  other 
texts,  for  example  Foundation  of  Modern  Physical 
Science  by  Holton  and  Roller,  sections  34.4  and 
34.7.)  Program  and  hints: 

i.  The  total  energy  E  of  the  system  is  the  kinetic 
and  potential  energy' KE  +  PE  of  the  electron  in 
its  orbit.  Since  mv-jr  =  k  q/lr'^  (see  p.  73), 
KE  =  imi;-  can  be  quickly  calculated. 

ii.  The  electrical  potential  energy  PE  of  a  charged 
point  object  (electron)  is  simply  given  by  the 
electrical  potential  V  of  the  region  in  which  it 
finds  itself,  times  its  own  charge.  The  value  of 
V  set  up  by  the  (positive)  nucleus  at 
distance  r  is  given  by  Kqjr  and  the  charge  on  the 
electron  (including  sign,  for  once!)  is  — q^.  Hence 
PE  =  —kq^glr.  The  meaning  of  the  negative  sign 
is  simply  that  PE  is  taken  to  be  zero  if  the  elec- 
tron is  infinitely  distant;  the  system  radiates 
energy  as  the  electron  is  placed  closer  to  the 
nucleus,  or  conversely  that  energy  must  be  sup- 
phed  to  move  the  electron  away  from  the  nucleus. 

iii.  Now  you  can  show  that  the  total  energy  E  is 


E  =  KE  +  PE  =  -k 


2r 


iv.  Using  the  equation  derived  on  p.  73,  namely 

n'^h^ 
r  =  -7—; —,  show  that 


47r^Treg^ 


E   = 


^_k^2nhnq^_  1 


n^h^ 


£, 


where  E^  =  k^2TT^niq/lh^. 


The  numerical  value  for  this  can  be  computed  by 
using  the  known  values  (in  consistent  units) 
for  k,  m,  q^  and  h. 

V.  Find  the  numerical  value  of  the  energy  of 


19.12  Using  the  Bohr  theory,  how  would  you 
account  for  the  existence  of  the  dark  lines  in 
the  absorption  spectrum  of  hydrogen? 

19.13  A  group  of  hydrogen  atoms  is  excited  (by 
collision,  or  by  absorption  of  a  photon  of  proper 
frequency),  and  they  all  are  in  the  stationary  state 
for  which  n  =  5.  Refer  to  the  figure  in  the  margin 
on  p.  78  and  list  all  possible  lines  emitted  by  this 
sample  of  hydrogen  gas. 

19.14  Make  an  energy  level  diagram  to  represent 
the  results  of  the  Franck-Hertz  experiment. 

19.15  Many  substances  emit  visible  radiation 
when  illuminated  with  ultraviolet  hght;  this 
phenomenon  is  an  example  of  fluorescence. 
Stokes,  a  British  physicist  of  the  nineteenth 
century,  found  that  in  fluorescence  the  wave- 
length of  the  emitted  light  usually  was  the  same 
or  longer  than  the  illuminating  light.  How  would 
you  account  for  this  phenomenon  on  the  basis  of 
the  Bohr  theory? 

19.16  In  Query  31  of  his  Opticks,  Newton  wrote: 

All  these  things  being  consider'd,  it 
seems  probable  to  me  that  God  in  the 
beginning  formed  matter  in  solid,  massy, 
hard,  impenetrable,  moveable  particles,  of 
such  sizes  and  figures,  and  with  such 
other  properties,  and  in  such  proportion  to 
the  end  for  which  he  formed  them;  and 
that  these  primitive  particles  being  sohds, 
are  incomparably  harder  than  any  porous 
bodies  compounded  of  them,  even  so  very 
hard,  as  never  to  wear  or  break  in  pieces; 
no  ordinary  power  being  able  to  divide 
what  God  himself  made  one  in  first  the 
creation.  While  the  particles  continue 
entire,  they  may  compose  bodies  of  one 
and  the  same  nature  texture  and  in  all 
ages:  But  should  they  wear  away,  or  break 
in  pieces,  the  nature  of  things  depending 
on  them  would  be  changed.  Water  and 
earth,  composed  of  old  worn  particles  and 
fragments  of  particles,  would  not  be  of  the 
same  nature  and  texture  now,  with  water 
and  earth  composed  of  entire  particles  in 
the  beginning.  And  therefore  that  nature 
may  be  lasting,  the  changes  of  corporeal 
things  are  to  be  placed  only  in  the  various 
separations  and  new  associations  and 
motions  of  these  permanent  particles; 
compound  bodies  being  apt  to  break,  not 
in  the  midst  of  solid  particles,  but  where 
those  particles  are  laid  together,  and  only 
touch  in  a  few  points. 

Compare  what  Newton  says  here  about  atoms  with 

91 


(a)  the  views  attributed  to  Leucippus  and 
Democritus  concerning  atoms  (see  the 
Prologue  to  this  unit); 

(b)  Dalton's  assumptions  about  atoms  (see  the 
end  of  the  prologue  to  this  unit); 

(c)  the  Rutherford-Bohr  model  of  the  atom 

19-17    Use  the  chart  on  p.  85  to  explain  why 
atoms  of  potassium  (Z  =  19)  have  electrons  in 
the  N  shell  even  though  the  M  shell  is  not  filled. 

19.18  Use  the  chart  on  p.  85  to  predict  the 
atomic  number  of  the  text  inert  gas  after  argon. 
That  is,  imagine  filling  the  electron  levels  with 
pairs  of  electrons  until  you  reach  an  apparently 
stable,  or  complete,  pattern.  Do  the  same  for  the 
next  inert  gas  following. 

19.19  Make  up  a  glossary,  with  definitions,  of 
terms  which  appeared  for  the  first  time  in  this 
chapter. 

19.20  The  philosopher  John  Locke  (1632-1704) 
proposed  a  science  of  human  nature  which  was 
strongly  influenced  by  Newton's  physics.  In 
Locke's  atomistic  view,  elementary  ideas  ("atoms") 
are  produced  by  elementary  sensory  experiences 
and  then  drift,  collide  and  interact  in  the  mind. 
Thus  the  association  of  ideas  was  but  a  special 
case  of  the  universal  interactions  of  particles. 

Does  such  an  approach  to  the  sulaject  of 
human  nature  seem  reasonable  to  you?  What 
argument  for  and  against  this  sort  of  theory  can 
you  think  of? 

19.21  In  a  recently  published  textbook  of  physics, 
the  following  statement  is  made: 

Arbitrary  though  Bohr's  new  postulate 
may  seem,  it  was  just  one  more  step  in 
the  process  by  which  the  apparently  con- 
tinuous macroscopic  world  was  being 


analyzed  in  terms  of  a  discontinuous, 
quantized,  microscopic  world.  Although 
the  Greeks  had  speculated  about  quan- 
tized matter  (atoms),  it  remained  for  the 
chemists  and  physicists  of  the  nineteenth 
century  to  give  them  reality.  In  1900 
Planck  found  it  necessary  to  quantize  the 
energy  of  electromagnetic  waves.  Also,  in 
the  early  1900's  a  series  of  experiments 
culminating  in  Millikan's  oil-drop  experi- 
ment conclusively  showed  that  electric 
charge  was  quantized.  To  this  hst  of 
quantized  entities.  Bohr  added  angular 
momentum  (mvr). 

(a)  What  other  properties  or  things  in  physics 
can  you  think  of  that  are  "quantized?" 

(b)  What  properties  or  things  can  you  think  of 
outside  physics  that  might  be  said  to  be 
"quantized?"" 

19.22  Write  an  essay  on  the  successes  and 
failures  of  the  Bohr  model.  Can  it  be  called  a  good 
model?  A  simple  model?  A  beautiful  model? 

19.23  In  1903  a  philosopher  wrote: 

The  propounders  of  the  atomic  view  of 
electricity  disagree  with  theories  which 
would  restrict  the  method  of  science  to  the 
use  of  only  such  quantities  and  data  as 
can  be  actually  seen  and  directly  mea- 
sured, and  which  condemn  the  introduc- 
tion of  such  useful  conceptions  as  the 
atom  and  the  electron,  which  cannot  be 
directly  seen  and  can  only  be  measured  by 
indirect  processes. 

On  the  basis  of  the  information  now  available 
to  you,  with  which  view  do  you  agree:  the  view  of 
those  who  think  in  terms  of  atoins  and  electrons, 
or  the  view  that  we  must  use  only  such  things  as 
can  be  actually  seen  and  measured? 


92 


This  construction  is  meant  to  represent  the  arrangement  of  mutually 
attracting  sodium  and  chlorine  ions  in  a  crystal  of  common  salt.  Notice 
that  the  outermost  electrons  of  the  sodium  atoms  have  been  lost  to  the 
chlorine  atoms,  leaving  positively  charged  sodium  ions  with  completed 
K  and  L  shells,  and  negatively  charged  chlorine  ions  with  completed  K, 
L  and  M  shells. 


20.1  Some  results  of  relativity  theory  95 

20.2  Particle-like  behavior  of  radiation  99 

20.3  Wave-like  behavior  of  particles  101 

20.4  Mathematical  vs  visualizable  atoms  104 

20.5  The  uncertainty  principle  108 

20.6  Probability  interpretation  111 


The  diffraction  pattern  on  the  left  was  made  by  a  beam  of  x  rays  passing  through 
thin  aluminum  foil.  The  diffraction  pattern  on  the  right  was  made  by  a  beam  of 
electrons  passing  through  the  same  foil. 


CHAPTER  TWENTY 


Some  Ideas  from  Modern 
Physical  Theories 


20.1     Some  results  of  relativity  theory 

Progress  in  atomic  and  nuclear  physics  has  been  based  on  two 
great  advances  in  physical  thought:  quantum  theory  and  relativity. 
In  so  short  a  space  as  a  single  chapter  we  cannot  even  begin  to  give        SG  20.1 
a  coherent  account  of  the  actual  development  of  physical  and 
mathematical  ideas  in  these  fields.  All  we  can  do  is  offer  you  some 
idea  of  what  kind  of  problems  led  to  the  development,  suggest 
some  of  the  unexpected  conclusions,  prepare  for  material  in  later 
chapters,  and  — very  important!— introduce  you  to  the  beautiful 
ideas  on  relativity  theory  and  quantum  mechanics  — offered  in 
articles  in  Reader  5. 

In  Chapters  18  and  19  we  saw  how  quantum  theory  entered 
into  atomic  physics.  To  follow  its  further  development  into  quantum 
mechanics,  we  need  to  learn  some  of  the  results  of  the  relativity 
theory.  These  results  will  also  be  essential  to  our  treatment  of 
nuclear  physics  in  Unit  6.  We  shall,  therefore,  devote  this  section 
to  a  brief  discussion  of  one  essential  result  of  the  theory  of  relativity 
introduced  by  Einstein  in  1905  — the  same  year  in  which  he 
published  the  theory  of  the  photoelectric  effect. 

In  Unit  1  we  discussed  the  basic  idea  of  relativity  — that  certain 
aspects  of  physical  events  appear  the  same  from  different  frames 
of  reference,  even  if  the  reference  frames  are  moving  with  respect 
to  one  another.  We  said  there  that  mass,  acceleration,  and  force 
seemed  to  be  such  invariant  quantities,  and  Newton's  laws  relating 
them  were  equally  good  in  all  reference  frames. 

By  1905  it  had  become  clear  that  this  is  true  enough  for  all 
ordinary  cases  of  motion,  but  not  if  the  bodies  involved  move  with 
respect  to  the  observer  at  a  speed  more  than  a  few  percent  of  that 
of  light.  Einstein  considered  whether  the  same  relativity  principle 
could  be  extended  to  include  not  only  the  mechanics  of  rapidly 
moving  bodies,  but  also  the  description  of  electromagnetic  waves. 
He  found  this  could  be  done  by  replacing  Newton's  definitions  of 
length  and  time  by  others  that  produce  a  more  consistent  physics. 

95 


96 


Some  Ideas  From  Modern  Physical  Theories 


Topics  in  relativity  theory  are 
developed  further  in  Reader  5. 
See  the  articles: 
"The  Clock  Paradox" 
"Mr.  Tompkins  and  Simultaneity" 
"Mathematics  and  Relativity" 
"Parable  of  the  Surveyors  ' 
"Outside  and  Inside  the  Elevator" 
"Space  Travel:  Problems  of  Physics 
and  Engineering" 


one  that  resulted  in  a  new  viewpoint.  The  viewpoint  is  the  most 
interesting  part  of  Einstein's  thinking,  and  parts  of  it  are  discussed 
in  articles  in  Reader  5  and  Reader  6;  but  here  we  will  deal  with 
high-speed  phenomena  from  an  essentially  Newtonian  viewpoint, 
in  terms  of  corrections  required  to  make  Newtonian  mechanics  a 
better  fit  to  a  new  range  of  phenomena. 

For  bodies  moving  at  speeds  which  are  small  compared  to  the 
speed  of  light,  measurements  predicted  by  relativity  theory  are  only 
negligibly  different  from  measurements  predicted  by  Newtonian 
mechanics.  This  must  be  true  because  we  know  that  Newton's  laws 
account  very  well  for  the  motion  of  the  bodies  with  which  we  are 
familiar  in  ordinary  life.  The  differences  between  relativistic 
mechanics  and  Newtonian  mechanics  become  noticeable  in 
experiments  involving  high-speed  particles. 

We  saw  in  Sec.  18.2  that  J.  J.  Thomson  devised  a  method  for 
determining  the  speed  v  and  the  ratio  of  charge  to  mass  qjm  for 
electrons.  Not  long  after  the  discovery  of  the  electron  by  Thomson, 
it  was  found  that  the  value  of  Qp/m  seemed  to  vary  with  the  speed 
of  the  electrons.  Between  1900  and  1910,  several  physicists  found 
that  electrons  have  the  value  ^p/m  =  1.76  x  10"  coul/kg  only  for 
speeds  that  are  very  small  compared  to  the  speed  of  light;  the  ratio 
became  smaller  as  electrons  were  given  greater  speeds.  The  relativity 
theory  offered  an  explanation  for  these  results:  the  electron  charge 
is  invariant  — it  does  not  depend  on  the  speed  of  the  electrons;  but 
the  mass  of  an  electron,  as  an  observer  in  a  laboratory  would 
measure  it,  should  vary  with  speed,  increasing  according  to  the 
formula: 


The  Relativistic  Increase  of 
Mass  with  Speed 

v/c  m/m,,        v/c  m/m„ 


0.0 

1.000 

0.95 

3.203 

0.01 

1.000 

0.98 

5.025 

0.10 

1.005 

0.99 

7.089 

0.50 

1.155 

0.998 

15.82 

0.75 

1.538 

0.999 

22.37 

0.80 

1.667 

0.9999 

70.72 

0.90 


2.294 


0.99999 


223.6 


VI  -  y-'lc' 
In  this  formula,  v  is  the  speed  the  electron  has  relative  to  the 
observer,  c  is  the  speed  of  light  in  a  vacuum,  and  m„  is  the  rest 
mass  — the  electron's  mass  measured  by  an  observer  when  an 
electron  is  at  rest  with  respect  to  the  observer;  m  is  the  mass  of 
an  electron  measured  while  it  moves  with  speed  v  relative  to  the 
observer.  We  may  call  m  the  relativistic  mass.  It  is  the  mass 
determined,  for  example,  by  means  of  J.  J.  Thomsons  method. 

The  ratio  of  relativistic  mass  to  rest  mass,  mlm^.  which  is  equal 
to  1/Vl  -  v-lc'\  is  listed  in  the  table  in  the  margin  for  values  of  vie 
which  approach  1.  The  value  of  mlm^^  becomes  very  large  as  v 
approaches  c. 

The  formula  for  the  relativistic  mass,  which  was  derived  by 
Einstein  from  fundamental  ideas  of  space  and  time,  has  been  tested 
experimentally;  some  of  the  results,  for  electrons  with  speeds  so 
high  that  the  value  of  v  reaches  about  0.8  c,  are  shown  as  points  on 
the  graph  on  the  next  page.  At  z;  =  0.8  c  the  relativistic  mass  m  is 
about  1.7  times  the  rest  mass  m„.  The  curve  shows  the  theoretical 
variation  of  m  as  the  value  of  v  increases,  and  the  dots  and  crosses 
are  results  from  two  different  experiments.  The  agreement  of 
experiment  and  theory  is  excellent.  The  increase  in  mass  with 
speed  fully  accounts  for  the  shrinking  of  the  ratio  qjm  with  speed, 
which  was  mentioned  earlier. 


Section  20.1 


97 


Variation  of  relativistic  mass  with 
speed  (expressed  as  a  fraction  of 
the  speed  of  light). 


The  formula  for  variation  of  mass  with  speed  is  vahd  for  all 
moving  bodies,  not  just  for  electrons  and  other  atoinic  particles.  But 
larger  bodies,  such  as  those  with  which  we  are  familiar  in  everyday 
life,  we  observe  at  speeds  so  small  compared  to  that  of  light  that 
the  value  of  vie  is  very  small.  The  value  of  v^lc'  in  the  denominator 
is  then  extremely  small,  and  the  values  of  m  and  m^  are  so  nearly 
the  same  that  we  cannot  tell  the  difference.  In  other  words,  the 
relativistic  increase  in  mass  can  be  detected  in  practice  only  for 
particles  of  atomic  or  sub-atomic  size,  those  that  can  be  given 
speeds  higher  than  a  small  fraction  of  c. 

The  effects  discussed  so  far  are  mainly  of  historical  interest 
because  they  helped  to  convince  physicists  (eventually)  of  the 
correctness  of  relativity  theory.  Experiments  done  more  recently 
provide  more  striking  evidence  of  the  inadequacy  of  Newtonian 
physics  for  particles  with  very  high  speeds.  Electrons  can  be  given 
very  high  energies  by  accelerating  them  in  a  vacuum  by  means  of 
a  high  voltage  V.  Since  the  electron  charge  q^  is  known,  the  energy 
increase,  q^v,  is  known,  the  rest  mass  m,,  of  an  electron  is  also 
known  (see  Sec.  18.3),  and  the  speed  v  can  be  measured  by  timing 
the  travel  over  a  known  distance.  It  is,  therefore,  possible  to 
compare  the  values  of  the  energy  supplied,  q^V,  with  the  expression 
for  kinetic  energy  in  classical  mechanics,  Ttriov''^.  When  experiments 
of  this  kind  are  done,  it  is  found  that  when  the  electrons  have 
speeds  that  are  small  compared  to  the  speed  of  light,  TTnoi^"'  =  ^pV. 
We  used  this  relation  in  Sec.  18.5  in  discussing  the  photoelectric 


SG  20.2-20.4 


98 


Some  Ideas  From  Modern  Physical  Theories 


SPtBD  Sfi/A/?ei> 


i 

, 

2c'* 

y  CLASSICAL 

/ 

/ 

/ 

/ 
. _^ 

£. (   . ,     ■( 1 >          t          »>■ 

I     o-i     luf.    OS     o-i, 
fCinBTic  fvEftjy   (MeV) 


Unit  6  deals  further  with  acceler- 
ators, and  the  operation  of  the  CEA 
apparatus  is  also  the  subject  of  a 
Project  Physics  film  Synchrotron. 


effect.  We  could  do  so  quite  correctly  because  photoelectrons  do, 
indeed,  have  small  speeds,  and  m  and  mo  have  nearly  the  same 
value.  But  when  the  speed  of  the  electron  becomes  large,  so  that 
vie  is  no  longer  a  small  fraction,  it  is  found  that  the  quantity  Tmoi^^ 
does  not  increase  in  proportion  to  q^V;  this  discrepancy  increases  as 
QgV  increases.  The  increase  in  kinetic  energy  still  is  equal  to  the 
amount  of  work  done  by  the  electrical  field,  q^V,  but  the  mass  is  no 
longer  mo  and  so  kinetic  energy  can't  be  measured  by  j'^o^^-  The 
value  of  v^,  instead  of  steadily  increasing  with  energy  supplied, 
approaches  a  limiting  value:  c^. 

In  the  Cambridge  Electron  Accelerator  (CEA)  operated  in 
Cambridge,  Massachusetts,  by  Harvard  University  and  the 
Massachusetts  Institute  of  Technology,  electrons  are  accelerated 
to  an  energy  which  is  equivalent  to  what  they  would  gain  in  being 
accelerated  by  a  potential  difference  of  6  x  10^  volts;  it  is  an 
enormous  energy  for  electrons.  The  speed  attained  by  the  electrons 
is  0.999999996  c;  at  this  speed  the  relativistic  mass  m  (both  by 
calculation  and  by  experiment)  is  over  10,000  times  greater  than 
the  rest  mass  mo! 

Another  way  of  saying  mass  increases  with  speed  is  this:  any 
increase  in  kinetic  energy  is  consistently  accompanied  by  an 
increase  in  mass.  If  the  kinetic  energy  measured  from  a  frame  of 
reference  is  KE,  the  increase  in  mass  Am  (above  the  rest  mass) 
measured  in  that  frame  is  proportional  to  KE: 


Am  ^  KE 


To  increase  the  mass  of  a  body 
by  1  gram,  it  would  have  to  be  given 
a  kinetic  energy  of  10"  joules  (about 
6  million  mile-tons). 


The  rest  energy  m„c-  includes  the 
potential  energy,  if  there  is  any. 
Thus  a  compressed  spring  has  a 
somewhat  larger  rest  mass  and 
rest  energy  than  the  same  spring 
when  relaxed. 


But  it  takes  a  great  deal  of  kinetic  energy  to  give  a  measurable 
increase  in  mass;  the  proportionality  constant  is  very  small  — in 
fact,  Einstein  showed  it  would  be  1/c-,  where  c  is  the  speed  of  light 
in  a  vacuum: 


Am  = 


KE 


Thus  the  total  mass  tti  of  a  body  is  its  rest  mass  mo  plus  KEIc-: 


m 


KE 
c- 


Einstein  proposed  that  the  "mass  equivalent"  of  kinetic  energy 
is  only  a  special  case,  and  that  there  is  in  general  a  precise 
equivalence  between  mass  and  energy.  Thus  one  might  expect  that 
the  rest  mass  mo  also  would  correspond  to  an  equivalent  amount  of 
"rest  energy"  Eo:  mo  =  EqIc^.  That  is, 

Eo^KE 
C'        c^ 


m 


If  we  use  the  symbol  E  for  the  total  energy  of  a  body,  E  =  Eo  +  KE, 
we  could  then  write 

E 

m  =  — 

This  is  just  that  Einstein  concluded  in  1905:  "The  mass  of  a  body  is 
a  measure  of  its  energy  content."  We  can  write  this  in  a  more 


Section  20.2  99 

familiar  form,  as  what  is  probably  the  most  famous  equation  in 
physics: 

£  —  -YYiQ-i  Do  not  confuse  E  with  symbol  for 

electric  field. 
The  last  four  equations  all  represent  the  same  idea  — that  mass  and 
energy  are  different  expressions  for  the  same  characteristic  of  a 
system.  It  is  not  appropriate  to  think  of  mass  being  "converted"  to  SG  20.5,  20.6 

energy  or  vice  versa.  Rather,  a  body  with  a  measured  mass  m  has 
an  energy  E  equal  to  mc'-;  and  vice  versa  — a  body  of  total  energy  E 
has  a  mass  equal  to  £/c-. 

The  implications  of  this  equivalence  are  exciting.  First,  two  of 
the  great  conservation  laws  have  become  alternative  statements 
of  a  single  law:  in  any  system  whose  total  mass  is  conserved,  the 
total  energy  will  be  conserved  also.  Second,  the  idea  arises  that 
some  of  the  rest  energy  might  be  transformed  into  a  more  familiar 
form  of  energy.  Since  the  energy  equivalent  of  mass  is  so  great,  a 
very  small  reduction  in  rest  mass  would  be  accompanied  by  the 
release  of  a  tremendous  amount  of  energy,  for  example,  kinetic 
energy  or  electromagnetic  radiation. 

In  Chapters  23  and  24,  we  shall  see  how  such  changes  come 
about  experimentally,  and  see  additional  experimental  evidence 
which  supports  this  relationship. 

Q1     What  happens  to  the  measurable  mass  of  a  particle  as  its 
kinetic  energy  is  increased? 

Q2     What  happens  to  the  speed  of  a  particle  as  its  kinetic 
energy  is  increased? 

20.2    Particle-like  behavior  of  radiation 

We  shall  now  make  use  of  one  of  these  relations  in  the  further 
study  of  light  quanta  and  of  their  interaction  with  atoms.  Study  of 
the  photoelectric  effect  taught  us  that  a  light  quantum  has  energy 
hf,  where  h  is  Planck's  constant  and /is  the  frequency  of  the  light. 
This  concept  also  applies  to  x  rays  which,  like  visible  light,  are 
electromagnetic  radiation,  but  of  higher  frequency  than  visible 
light.  The  photoelectric  effect,  however,  did  not  tell  us  anything 
about  the  momentum  of  a  quantum.  We  may  raise  the  question:  if 
a  light  quantum  has  energy,  does  it  also  have  momentum? 

The  magnitude  of  the  momentum  ^  of  a  body  is  defined  as  the  SG  20.7 

product  of  its  mass  m  and  speed  v:  p  —  mv.  If  we  replace  m  with 
its  energy  equivalent  £/c^  we  can  write 

Note  that  the  last  equation  is  an  expression  for  the  momentum  in 
which  there  is  no  explicit  reference  to  mass.  If  we  now  speculate 
that  this  same  equation  might  define  the  momentum  of  a  photon  of 
energy  E,  v  would  be  replaced  by  the  speed  of  light  c  and  we  would 
get 

^Ec^E 
^      c^~  c 


100 


Some  Ideas  From  Modern  Physical  Theories 


SG  20.8 


X-RA/BCAH 


Foil 


Arthur  H.  Compton  (1892-1962)  was 
born  in  Wooster,  Ohio  and  graduated 
from  the  College  of  Wooster.  After  re- 
ceiving his  doctor's  degree  in  physics 
from  Princeton  University  in  1916,  he 
taught  physics  and  then  worked  in  in- 
dustry. In  1919-1920  he  did  research 
under  Rutherford  at  the  Cavendish 
Laboratory  of  the  University  of  Cam- 
bridge. In  1923,  while  studying  the 
scattering  of  x  rays,  he  discovered 
and  interpreted  the  changes  in  the 
wavelengths  of  x  rays  when  the  rays 
are  scattered.  He  received  the  Nobel 
Prize  in  1927  for  this  work. 


Now,  E^hf  for  a  light  quantum,  and  if  we  substitute  this  expres- 
sion for  E  in  p  =  Ejc,  we  would  get  the  momentum  of  a  light 
quantum: 

Or,  using  the  wave  relation  that  the  speed  equals  the  frequency 
times  the  wavelength,  c  =fK  we  could  express  the  momentum  as 

h 

Does  it  make  sense  to  define  the  momentum  of  a  photon  in  this 
way?  It  does,  if  the  definition  is  of  help  in  understanding  experi- 
mental results.  The  first  example  of  the  successful  use  of  the 
definition  was  in  the  analysis  of  an  effect  discovered  by  Arthur  H. 
Compton  which  we  will  now  consider. 

According  to  classical  electromagnetic  theory,  when  a  beam  of 
light  (or  X  rays)  strikes  the  atoms  in  a  target  (such  as  a  thin  sheet 
of  metal),  the  light  will  be  scattered  in  various  directions,  but  its 
frequency  will  not  be  changed.  The  absorption  of  light  of  a  certain 
frequency  by  an  atom  may  be  followed  by  re-emission  of  light 
of  another  frequency;  but,  if  the  light  wave  is  simple  scattered, 
then  according  to  classical  theory  there  should  be  no  change  in 
frequency. 

According  to  quantum  theory,  however,  light  is  made  up  of 
photons.  Compton  reasoned  that  if  photons  have  momentum  in 
accord  with  the  argument  for  relativity  theory,  then  in  a  collision 
between  a  photon  and  an  atom  the  law  of  conservation  of  momen- 
tum should  apply.  According  to  this  law  (see  Chapter  9),  when  a 
body  of  small  mass  collides  with  a  massive  object  at  rest,  it  simply 
bounces  back  or  glances  off  with  little  loss  in  speed  — that  is,  with 
very  little  change  in  energy.  But  if  the  masses  of  the  two  colliding 
objects  are  not  very  much  different,  a  significant  amount  of  energy 
can  be  transferred  in  the  collision.  Compton  calculated  how  much 
energy  a  photon  should  lose  in  a  collision  with  an  atom,  if  the 
momentum  of  the  photon  is  hflc.  He  concluded  that  the  change 
in  energy  is  too  small  to  observe  if  a  photon  simply  bounces  off  an 
entire  atom.  If,  however,  a  photon  strikes  an  electron,  which  has 
a  small  mass,  the  photon  should  transfer  a  significant  amount  of 
energy  to  the  electron. 

In  experiments  up  to  1923,  no  difference  has  been  observed 
between  the  frequencies  of  the  incident  and  scattered  light  (or 
X  rays)  when  electromagnetic  radiation  was  scattei'ed  by  matter.  In 
1923  Compton  was  able  to  show  that  when  a  beam  of  x  rays  is 
scattered,  the  scattered  beam  consists  of  two  parts:  one  part  has  the 
same  frequency  as  the  incident  x  rays;  the  other  part  has  slightly 
lower  frequency.  The  reduction  in  frequency  of  some  of  the 
scattered  x  rays  is  called  the  Compton  effect.  The  scattered  x  rays 
of  unchanged  frequency  have  been  scattered  by  whole  atoms, 
whereas  the  component  of  x  rays  with  changed  frequency  indicates 
a  transfer  of  energy  from  some  photons  to  electrons,  in  accordance 
with  the  laws  of  conservation  of  momentum  and  energy.  The 


Section  20.3 


101 


observed  change  in  frequency  is  just  what  would  be  predicted  if 
the  photons  were  acting  hke  particle-hke  projectiles  having 
momentum  p  =  hflc. 

Furthermore,  the  electrons  which  were  struck  by  the  photons 
could  also  be  detected,  because  they  were  knocked  out  of  the  target. 
Compton  found  that  the  momentum  of  these  electrons  was  related 
to  their  direction  in  just  the  way  that  would  be  expected  if  they 
had  been  struck  by  particles  with  momentum  equal  to  hflc. 

Compton's  experiment  showed  that  a  photon  can  be  regarded  as 
a  particle  with  a  definite  momentum  as  well  as  energy;  it  also 
showed  that  collisions  between  photons  and  electrons  obey  the  laws 
of  conservation  of  momentum  and  energy. 

Photons  are  not  like  ordinary  particles  — if  only  because  they 
do  not  exist  at  speeds  other  than  that  of  light.  (There  can  be  no 
resting  photons,  and  therefore  no  rest  mass  for  photons.)  But  in 
other  ways,  as  in  their  scattering  behavior,  photons  act  much  like 
particles  of  matter,  having  momentum  as  well  as  energy;  and  they 
also  act  like  waves,  having  frequency  and  wavelength.  In  other 
words,  the  behavior  of  electromagnetic  radiation  is  in  some  experi- 
ments similar  to  what  we  are  used  to  thinking  of  as  particle 
behavior,  and  in  other  experiments  is  similar  to  what  we  are  used 
to  thinking  of  as  wave  behavior.  This  behavior  is  often  referred  to 
as  the  wave-particle  dualism  of  radiation.  The  question,  "Is  a 
photon  a  wave  or  a  particle?"  can  only  be  answered:  it  can  act 
like  either,  depending  on  what  we  are  doing  with  it.  (This  fascinating 
topic  is  elaborated  in  several  of  the  Reader  5  articles.) 

Q3    How  does  the  momentum  of  a  photon  depend  on  the 
frequency  of  the  light? 

Q4    What  is  the  Compton  effect,  and  what  did  it  prove? 


aA/l/lAr^     • 

f 

"'    / 

/.• 

o 

(b)       \v 

¥ 

^  /1^ 

(c)  %.P 

SG  20.9 


SG  20.10 


20.3    Wave-like  behavior  of  particles 

In  1923,  the  French  physicist  Louis  de  Broglie  suggested  that 
the  wave-particle  dualism  which  applies  to  radiation  might  also 
apply  to  electrons  and  other  atomic  particles.  Perhaps,  he  said,  the 
wave-particle  dualism  is  a  fundamental  property  of  all  quantum 
processes,  and  what  we  have  always  thought  of  as  material  particles 
can,  in  some  circumstances,  act  like  waves.  He  sought  an  expres- 
sion for  the  wavelength  that  might  be  associated  with  wave-hke 
behavior  of  an  electron,  and  he  found  one  by  means  of  a  simple 
argument. 

The  momentum  of  a  photon  of  wavelength  X  is  p  =  h/X.  De 
Broglie  suggested  that  this  relation,  derived  for  photons,  would 
also  apply  to  electrons  with  the  momentum  p  =  mv.  He  therefore 
boldly  suggested  that  the  wavelength  of  an  electron  is: 

mv 
where  m  is  the  mass  of  the  electron  and  v  its  speed. 

What  does  it  mean  to  say  that  an  electron  has  a  wavelength 
equal  to  Planck's  constant  divided  by  its  momentum?  If  this 


The  "de  Broglie  wavelength"  of  a 
material  particle  does  not  refer  to 
anything  having  to  do  with  light, 
but  to  some  new  wave  property 
associated  with  the  motion  of 
matter  itself. 


102 


Some  Ideas  From  Modern  Physical  Theories 


Diffraction  pattern  produced  by  di- 
recting a  beam  of  electrons  through 
polycrystalline  aluminum.  With  a 
similar  pattern,  G.  P.  Thomson  dem- 
onstrated the  wave  properties  of 
electrons— 28  years  after  their  par- 
ticle properties  were  first  demon- 
strated by  J.  J.  Thomson,  his  father. 


statement  is  to  have  any  physical  meaning,  it  must  be  possible  to 
test  it  by  some  kind  of  experiment.  Some  wave  property  of  the 
electron  must  be  measured.  The  first  such  property  to  be  measured 
was  diffraction. 

The  relationship  X  =  himv  implies  that  the  wavelengths 
associated  with  electrons  will  be  very  short,  even  for  fairly  slow 
electrons;  an  electron  accelerated  across  a  potential  difference  of 
only  lOOV  would  have  a  wavelength  of  only  10"'"  meter.  So  small 
a  wavelength  would  not  give  noticeable  diffraction  effects  on 
encountering  any  object  of  appreciable  size  — even  microscopically 
small  size  (say,  10"^  meter). 

By  1920  it  was  known  that  crystals  have  a  regular  lattice 
structure;  the  distance  between  rows  or  planes  of  atoms  in  a  crystal 
is  about  lO"'"  m.  After  de  Broglie  proposed  his  hypothesis  that 
electrons  have  wave  properties,  several  physicists  suggested  that 
the  existence  of  electron  waves  might  be  shown  by  using  crystals  as 
diffraction  gratings.  Experiments  begun  in  1923  by  C.  J.  Davisson 
and  L.  H.  Germer  in  the  United  States,  yielded  diffraction  patterns 
similar  to  those  obtained  for  x  rays  (see  Sec.  18.6)  as  illustrated  in 
the  two  drawings  at  the  left  below.  The  experiment  showed  two 
things:  first  that  electrons  do  have  wave  properties  — one  may  say 
that  an  electron  moves  along  the  path  taken  by  the  de  Broglie  wave 
that  is  associated  with  the  electron.  Also,  it  showed  that  their 
wavelengths  are  correctly  given  by  de  Broglie's  relation,  X  =  hImv. 
These  results  were  confirmed  in  1927  by  G.  P.  Thomson,  who  directed 
an  electron  beam  through  thin  gold  foil  to  produce  a  pattern  like  the 
one  in  the  margin,  similar  to  diffraction  patterns  produced  by 
light  beams  going  through  thin  slices  of  materials.  By  1930.  diffrac- 
tion from  crystals  had  been  used  to  demonstrate  the  wave-like 
behavior  of  helium  atoms  and  hydrogen  molecules,  as  illustrated 
in  the  drawing  on  page  103. 


The  de  Broqiie  wavelength:  examples. 

A  body  of  mass  1  kg  moves  with 
a  speed  of  1  m/sec.  What  is  its 
de  Broglie  wavelength? 

An  electron  of  mass  9.1  x  IQ-^'  kg 
moves  with  a  speed  of  2  x  io^  m/sec. 
What  is  its  de  Broglie  wavelength? 

mv 

x  =  A 

mv 

rt  =  6.6  X  10  •''  joulesec 

/7  =  6.6  X  10-^'  joulesec 

mv  =  1  kg- m/sec 

mv=  1.82  X  10"-'  kg  m/sec 

6.6  X  10--"  joule.sec 
1  kgm/sec 

6.6  X  iQ--'^  joulesec 
1.82  X  10--'  kg  m/sec 

so 

so 

X  =  6.6x  10 -'^  m 

\  =  3.6  X  10"'"  m 

The  de  Broglie  wavelength  is  many 
orders  of  magnitude  smaller  than  an 
atom,  and  so  is  much  too  small  to  be 
detected— there  are,  for  example,  no 
slits  or  obstacles  small  enough  to 
show  diffraction  effects.  We  would 
expect  to  detect  no  wave  aspects  in 
the  motion  of  this  body. 

The  de  Broglie  wavelength  is  of 
atomic  dimensions;  for  example, 
it  is  of  the  same  order  of  magnitude 
as  the  distances  between  atoms  in 
a  crystal.  So  we  expect  to  see  wave 
aspects  in  the  interaction  of  elec- 
trons with  crystals. 

Section  20.3 


103 


(a) 


(b) 


oerecnsK. 


O 


X 


■■\-.f,'  ■ 


Vat£TXH. 


a.  One  way  to  demonstrate  the  wave 
behavior  of  x  rays  is  to  direct  a  beam 
at  the  surface  of  a  crystal.  The  reflec- 
tions from  different  planes  of  atoms 
in  the  crystal  interfere  to  produce 
reflected  beams  at  angles  other  than 
the  ordinary  angle  of  reflection. 


b.  A  very  similar  effect  can  be  demon- 
strated for  a  beam  of  electrons.  The 
electrons  must  be  accelerated  to  an 
energy  that  corresponds  to  a  de 
Broglie  wavelength  of  about  10"'"  m 
(which  requires  an  accelerating  volt- 
age of  only  about  100  volts). 


c.  Like  any  other  beam  of  particles, 
a  beam  of  molecules  directed  at  a 
crystal  will  show  a  similar  diffraction 
pattern.  The  diagram  above  shows 
how  a  beam  of  hydrogen  molecules 
(Ho)  can  be  formed  by  slits  at  the 
opening  of  a  heated  chamber;  the 
average  energy  of  the  molecules  is 
controlled  by  adjusting  the  tempera- 
ture of  the  oven.  The  graph,  repro- 
duced from  Zeitschrift  fur  Physik, 
1930,  shows  results  obtained  by  I. 
Estermann  and  O.  Stern  in  Germany. 
The  detector  reading  is  plotted  against 
the  deviation  to  either  side  of  the 
angle  of  ordinary  reflection. 


zso'/f 


Li 


DireMer  StrohlSSOcm. 


I    I    I 


-S 


zo'  ^o"  '5'  W  zo' 

Diffraction  pattern  for  Ha  molecules 
glancing  off  a  crystal  of  lithium 
fluoride. 


According  to  de  Broglie's  hypothesis,  which  has  been  confirmed 
by  all  experiments,  wave-particle  dualism  is  a  general  property  not 
only  of  radiation  but  also  of  matter.  It  is  now  customary  to  use  the 
word  "particle"  to  refer  to  electrons  and  photons  while  recognizing 
that  they  both  have  properties  of  waves  as  well  as  of  particles  (and, 
of  course,  that  there  are  important  differences  between  them). 

De  Broglie's  relation,  A.  =  h/mv,  has  an  interesting  yet  simple 
application  which  makes  more  reasonable  Bohr's  postulate  that  the 
quantity  mvr  (the  angular  momentum)  of  the  electron  in  the 
hydrogen  atom  can  only  have  certain  values.  Bohr  assumed  that 
mvr  can  have  only  the  values: 

h 
mvr  —  n  r—  where  n  =  1,  2,  3,  .  .  . 
2n 

Now,  suppose  that  an  electron  wave  is  somehow  spread  over  an 
orbit  of  radius  r  — that,  in  some  sense,  it  "occupies"  an  orbit  of 
radius  r.  We  may  ask  if  standing  waves  can  be  set  up  as  indicated, 
for  example,  in  the  sketch  in  the  margin.  The  condition  for  such 
standing  waves  is  that  the  circumference  of  the  orbit  is  equal  in 
length  to  a  whole  number  of  wavelengths,  that  is,  to  nX.  The 
mathematical  expression  for  this  condition  of  "fit"  is: 


SG  20.11-20.13 

Only    certain    wavelengths   will    "fit" 
around  a  circle. 


-it": 


27rr  =  nX 


104  Some  Ideas  From  Modern  Physical  Theories 

If  we  now  replace  X  by  himv  according  to  de  Broglie's  relation 
we  get 

o               ^ 
zTtr  =  n 

mv 

h 
or  mvr  =  n  -^r— 

277 

But,  this  is  just  Bohr's  quantization  condition!  The  de  BrogUe 
relation  for  electron  waves  — and  the  idea  that  the  electron  is  in 
SG  20.14       orbits  that  allow  a  standing  wave  — allows  us  to  derive  the  quantiza- 
tion that  Bohr  had  to  assume. 

The  result  obtained  indicates  that  we  may  picture  the  electron 
in  the  hydrogen  atom  in  two  ways:  either  as  a  particle  moving  in 
Either  way  is  incomplete  by  itself.  an  orbit  with  a  certain  quantized  value  of  mvr,  or  as  a  standing 

de  Broglie-type  wave  occupying  a  certain  region  around  the  nucleus. 


Q5    Where  did  de  Broghe  get  the  relation  X  =  hImv  for  electrons? 
Q6    Why  were  crystals  used  to  get  diffraction  patterns  of 
electrons? 

20.4    Mathematical  vs.  visualizable  atoms 

The  proof  that  "things"  (electrons,  atoms,  molecules)  which  had 
been  regarded  as  particles  also  show  properties  of  waves  has 
served  as  the  basis  for  the  currently  accepted  theory  of  atomic 
structure.  This  theory,  quantum  mechanics,  was  introduced  in 
1925;  its  foundations  were  developed  with  great  rapidity  during 
the  next  few  years,  primarily  by  Heisenberg,  Born,  Schrddinger, 
Bohr,  and  Dirac.  Initially  the  theory  appeared  in  two  different 
mathematical  forms,  proposed  independently  by  Heisenberg  and 
Schrodinger.  A  few  years  later,  these  two  forms  were  shown  by 
Dirac  to  be  equivalent,  different  ways  of  expressing  the  same 
relationships.  The  form  of  the  theory  that  is  closer  to  the  ideas  of 
de  Broglie.  discussed  in  the  last  section,  was  that  of  Schrodinger. 
It  is  often  referred  to  as  "wave  mechanics". 

One  of  the  fundamental  requirements  for  a  physical  theory  is 
that  it  predict  the  path  taken  by  a  particle  when  it  interacts  with 
other  particles.  It  is  possible,  as  we  have  already  indicated  for  light, 
to  write  an  equation  describing  the  behavior  of  waves  that  will 
imply  the  path  of  the  waves  — the  "rays." 

Schrodinger  sought  to  express  the  dual  wave  and  particle  nature 
of  matter  mathematically.  Maxwell  had  formulated  the  electro- 
magnetic theory  of  light  in  terms  of  a  wave  equation,  and  physicists 
were  familiar  with  this  theory  and  its  applications.  Schrodinger 
reasoned  that  the  de  Broglie  waves  associated  with  electrons  would 
resemble  the  classical  waves  of  light,  including  also  that  there  be 
a  wave  equation  that  holds  for  matter  waves  just  as  there  is  a  wave 
equation  for  electromagnetic  waves.  We  cannot  discuss  this 
mathematical  part  of  wave  mechanics  even  adequately  without 
using  an  advanced  part  of  mathematics,  but  the  physical  ideas 


Section  20.4 


105 


involved  require  only  a  little  mathematics  and  are  essential  to  an 
understanding  of  modern  physics.  So,  in  the  rest  of  this  chapter, 
we  shall  discuss  some  of  the  physical  ideas  of  the  theory  to  try  to 
make  them  seem  plausible;  and  we  shall  consider  some  of  the 
results  of  the  theory  and  some  of  the  implications  of  these  results. 
But  again  our  aim  is  not  (and  cannot  honestly  be  in  the  available 
time  and  space)  a  full  presentation.  We  want  only  to  prepare  for  the 
use  of  specific  results,  and  for  reading  in  Reader  5  and  Reader  6. 

Schrbdinger  was  successful  in  deriving  an  equation  for  the 
waves  presumed  to  "guide"  the  motion  of  electrons.  This  equation, 
which  has  been  named  after  him.  defines  the  wave  properties  of 
electrons  and  also  predicts  particle-hke  behavior.  The  Schrbdinger 
equation  for  an  electron  bound  in  an  atom  has  a  solution  only 
when  a  constant  in  the  equation  has  the  whole-number  values  1, 
2,  3.  ...  It  turns  out  that  these  numbers  correspond  to  different 
energies,  so  the  Schrodinger  equation  predicts  that  only  certain 
electron  energies  are  possible  in  an  atom.  In  the  hydrogen  atom,  for 
example,  the  single  electron  can  only  be  in  those  states  for  which 
the  energy  of  the  electron  has  the  values: 

_  2TT-mqe* 

with  n  having  only  whole  number  values.  But  these  values  of  the 
energies  are  what  are  found  experimentally  — and  are  just  the  ones 
given  by  the  Bohr  theory!  In  Schrodinger's  theory,  this  result  follows 
directly  from  the  mathematical  formulation  of  the  wave  and 
particle  nature  of  the  electron.  The  existence  of  these  stationary 
states  has  not  been  assumed,  and  no  assumptions  have  been  made 
about  orbits.  The  new  theory  yields  all  the  results  of  the  Bohr  theory 
without  having  any  of  the  inconsistent  hypotheses  of  the  earlier 
theory.  The  new  theory  also  accounts  for  the  experimental  informa- 
tion for  which  the  Bohr  theory  failed  to  account,  such  as  the  prob- 
ability of  an  electron  changing  from  one  energy  state  to  another. 

On  the  other  hand,  quantum  mechanics  does  not  supply  a 
physical  model  or  visualizable  picture  of  what  is  going  on  in  the 
world  of  the  atom.  The  planetary  model  of  the  atom  has  had  to  be 
given  up,  and  has  not  been  replaced  by  another  simple  picture. 
There  is  now  a  highly  successful  mathematical  model,  but  no  easily 
visualized  physical  model.  The  concepts  used  to  build  quantum 
mechanics  are  more  abstract  than  those  of  the  Bohr  theory;  it  is 
hard  to  get  an  intuitive  feeling  for  atomic  structure  without 
training  in  the  field.  But  the  mathematical  theory  of  quantum 
mechanics  is  much  more  powerful  than  the  Bohr  theory,  in 
predicting  and  explaining  phenomena,  and  many  problems  that  were 
previously  unsolvable  have  been  solved  with  quantum  mechanics. 
Physicists  have  learned  that  the  world  of  atoms,  electrons,  and 
photons  cannot  be  thought  of  in  the  same  mechanical  terms  as  the 
world  of  everyday  experience.  The  world  of  atoms  has  presented  us 
with  some  fascinating  concepts  which  will  be  discussed  in  the  next 
two  sections;  what  has  been  lost  in  easy  visualizability  is  amply 
made  up  for  by  the  increased  range  of  fundamental  understanding. 


Topics  in  quantum  physics  are 

developed  further  in  Reader  5. 

See  the  articles: 

"Ideas  and  Theories" 

"The  New  Landscape  of  Science" 

"The  Evolution  of  the  Physicist's 

Picture  of  Nature" 
"Dirac  and  Born" 
"I  am  the  Whole  World:  Erwin 

Schrbdinger" 
"The  Fundamental  Idea  of  Wave 

Mechanics" 
"The  Sea-Captain's  Box" 


Visualizability  is  an  unnecessary 
luxury  when  it  is  bought  at  the  cost 
of  clarity.  For  the  same  reason  we 
learned  to  do  without  visualizability 
in  many  other  fields.  For  example, 
we  no  longer  think  of  the  action  of 
an  ether  to  explain  light  propaga- 
tion. (Nor  do  we  demand  to  see 
pieces  of  silver  or  gold  or  barter 
goods  when  we  accept  a  check  as 
payment.) 


p.  A.  M.  Dirac  (1902-),  an  English  physicist,  was  one 
of  the  developers  of  modern  quantum  mechanics. 
In  1932,  at  the  age  of  30,  Dirac  was  appointed 
Lucasian  Professor  of  Mathematics  at  Cambridge 
University,  the  post  held  by  Newrton. 


Max  Born  (1882-1969)  was  born  in  Germany,  but  left  that 
country  for  England  in  1933  when  Hitler  and  the  Nazis  gained 
control.  Born  was  largely  responsible  for  introducing  the 
statistical  interpretation  of  wave  mechanics. 


Prince  Louis  Victor  de  Broglie  (1892-)  comes 
of  a  noble  French  family.  His  ancestors 
served  the  French  kings  as  far  back  as  the 
time  of  Louis  XIV.  He  was  educated  at  the 
Sorbonne  in  Paris,  and  proposed  the  idea  of 
wave  properties  of  electrons  in  his  PhD 
thesis. 


Erwin  Schrodinger  (1887-1961)  was  born  in 
Austria.  He  developed  wave  mechanics  in 
1926,  fled  from  Germany  in  1933  when  Hitler 
and  the  Nazis  came  to  power.  From  1940  to 
1956,  when  he  retired,  he  was  professor  of 
physics  at  the  Dublin  Institute  for  Advanced 
Studies. 


Werner  Heisenberg  (1 901  -).  a  german  physicist,  was  one  of  the  developers 
of  modern  quantum  mechanics  (at  the  age  of  23).  He  first  stated  the  un- 
certainty principle,  and  after  the  discovery  of  the  neutron  in  1932,  pro- 
posed the  proton-neutron  theory  of  nuclear  structure. 


108  Some  Ideas  From  Modern  Physical  Theories 

Q7    The  set  of  energy  states  of  hydrogen  could  be  derived  from 
Bohr's  postulate  that  mvr  =  nhl2TT.  In  what  respect  was  the 
derivation  from  Schrodinger's  equation  better? 

Q8    Quantum  (or  wave)  mechanics  has  had  great  success. 
What  is  its  drawback  for  those  trained  on  physical  models? 

20.5    The  uncertainty  principle 

Up  to  this  point  we  have  always  talked  as  if  we  could  measure 
any  physical  property  as  accurately  as  we  pleased;  to  reach  any 
desired  degree  of  accuracy  we  would  have  only  to  design  a 
sufficiently  precise  instrument.  Wave  mechanics  showed,  however, 
that  even  in  thought  experiments  with  ideal  instruments  there  are 
limitations  on  the  accuracy  with  which  measurements  can  be  made. 

Think  how  you  would  go  about  measuring  the  positions  and 
velocity  of  a  car  that  moves  slowly  along  a  driveway.  We  can  mark 
the  position  of  the  front  end  of  the  car  at  a  given  instant  by  making 
a  scratch  on  the  ground;  at  the  same  time,  we  start  a  stop-watch. 
Then  we  can  run  to  the  end  of  the  driveway,  and  at  the  instant  that 
the  front  end  of  the  car  reaches  another  mark  placed  on  the  ground 
we  stop  the  watch.  We  then  measure  the  distance  between  the 
marks  and  get  the  average  speed  of  the  car  by  dividing  the 
measured  distance  traversed  by  the  measured  time  elapsed.  Since 
we  know  the  direction  of  the  car's  motion,  we  know  the  average 
velocity.  Thus  we  know  that  at  the  moment  the  car  reached  the 
second  mark  it  was  at  a  certain  distance  from  its  starting  point 
and  had  traveled  at  a  certain  average  velocity.  By  the  process  of 
going  to  smaller  and  smaller  intervals  we  could  also  get  the 
instantaneous  velocity  at  any  point  along  its  path. 

How  did  we  get  the  needed  information?  We  located  the  position 
of  the  car  by  sunlight  bounced  off  the  front  end  into  our  eyes;  that 
permitted  us  to  see  when  the  car  reached  a  certain  mark  on  the 
ground.  To  get  the  average  speed  we  had  to  locate  the  front  end 
twice. 

But  suppose  that  we  had  decided  to  use  reflected  radio  waves 
instead  of  light  of  visible  wavelength.  At  1000  kilocycles  per  second, 
\  =  f  =     ^  TlO  "'^sec     gQQ  ^  ^  typical  value  for  radio  signals,  the  wavelength  is  300  meters. 


f  10'7sec 


With  radiation  of  this  wavelength,  which  is  very  much  greater  than 
the  dimensions  of  the  car,  it  is  impossible  to  locate  the  car  with 
any  accuracy.  The  wave  would  reflect  from  the  car  ("scatter"  is  a 
more  appropriate  term)  in  all  directions,  just  as  it  would  sweep 
around  any  man-sized  device  we  used  to  detect  the  wave  direction. 
The  wavelength  has  to  be  comparable  with  or  smaller  than  the 
dimensions  of  the  object  before  the  object  can  be  located  well. 
Radar  uses  wavelengths  from  about  0.1  cm  to  about  3  cm;  so  a 
radar  apparatus  could  have  been  used  instead  of  sunlight,  but 
would  leave  uncertainties  as  large  as  several  centimeters  in  the  two 
measurements  of  position.  With  visible  light  whose  wavelength  is 
less  than  10"  m,  we  could  design  instruments  that  would  locate  the 
position  of  the  car  to  an  accuracy  of  a  few  thousandths  of  a  millimeter. 


Section  20.5 


109 


The  extreme  smallness  of  the  atomic  scale  is  indicated  by  these  pictures  made 
with  techniques  that  are  near  the  very  limits  of  magnification-about  10,000,000 
times  in  these  reproductions. 


!li^'L 


^-^'.i'v,' 


Pattern  produced  by  electron  beam  scattered 
from  a  section  of  a  single  gold  crystal.  The 
entire  section  of  crystal  shown  is  only  100A 
across-smaller  than  the  shortest  wavelength 
of  ultraviolet  light  that  could  be  used  in  a  light 
microscope.  The  finest  detail  that  can  be  re- 
solved with  this  "electron  microscope"  is  just 
under  2A,  so  the  layers  of  gold  atoms  (spaced 
slightly  more  than  2A)  show  as  a  checked  pat- 
tern; individual  atoms  are  beyond  the  resolving 
power. 


Let  us  now  turn  from  car  and  driveway,  and  think  of  an  electron 
moving  across  an  evacuated  tube.  We  shall  try  to  measure  the 
position  and  speed  of  the  electron.  But  some  changes  have  to  be 
made  in  the  method  of  measurement.  The  electron  is  so  small  that 
we  cannot  locate  its  position  by  using  visible  light:  the  wavelength 
of  visible  light,  small  as  it  is,  is  still  at  least  10*  times  greater 
than  the  diameter  of  an  atom. 

To  locate  an  electron  within  a  region  the  size  of  an  atom  (about 
10~*°  m  across)  we  must  use  a  light  beam  whose  wavelength  is 
comparable  to  the  size  of  the  atom,  preferably  smaller.  Now  a 
photon  of  such  a  short  wavelength  k  (and  high  frequency/)  has 
very  great  momentum  (h/X)  and  energy  (hf);  and,  from  our  study 
of  the  Compton  effect,  we  know  that  the  photon  will  give  the 
electron  a  strong  kick  when  it  is  scattered  by  the  electron.  As  a 
result,  the  velocity  of  the  electron  will  be  greatly  changed,  into  a 
new  and  unknown  direction.  (This  is  a  new  problem,  one  we  did 
not  even  think  about  when  speaking  about  measuring  the  position 
of  the  car!)  Hence,  when  we  receive  the  scattered  photon  we  can 
deduce  from  its  direction  where  the  electron  had  been  — and  so  we 
have  "located"  the  electron.  But  in  the  process  we  have  altered  the 
velocity  of  the  electron  (in  both  magnitude  and  direction). 


Pattern  produced  by  charged  par- 
ticles repelled  from  the  tip  of  a  micro- 
scopically thin  tungsten  crystal.  The 
entire  section  shown  is  only  about 
100A  across.  The  finest  detail  that  can 
be  revealed  by  this  "field-ion  micro- 
scope" is  about  1A,  but  the  bright 
spots  indicate  the  locations  of  atoms 
along  edges  of  the  crystal,  and  should 
not  be  thought  of  as  pictures  of  the 
atoms. 


SG  20.15 


110 


Some  Ideas  From  Modern  Physical  Theories 


SG  20.16-20.18 


To  say  this  more  directly:  the  more  accurately  we  locate  the  electron 
(by  using  photons  of  shorter  wavelength)  the  less  accurately  we  can 
know  its  velocity.  We  could  try  to  disturb  the  electron  less  by  using 
less  energetic  photons.  But  because  light  exists  in  quanta  of  energy 
hf,  a  lower-enevgy  photon  will  have  a  longer  wavelength  — and 
therefore  would  give  us  greater  uncertainty  in  the  electron's  position! 

To  summarize:  we  are  unable  to  measure  both  the  position 
and  velocity  of  an  electron  to  unlimited  accuracy.  This  conclusion 
is  expressed  in  the  uncertainty  principle,  and  was  first  stated  by 
Werner  Heisenberg.  The  uncertainty  principle  can  be  expressed 
quantitatively  in  a  simple  formula,  derived  from  Schrbdinger's 
wave  equation  for  the  motion  of  particles.  If  Ax  is  the  uncertainty 
in  position,  and  Ap  is  the  uncertainty  in  momentum,  then  the 
product  of  the  two  uncertainties  must  be  equal  to,  or  greater  than, 
Planck's  constant  divided  by  27r: 

AjcAp  ^  — 

The  same  reasoning  (and  equation)  holds  for  the  experiment  on 
the  car,  but  the  limitation  is  of  no  practical  consequence  with  such 
a  massive  object.  (See  the  worked-out  example  below.)  It  is  only  on 
the  atomic  scale  that  the  limitation  becomes  evident  and  important. 


The  chief  use  made  of  the  un- 
certainty principle  is  in  general 
arguments  in  atomic  theory  rather 
than  in  particular  numerical 
problems.  We  do  not  really  need 
to  know  exactly  where  an  electron 
is,  but  we  sometimes  want  to  know 
if  it  could  be  in  some  region  of 
space. 


The  uncertainty  principle:  examples 


A  large  mass. 
Consider  a  car,  with  a  mass  of 
1000  kg,  moving  with  a  speed  of 
about  1  m/sec.  Suppose  that  in  this 
experiment  the  inherent  uncertainty 
Ai'  in  the  measured  speed  is  0.1  m/sec 
(10%  of  the  speed).  What  is  the  un- 
certainty in  the  position  of  the  car? 


A  small  mass. 
Consider  an  electron,  with  a 
mass  of  9.1  X  10"^'  kg,  moving  with 
a  speed  of  about  2  x  lO*'  m/sec. 
Suppose  that  the  uncertainty  Ai/  in 
the  speed  is  0.2  x  10*^  m/sec  (10%  of 
the  speed).  What  is  the  uncertainty  in 
the  position  of  the  electron? 


AxAp  > 


273- 


AxAp 


277 


Ap  =  mAv  =  100  kgm/sec 
h  =  6.63  X  10"''*  joulesec 


Ap  =  mlv  =  1 .82  X  10~"  kgm/sec 
h  =  6.63  X  10"'^  joule/sec 


Ax  = 
Ax 


6.63 
6.28 


lO'^-*  joulesec 


10-  kgm/sec 
1  X  io-''«  m. 


Ax 


6.63 


10'^^  joule/sec 


6.28     1.82  X  10-"  kgm/sec 

Ax  >5x  10-'"  m. 

The  uncertainty  in  position  is  of 
the  order  of  atomic  dimensions,  and 
is  significant  in  atomic  problems. 
It  is  impossible  to  specify  where 
an  electron  is  in  an  atom. 


This  uncertainty  in  position— many 
of  orders  smaller  than  the  size  of 
atoms— is  much  too  small  to  be 
observable.  In  this  case  we  can 
determine  the  position  of  the  body 
with  as  high  an  accuracy  as  we 
would  ever  need. 

The  reason  for  the  difference  between  these  two  results  is  that 
Planck's  constant  h  is  very  small;  so  small  that  the  uncertainty 
principle  becomes  important  only  on  the  atomic  scale.  Ordinary 
objects  behave  as  if,  in  the  equations  used  here,  h  is  effectively 
equal  to  zero. 


Section  20.6 


111 


Q9  If  photons  used  in  finding  the  velocity  of  an  electron 
disturb  the  electron  too  much,  why  cannot  the  observation  be 
improved  by  using  less  energetic  photons? 

Q10     If  the  wavelength  of  light  used  to  locate  a  particle  is  too 
long,  why  cannot  the  location  be  found  more  precisely  by  using 
light  of  shorter  wavelength? 


To  explore  further  the  implications  of  dualism  we  need  to 
review  some  ideas  of  probability.  Even  in  situations  in  which  no 
single  event  can  be  predicted  with  certainty,  it  may  still  be  possible 
to  make  predictions  of  the  statistical  probabilities  of  certain  events. 
On  a  holiday  weekend  during  which  perhaps  25  million  cars  are  on 
the  road,  the  statisticians  report  a  high  probability  that  about  600 
people  will  be  killed  in  accidents.  It  is  not  known  which  cars  in 
which  of  the  50  states  will  be  the  ones  involved  in  the  accidents, 
but  on  the  basis  of  past  experience  the  average  behavior  is  still 
quite  accurately  predictable. 

It  is  in  this  way  that  physicists  think  about  the  behavior  of 
photons  and  material  particles.  As  we  have  seen,  there  are 
fundamental  limitations  on  our  ability  to  describe  the  behavior 
of  an  individual  particle.  But  the  laws  of  physics  often  enable  us 
to  describe  the  behavior  of  large  collections  of  particles  with  good 
accuracy.  The  solutions  of  Schrodinger's  wave  equations  for  the 
behavior  of  waves  associated  with  particles  give  us  the  probabilities 
for  finding  the  particles  at  a  given  place  at  a  given  time. 

To  see  how  probability  fits  into  the  picture,  consider  the 
situation  of  a  star  being  photographed  through  a  telescope.  As  you 
have  already  seen  (for  example  on  the  page  on  Diffraction  and  Detail 
in  Chapter  13),  the  image  of  a  point  source  is  not  a  precise  point 
but  is  a  diffraction  pattern  — a  central  spot  with  a  series  of 
progressively  fainter  circular  rings. 

The  image  of  a  star  on  the  photographic  film  in  the  telescope 
would  be  a  similar  pattern.  Imagine  now  that  we  wished  to 
photograph  a  very  faint  star.  If  the  energy  in  light  rays  were  not 
quantized,  but  spread  continuously  over  ever-expanding  wave 
fronts,  we  would  expect  that  the  image  of  a  very  faint  star  would 
be  exactly  the  same  as  that  of  a  much  brighter  star  — except  that 
the  intensity  of  light  would  be  less  over  the  whole  pattern.  However, 
the  energy  of  light  is  quantized  — it  exists  in  separate  quanta, 
"photons,"  of  definite  energy.  When  a  photon  strikes  a  photographic 
emulsion,  it  produces  a  chemical  change  in  the  film  at  a  single 
location -not  all  over  the  image  area.  If  the  star  is  very  remote, 
only  a  few  photons  per  second  may  arrive  at  the  film.  The  effect  on 
the  film  after  a  very  short  period  of  exposure  would  not  be  at  all 
like  the  diffraction  pattern  in  drawing  C  in  the  margin,  but 
something  like  the  scatter  in  A.  As  the  exposure  continued,  the 
effect  on  the  film  would  begin  to  look  like  B.  Eventually,  a  pattern 
like  C  would  be  produced,  just  like  the  image  produced  by  a  bright 
star  with  a  much  shorter  exposure. 


These  sketches  represent  greatly  en- 
larged images  of  a  distant  star  on  a 
photographic  plate. 


112 


Some  Ideas  From  Modern  Physical  Theories 


As  we  have  already  discussed  in 
connection  with  l<inetic  theory  and 
disorder,  it  is  easy  to  predict  the 
average  behavior  of  very  large 
numbers  of  particles,  even  though 
nothing  at  all  is  known  about  the 
behavior  of  any  single  one  of  them. 


If  there  are  tremendous  numbers  of  quanta,  then  then'  overall 
distribution  will  be  very  well  described  by  the  distribution  of  wave 
intensity.  For  small  numbers  of  quanta,  the  wave  intensity  will  not 
be  very  useful  for  predicting  where  they  will  go.  We  expect  them  to 
go  mostly  to  the  "high-intensity"  parts  of  the  image  but  we  cannot 
predict  exactly  where.  These  facts  fit  together  beautifully  if  we 
consider  the  wave  intensity  at  a  location  to  indicate  the  probability 
of  the  photon  going  there! 

A  similar  connection  can  be  made  for  de  Broglie  waves  and 
particles  of  matter.  Rather  than  considering  an  analogous  example, 
such  as  a  diffraction  pattern  formed  by  an  electron  beam,  we  can 
consider  a  bound  electron  wave  — a  wave  confined  to  a  region  of 
space  by  the  electric  attraction  of  a  positive  nucleus  and  a  negative 
electron.  For  example,  the  de  Broglie  wave  associated  with  an 
electron  is  spread  out  all  over  an  atom  — but  we  need  not  think  of 
the  electron  as  spread  out.  It  is  quite  useful  to  think  of  the  electron 
as  a  particle  moving  around  the  nucleus,  and  the  wave  amplitude 
at  some  location  represents  the  probability  of  the  electron  being 
there. 

According  to  modern  quantum  theory,  the  hydrogen  atom  does 
not  consist  of  a  localized  negative  particle  moving  around  a  nucleus 
as  in  the  Bohr  model.  Indeed,  the  theory  does  not  provide  any  picture 
of  the  hydrogen  atom.  A  description  of  the  probability  distribution 
is  the  closest  thing  that  the  theory  provides  to  a  picture.  The  proba- 
bility distribution  for  the  lowest  energy  state  of  the  hydrogen  atom 
is  represented  in  the  drawing  at  the  left  below,  where  whiter 
shading  at  a  point  indicates  greater  probability.  The  probability 
distribution  for  a  higher  energy  state,  still  for  a  single  electron,  is 
represented  in  the  drawing  at  the  right. 

Quantum  theory  is,  however,  not  really  concerned  with  the 
position  of  any  individual  electron  in  any  individual  atom.  Instead, 
the  theory  gives  a  mathematical  representation  that  can  be  used  to 
predict  interaction  with  particles,  fields,  and  radiation.  For  example, 
it  can  be  used  to  calculate  the  probability  that  hydrogen  will  emit 
light  of  a  particular  wavelength;  the  intensity  and  wavelength  of 
light  emitted  by  a  large  number  of  hydrogen  atoms  can  then  be 
compared  with  these  calculations.  Comparisons  such  as  these  have 
shown  that  the  theory  agrees  with  experiment. 


Section  20.6 


113 


To  understand  atomic  physics,  we  deal  with  the  average 
behavior  of  many  atomic  particles;  the  laws  governing  this  average 
behavior  turn  out  to  be  those  of  wave  mechanics.  The  waves,  it 
seems,  are  waves  that  measure  probability.  The  information  about 
the  probability  (that  a  particle  will  have  some  position  at  a  given 
time)  travels  through  space  in  waves.  These  waves  can  interfere 
with  each  other  in  exactly  the  same  way  that  water  waves  do.  So, 
for  example,  if  we  think  of  a  beam  of  electrons  passing  through 
two  slits,  we  consider  the  electrons  to  be  waves  and  compute  the 
interference  patterns  which  determine  the  directions  in  which  there 
are  high  wave  amplitudes  (high  probability  of  electrons  going 
there).  Then,  as  long  as  there  are  no  more  slits  or  other  interactions 
of  the  waves  with  matter,  we  can  return  to  our  description  in  terms 
of  particles  and  say  that  the  electrons  are  likely  to  (and  on  the 
average  will)  end  up  going  in  such  and  such  directions  with  such 
and  such  speeds. 

The  success  of  wave  mechanics  emphasized  the  importance  of 
the  dual  wave-and-particle  nature  of  radiation  and  matter.  But  it  is 
natural  to  ask  how  a  particle  can  be  thought  of  as  "really"  having 
wave  properties.  The  answer  is  that  matter,  particularly  on  the 
scale  of  the  atom,  does  not  have  to  be  thought  of  as  being  either 
"really"  particles  or  "really"  waves.  Our  ideas  of  waves  and  of 
particles  are  taken  from  the  world  of  visible  things  and  just  do  not 
apply  on  the  atomic  scale. 

When  we  try  to  describe  something  that  no  one  has  ever  seen  or 
can  ever  see  directly,  it  would  be  surprising  if  the  concepts  of  the 
visible  world  could  be  used  unchanged.  It  appeared  natural  before 
1925  to  try  to  talk  about  the  transfer  of  energy  in  either  wave  terms 
or  particle  terms,  because  that  was  all  physicists  needed  or  knew 
at  the  time.  Almost  no  one  was  prepared  to  find  that  both  wave  and 
particle  descriptions  could  apply  to  light  and  to  matter.  But  as  long 
as  our  imagination  and  language  has  only  these  two  ideas  — waves 
and  particles  — to  stumble  along  on,  this  dualism  cannot  be  wished 
away;  it  is  the  best  way  to  handle  experimental  results. 

Max  Born,  one  of  the  founders  of  quantum  mechanics,  has 
written : 


The  ultimate  origin  of  the  difficulty  lies  in  the  fact  (or 
philosophical  principle)  that  we  are  compelled  to  use  the 
words  of  common  language  when  we  wish  to  describe  a 
phenomenon,  not  by  logical  or  mathematical  analysis,  but 
by  a  picture  appealing  to  the  imagination.  Common 
language  has  grown  by  everyday  experience  and  can 
never  surpass  these  limits.  Classical  physics  has  restricted 
itself  to  the  use  of  concepts  of  this  kind;  by  analyzing 
visible  motions  it  has  developed  two  ways  of  representing 
them  by  elementary  processes:  moving  particles  and 
waves.  There  is  no  other  way  of  giving  a  pictorial  descrip- 
tion of  motions  — we  have  to  apply  it  even  in  the  region  of 
atomic  processes,  where  classical  physics  breaks  down. 


See  "Dirac  and  Born"  in  Reader  5. 


Despite  the  successes  of  the  idea  that  the  wave  represents  the 


114 


Some  Ideas  From  Modern  Physical  Theories 


probability  of  finding  its  associated  particle  in  some  specific  condi- 
SG  20.23        tion  of  motion,  many  scientists  found  it  hard  to  accept  the  idea 

that  men  cannot  know  exactly  what  any  one  particle  is  doing.  The 
most  prominent  of  such  disbelievers  was  Einstein.  In  a  letter  to 
Born  written  in  1926,  he  remarked: 

The  quantum  mechanics  is  very  imposing.  But  an  inner 
voice  tells  me  that  it  is  still  not  the  final  truth.  The  theory 
yields  much,  but  it  hardly  brings  us  nearer  to  the  secret 
of  the  Old  One.  In  any  case,  I  am  convinced  that  He  does 
not  play  dice. 


"Deterministic"  means  here  that  if 
all  the  conditions  of  an  isolated 
system  are  known  and  the  laws 
describing  interaction  are  known, 
then  it  is  possible  to  predict 
precisely  what  happens  next,  without 
any  need  for  probability  ideas. 


SG  20.19-20.23 


Thus,  Einstein,  while  agreeing  with  the  usefulness  and  success 
of  wave  mechanics  so  interpreted,  refused  to  accept  probability- 
based  laws  as  the  final  level  of  explanation  in  physics;  in  the 
remark  about  not  believing  that  God  played  dice  — an  expression  he 
used  many  times  later  — he  expressed  his  faith  that  there  are  more 
basic,  deterministic  laws  yet  to  be  found.  Yet  despite  the  refusal  of 
Einstein  (and  some  others)  to  accept  the  probability  laws  in 
mechanics,  neither  he  nor  other  physicists  have  yet  succeeded  in 
replacing  Born's  probability  interpretation  of  quantum  mechanics. 

Scientists  agree  that  quantum  mechanics  works;  its  gives  the 
right  answers  to  many  questions  in  physics,  it  unifies  ideas  and 
occurrences  that  were  once  unconnected,  and  it  has  been  wonder- 
fully productive   of  new  experiments  and  new  concepts.  On  the 
other  hand,  there  is  still  vigorous  argument  about  its  basic 
significance.  It  yields  probability  functions,  not  precise  trajectories. 
Some  scientists  see  in  this  aspect  of  the  theory  an  important 
revelation  about  the  nature  of  the  world;  for  other  scientists  this 
same  fact  indicates  that  quantum  theory  is  incomplete.  Some  in 
this  second  group  are  trying  to  develop  a  more  basic,  non-statistical 
theory  for  which  the  present  quantum  theory  is  only  a  limiting  case. 
As  in  other  fields  of  physics,  the  greatest  discoveries  here  may  be 
those  yet  to  be  made. 

Q11     In  wave  terms,  the  bright  lines  of  a  diffraction  pattern  are 
regions  where  there  is  a  high  field  intensity  produced  by  constructive 
interference.  What  is  the  probability  interpretation  of  quantum 
mechanics  for  the  bright  lines  of  a  diffraction  pattern? 

Q12     If  quantum  mechanics  can  predict  only  probabilities  for 
the  behavior  of  any  one  particle,  how  can  it  predict  many 
phenomena,  for  example,  half-lives  and  diffraction  patterns,  with 
great  certainty? 


"Sea  and  Sky",  by  M.  C.  Escher 


Models  of  the  Atom 

EPILOGUE      In  this  unit  we  have  traced  the  concept  of  the  atom  from 
the  early  ideas  of  the  Greeks  to  the  quantum  mechanics  now  generally 
accepted  by  physicists.  The  search  for  the  atom  started  with  the 
qualitative  assumptions  of  Leucippus  and  Democritus  who  thought 
that  their  atoms  offered  a  rational  explanation  of  the  behavior  of 
matter.  For  many  centuries  most  natural  philosophers  thought  that 
other  explanations,  not  involving  atoms,  were  more  reasonable. 
Atomism  was  pushed  aside  and  received  only  occasional  consideration 
until  the  seventeenth  century.  With  the  growth  of  the  mechanical 
philosophy  of  nature  in  the  seventeenth  and  eighteenth  centuries, 
particles  (corpuscles)  became  important.  Atomism  was  reexamined, 
mostly  in  connection  with  physical  properties  of  matter.  Galileo,  Boyle, 
Newton  and  others  speculated  on  the  role  of  particles  for  explaining  the 
expansion  and  contraction  of  gases.  Chemists  speculated  about  atoms 
in  connection  with  chemical  change.  Finally,  Dalton  began  the  modern 
development  of  atomic  theory,  introducing  a  quantitative  conception 
that  had  been  lacking  — the  relative  atomic  mass. 

Chemists,  in  the  nineteenth  century,  found  that  they  could  correlate 
the  results  of  many  chemical  experiments  in  terms  of  atoms  and 
molecules.  They  also  found  that  there  are  relations  between  the 
properties  of  different  chemical  elements.  Quantitative  information 
about  atomic  masses  provided  a  framework  for  the  system  organizing 
these  relations  — the  periodic  table  of  Mendeleev.  During  the  nineteenth 
century,  physicists  developed  the  kinetic  theory  of  gases.  This  theory- 
based  on  the  assumption  of  very  small  corpuscles,  or  particles,  or 


Epilogue 


117 


molecules,  or  whatever  else  they  might  be  called  — helped  strengthen 
the  position  of  the  atomists.  Other  work  of  nineteenth-century  physicists 
helped  pave  the  way  to  the  study  of  the  structure  of  atoms-through 
the  study  of  the  spectra  of  the  elements  and  of  the  conduction  of 
electricity  in  gases,  the  discovery  of  cathode  rays,  electrons,  and 
X  rays. 

Nineteenth-century  chemistry  and  physics  converged,  at  the 
beginning  of  the  twentieth  century,  on  the  problem  of  atomic  structure. 
It  became  clear  that  the  uncuttable,  infinitely  hard  atom  was  too  simple 
a  model:  that  the  atom  itself  is  made  up  of  smaller  particles.  And  so  the 
search  for  a  model  with  structure  began.  Of  the  early  models,  that  of 
Thomson  gave  way  to  Rutherford's  nuclear  atom,  with  its  small,  heavy, 
positively  charged  nucleus,  surrounded  somehow  by  negative  charges. 
Then  came  the  atom  of  Bohr,  with  its  electrons  thought  to  be  moving  in 
orbits  like  planets  in  a  miniature  solar  system.  The  Bohr  theory  had 
many  successes  and  linked  chemistry  and  spectra  to  the  physics  of 
atomic  structure.  But  beyond  that,  it  could  not  advance  substantially 
without  giving  up  an  easily  grasped  picture  of  the  atom.  The  tool 
needed  is  the  mathematical  model,  not  pictures.  Quantum  mechanics 
enables  us  to  calculate  how  atoms  behave;  it  helps  us  understand  the 
physical  and  chemical  properties  of  the  elements.  But  at  the  most  basic 
level,  nature  still  has  secrets. 

The  next  stage  in  our  story.  Unit  6,  is  the  nucleus  at  the  center  of 
the  atom.  Is  the  nucleus  made  up  of  smaller  components?  Does  it  have 
laws  of  physics  all  its  own? 


o®©o® 

©  (D  ©  (D)  © 
OO©©© 

©®©®o 


20.1     The  Project  Physics  materials  particularly 
appropriate  for  Chapter  20  include: 


and  show  that  KE  =  jnioV^  is  a  good 
approximation  for  familiar  objects. 


Activities 

Standing  waves  on  a  band-saw  blade 
Turntable  oscillator  patterns  resembling 

de  Broglie  waves 
Standing  waves  in  a  wire  ring 

Reader  Articles 

The  Clock  Paradox 

Ideas  and  Theories 

Mr.  Tompkins  and  Simultaneity 

Mathematics  and  Relativity 

Parable  of  the  Surveyors 

Outside  and  Inside  the  Elevator 

Einstein  and  Some  Civilized  Discontents 

The  New  Landscape  of  Science 

The  Evolution  of  the  Physicist's  Picture 

of  Nature 
Dirac  and  Born 

I  am  the  Whole  World:  Erwin  Schrodinger 
The  Fundamental  Idea  of  Wave  Mechanics 
The  Sea-Captain's  Box 
Space  Travel:  Problems  of  Physics  and 

Engineering 
Looking  for  a  New  Law 

20.2  How  fast  would  you  have  to  move  to 
increase  your  mass  by  1%? 

20.3  The  centripetal  force  on  a  mass  moving 
with  relativistic  speed  v  around  a  circular  orbit  of 
radius  R  is  F  =  mv^lR,  where  m  is  the  relativistic 
mass.  Electrons  moving  at  a  speed  0.60  c  are  to 
be  deflected  in  a  circle  of  radius  1.0  m:  what 
must  be  the  magnitude  of  the  force  applied? 

(mo  =  9.1  X  lO-»'  kg.) 

20.4  The  formulas  (p  =  ruoV,  KE  =  jtnov'^)  used  in 
Newtonian  physics  are  convenient  approxima- 
tions to  the  more  general  relativistic  formulas. 
The  factor  1/Vl  -  v^lc'^  can  be  expressed  as  an 
infinite  series  of  steadily  decreasing  terms  by 
using  a  binomial  series  expansion.  When  this  is 
done  we  find  that 


1 


V 


1-^ 


V'  t;"  v" 

=  1  +  1/2  ^  -f  3/8  -^  +  5/16  ^  + 
c^  c*  c® 


35/128-^ 

c* 


20.5  According  to  relativity  theory,  changing 
the  energy  of  a  system  by  AE  also  changes  the 
mass  of  the  system  by  Am  =  A£/c-.  Something 
like  10^  joules  per  kilogram  of  substance  are 
commonly  released  as  heat  energy  in  chemical 
reactions. 

(a)  Why  then  aren't  mass  changes  detected  in 
chemical  reactions? 

(b)  Calculate  the  mass  change  associated  with 
a  change  of  energy  of  10^  joules. 

20.6  The  speed  of  the  earth  in  its  orbit  is  about 
18  miles/sec  (3  x  10^  m/sec).  Its  "rest"  mass  is 
6.0  X  102"  kg 

(a)  What  is  the  kinetic  energy  of  the  earth  in 
its  orbit? 

(b)  What  is  the  mass  equivalent  of  that  kinetic 
energy? 

(c)  By  what  percentage  is  the  earth's  "rest" 
mass  increased  at  orbital  speed? 

(d)  Refer  back  to  Unit  2  to  recall  how  the  mass 
of  the  earth  is  found;  was  it  the  rest  mass 
or  the  mass  at  orbital  speed? 

20.7  In  relativistic  mechanics  the  formula 
^=  mi/^  still  holds,  but  the  mass  m  is  given  by 
m  =  mo/Vi  -  v'^lc~.  The  rest  mass  of  an  electron 
is  9.1  X  lO-'"  kg. 

(a)  What  is  its  momentum  when  it  is  moving 
down  the  axis  of  a  linear  accelerator  from 
left  to  right  at  a  speed  of  0.4  c  with 
respect  to  the  accelerator  tube? 

(b)  What  would  Newton  have  calculated  for 
the  momentum  of  the  electron? 

(c)  By  how  much  would  the  relativistic  momen- 
tum increase  if  the  speed  of  the  electron 
were  doubled? 

(d)  What  would  Newton  have  calculated  its 
change  in  momentum  to  be? 

20.8  Calculate  the  momentum  of  a  photon  of 
wavelength  4000A.  How  fast  would  an  electron 
have  to  move  in  order  to  have  the  same 
momentum? 


(a)  Show,  by  simple  substitution,  that  when  — 

is  less  than  0.1,  the  values  of  the  terms 
drop  off  so  rapidly  that  only  the  first  few 
terms  need  be  considered. 

(b)  We  rarely  observe  familiar  objects  moving 
faster  than  about  3,000  m/sec;  the  speed  of 
light  is  3  X  10"  m/sec,  so  the  value  of  v/c 
for  familiar  objects  is  rarely  greater  than 
about  10  •'*.  What  error  do  we  suffer  by  using 
only  the  first  two  terms  of  the  series? 

(c)  Substitute  the  first  two  terms  of  the  series 
into  the  relativistic  expression 


20.9  Construct  a  diagram  showing  the  change 
that  occurs  in  the  frequency  of  a  photon  as  a 
result  of  its  collision  with  an  electron. 

20.10  What  explanation  would  you  offer  for  the 
fact  that  the  wave  aspect  of  light  was  shown  to 
be  valid  before  the  particle  aspect  was  demon- 
strated? 

20.11  The  electrons  which  produced  the  diffrac- 
tion photograph  on  p.  102  had  de  Broglie 
wavelengths  of  10""*  meter.  To  what  speed  must 
they  have  been  accelerated?  (Assume  that  the 


118 


speed  is  small  compared  to  c,  so  that  the  electron 
mass  is  about  10"^"  kg.) 

20.12  A  bilhard  ball  of  mass  0.2  kilograms 
moves  with  a  speed  of  1  meter  per  second.  What 
is  its  de  Broghe  wavelength? 

20.13  Show  that  the  de  Broglie  wavelength  of  a 
classical  particle  of  mass  m  and  kinetic  energy 
KE  is  given  by 

h 


V2Tn(K£) 

What  happens  when  the  mass  is  very  small  and 
the  speed  is  very  great? 

20.14    A  particle  confined  in  a  box  cannot  have 
a  kinetic  energy  less  than  a  certain  amount;  this 
least  amount  corresponds  to  the  longest  de  Broghe 
wavelength  which  produces  standing  waves  in 
the  box;  that  is,  the  box  size  is  one-half  wave- 
length. For  each  of  the  following  situations  find 
the  longest  de  Broglie  wavelength  that  would  fit 
in  the  box:  then  use  p  =  hi K  to  find  the  momen- 
tum p,  and  use  p  =  mv  to  find  the  speed  v. 

(a)  a  dust  particle  (about  lO-^*  kg)  in  a  display 
case  (about  1  m  across). 

(b)  an  argon  atom  (6.6  x  10"-'*  kg)  in  a  hght 
bulb  (about  10~*  m  across). 

(c)  a  protein  molecule  (about  10"--  kg)  in  a 
bacterium  (about  10"®  m  across). 


(d)  an  electron  (about  10"^' 
(about  10"'"  m  across). 


kg)  in  an  atom 


20.15  Suppose  that  the  only  way  you  could  obtain 
information  about  the  world  was  by  throwing 
rubber  balls  at  the  objects  around  you  and 
measuring  their  speeds  and  directions  of  rebound. 
What  kind  of  objects  would  you  be  unable  to 
learn  about? 

20.16  A  bullet  can  be  considered  as  a  particle 
having  dimensions  approximately  1  centimeter. 
It  has  a  mass  of  about  10  grams  and  a  speed  of 
about  3x10^  centimeters  per  second.  Suppose 
we  can  measure  its  speed  to  an  accuracy  of 

±1  cm/sec.  What  is  the  corresponding  uncertainty 
in  its  position  according  to  Heisenberg's 
principle? 

20.17  Show  that  if  Planck's  constant  were  equal 
to  zero,  quantum  effects  would  disappear  and 
even  atomic  particles  would  behave  according 

to  Newtonian  physics.  What  effect  would  this 
have  on  the  properties  of  light? 

20.18  Some  writers  have  claimed  that  the  un- 
certainty principle  proves  that  there  is  free  will. 
Do  you  think  this  extrapolation  from  atomic 
phenomena  to  the  world  of  animate  beings  is 
justified? 

20.19  A  physicist  has  written 


It  is  enough  that  quantum  mechanics  predicts 
the  average  value  of  observable  quantities 
correctly.  It  is  not  really  essential  that  the 
mathematical  symbols  and  processes  corre- 
spond to  some  intelligible  physical  picture  of 
the  atomic  world. 

Do  you  regard  such  a  statement  as  acceptable? 

Give  reasons. 

20.20  In  Chapters  19  and  20  we  have  seen  that 

it  is  impossible  to  avoid  the  wave-particle  duahsm 
of  light  and  matter.  Bohr  has  coined  the  word 
complementarity  for  the  situation  in  which  two 
opposite  views  seem  valid,  and  the  correct  choice 
depends  only  on  which  aspect  of  a  phenomenon 
one  chooses  to  consider.  Can  you  think  of  situa- 
tions in  other  fields  (outside  of  atomic  physics) 
to  which  this  idea  might  apply? 

20.21  In  Units  1  through  4  we  discussed  the 
behavior  of  large-scale  "classical  particles"  (for 
example,  tennis  balls)  and  "classical  waves" 
(for  example,  sound  waves),  that  is,  of  particles 
and  waves  that  in  most  cases  can  be  described 
without  any  use  of  ideas  such  as  the  quantum  of 
energy  or  the  de  Broghe  matter-wave.  Does  this 
mean  that  there  is  one  sort  of  physics  ("classical 
physics")  for  the  phenomena  of  the  large-scale 
world  and  quite  a  different  physics  ("quantum 
physics")  for  the  phenomena  of  the  atomic  world? 
Or  does  it  mean  that  quantum  physics  really 
applies  to  all  phenomena  but  is  not  distinguish- 
able from  classical  physics  when  applied  to  large- 
scale  particles  and  waves?  What  arguments  or 
examples  would  you  use  to  defend  your  answer? 

20.22  If  there  are  laws  that  describe  precisely 
the  behavior  of  atoms,  it  can  be  inferred  that  the 
future  is  completely  determined  by  the  present 
(and  the  present  was  determined  in  the  ancient 
past).  This  idea  of  complete  determinism^  was 
uncomfortable  to  many  philosophers  during  the 
centuries  following  the  great  success  of 
Newtonian  mechanics.  The  great  French  physi- 
cist Pierre  Laplace  (1748-1827)  wrote. 

Given  for  one  instant  an  intelligence  which 
could  comprehend  all  the  forces  by  which 
nature  is  animated  and  the  respective 
situation  of  the  beings  who  compose  it  — an 
intelhgence  sufficiently  vast  to  submit  these 
data  to  analysis  — it  would  embrace  in  the 
same  formula  the  movements  of  the  greatest 
bodies  of  the  universe  and  those  of  the 
hghtest  atom;  for  it,  nothing  would  be  un- 
certain and  the  future,  as  the  past,  would  be 
present  to  its  eyes  [A  Philosophical  Essay  on 
Probabilities.] 

Is  this  statement  consistent  with  modem 

physical  theory? 

20.23  (The  later  statistical  view  of  kinetic  theory 
may  have  emphasized  the  difficulty  of  actually 
predicting  the  future,  but  did  not  weaken  the 
idea  of  an  underlying  chain  of  cause  and  effect.) 

(a)  What  implications  do  you  see  in  relativity 
theorv  for  the  idea  of  determinism? 


119 


STUDY  G 


(b)  What  implications  do  you  see  for  determinism 
in  quantum  theory? 

20.24    Those  ancient  Greeks  who  beUeved  in 
natural  law  were  also  troubled  by  the  idea  of 
determinism.  How  do  the  Greek  ideas  expressed 
in  the  following  passage  from  Lucretius'  On  the 
Nature  of  Things  (about  80  B.C.)  compare  with 
modern  physics  ideas? 
If  cause  forever  follows  after  cause 
In  infinite,  undeviating  sequence 
And  a  new  motion  always  has  to  come 
Out  of  an  old  one,  by  fixed  law;  if  atoms 


Do  not,  by  swerving,  cause  new  moves  which 

break 

The  laws  of  fate;  if  cause  forever  follows, 

In  infinite  sequence,  cause  — where  would  we 

get 

This  free  will  that  we  have,  wrested  from  fate  . . . 

What  keeps  the  mind  from  having  inside  itself 

Some  such  compulsiveness  in  all  its  doings. 

What  keeps  it  from  being  matter's  absolute 

slave? 

The  answer  is  that  our  free-will  derives 

From  just  that  ever-so-slight  atomic  swerve 

At  no  fixed  time,  at  no  fixed  place  whatever. 


120 


» 


I 


Kenneth  Ford,  University  of  California,  Irvine 
Robert  Gardner,  Harvard  University 
Fred  Geis,  Jr.,  Harvard  University 
Nicholas  J.  Georgis,  Staples  High  School, 

Westport,  Conn. 
H.  Richard  Gerfin,  Somers  Middle  School, 

Somers,  N.Y. 
Owen  Gingerich,  Smithsonian  Astrophysical 

Observatory,  Cambridge,  Mass. 
Stanley  Goldberg,  Antioch  College,  Yellow  Springs, 

Ohio 
Leon  Goutevenier,  Paul  D.  Schreiber  High  School, 

Port  Washington,  N.Y. 
Albert  Gregory,  Harvard  University 
Julie  A.  Goetze,  Weeks  Jr.  High  School,  Newton, 

Mass. 
Robert  D.  Haas,  Clairemont  High  School,  San 
Diego,  Calif. 

Walter  G.  Hagenbuch,  Plymouth- Whitemarsh 

Senior  High  School,  Plymouth  Meeting,  Pa. 

John  Harris,  National  Physical  Laboratory  of 

Israel,  Jerusalem 
Jay  Hauben,  Harvard  University 
Peter  Heller,  Brandeis  University,  Waltham,  Mass. 
Robert  K.  Henrich,  Kennewick  High  School, 

Washington 
Ervin  H.  HofFart,  Raytheon  Education  Co.,  Boston 
Banesh  Hoffmann,  Queens  College,  Flushing,  N.Y. 
Elisha  R.  Huggins,  Dartmouth  College,  Hanover, 

N.H. 
Lloyd  Ingraham,  Grant  High  School,  Portland, 

Ore. 
John  Jared,  John  Rennie  High  School,  Pointe 

Claire,  Quebec 
Harald  Jensen,  Lake  Forest  College,  111. 
John  C.  Johnson,  Worcester  Polytechnic  Institute, 

Mass. 
Kenneth  J.  Jones,  Harvard  University 
LeRoy  Kallemeyn,  Benson  High  School,  Omaha, 

Neb. 
Irving  Kaplan,  Massachusetts  Institute  of 

Technology,  Cambridge 
Benjamin  Karp,  South  Philadelphia  High  School, 

Pa. 
Robert  Katz,  Kansas  State  University,  Manhattan, 

Kans. 
Harry  H.  Kemp,  Logan  High  School,  Utah 
Ashok  Khosla,  Harvard  University 
John  Kemeny,  National  Film  Board  of  Canada, 

Montreal 
Merritt  E.  Kimball,  Capuchino  High  School,  San 

Bruno,  Calif. 
Walter  D.  Knight,  University  of  California, 

Berkeley 
Donald  Kreuter,  Brooklyn  Technical  High  School, 

N.Y. 
Karol  A.  Kunysz,  Laguna  Beach  High  School, 

Calif. 
Douglas  M.  Lapp,  Harvard  University 
Leo  Lavatelli,  University  of  Illinois,  Urbana 


122 


Joan  Laws,  American  Academy  of  Arts  and 

Sciences.  Boston 
Alfred  Leitner,  Michigan  State  University,  East 

Lansing 
Robert  B.  Lillich.  Solon  High  School.  Ohio 
James  Lindblad,  Lowell  High  School,  Whittier 

Calif. 
Noel  C.  Little,  Bowdoin  College,  Brunswick,  Me. 
Arthur  L.  Loeb,  Ledgemont  Laboratory,  Lexington, 

Mass. 
Richard  T.  Mara.  Gettysburg  College,  Pa. 
Robert  H.  Maybury,  UNESCO,  Paris 
John  McClain,  University  of  Beirut,  Lebanon 
E.  Wesley  McNair,  W.  Charlotte  High  School. 

Charlotte,  N.C. 
William  K.  Mehlbach,  Wheat  Ridge  High  School, 

Colo. 
Priya  N.  Mehta.  Harvard  University 
Glen  Mervyn,  West  Vancouver  Secondary  School, 

B.C.,  Canada 
Franklin  Miller,  Jr..  Kenyon  College,  Gambler 

Ohio 
Jack  C.  Miller,  Pomona  College,  Claremont.  Calif. 
Kent  D.  Miller,  Claremont  High  School,  Calif. 
James  A.  Minstrell,  Mercer  Island  High  School, 

Washington 
James  F.  Moore,  Canton  High  School,  Mass. 
Robert  H.  Mosteller,  Princeton  High  School, 

Cincinnati,  Ohio 
William  Naison,  Jamaica  High  School.  N.Y. 
Henry  Nelson,  Berkeley  High  School,  Calif. 
Joseph  D.  Novak,  Purdue  University,  Lafayette 

Ind. 
Thorir  Olafsson,  Menntaskolinn  Ad,  Laugarvatni, 

Iceland 
Jay  Orear,  Cornell  University,  Ithaca,  N.Y. 
Paul  O'Toole.  Dorchester  High  School,  Mass. 
Costas  Papaliolios,  Harvard  University 
Jacques  Parent,  National  Film  Board  of  Canada. 

Montreal 
Father  Thomas  Pisors,  C.S.U.,  Griffin  High 

School,  Springfield,  111. 
Eugene  A.  Platten,  San  Diego  High  School,  Calif. 
L.  Eugene  Poorman,  University  High  School, 

Bloomington,  Ind. 
Gloria  Poulos,  Harvard  University 
Herbert  Priestley,  Knox  College,  Galesburg,  111. 
Edward  M.  Purcell,  Harvard  University 
Gerald  M.  Rees,  Ann  Arbor  High  School,  Mich. 
James  M.  Reid.  J.  W.  Sexton  High  School, 

Lansing,  Mich. 
Robert  Resnick,  Rensselaer  Polytechnic  Institute, 

Troy,  NY. 
Paul  I.  Richards,  Technical  Operations,  Inc., 

Burlington,  Mass. 
John  Rigden,  Eastern  Nazarene  College,  Quincy, 

Mass. 
Thomas  J.  Ritzinger,  Rice  Lake  High  School,  Wise. 
Nickerson  Rogers,  The  Loomis  School,  Windsor, 
Conn. 

(Continued  on  page  167) 


I 


The  Project  Physics  Course 


Models  of  the  Atom 


Contents 


HANDBOOK  SECTION 


Chapter  17    The  Chemical  Basis  of  Atomic  Theory 

Experiment 

40.  Electrolysis     126 

Activities 

Dalton's  Puzzle     129 
Electrolysis  of  Water     129 
Periodic  Table     129 
Single-electrode  Plating     131 
Activities  from  Scientific  American     1 31 

Film  Loop 

Film  Loop  46 :  Production  of  Sodium  by  Electrolysis     1 32 

Chapter  18     Electrons  and  Quanta 

Experiments 

41.  The  Charge-to-mass  Ratio  for  an  Electron     133 

42.  TheMeasurementof  Elementary  Charge     136 

43.  The  Photoelectric  Effect     139 

Activities 

Writings  By  or  About  Einstein     143 
Measuring  Q/M  for  the  Electron     143 
Cathode  Rays  in  a  Crookes  Tube     1 43 
X-rays  from  a  Crookes  Tube     143 
Lighting  an  Electric  Lamp  with  a  Match     1 43 

Film  Loop 

Film  Loop  47:  Thomson  Model  of  the  Atom     145 

Chapter  19    The  Rutherford-Bohr  Model  of  the  Atom 

Experiment 

44.  Spectroscopy     146 

Activities 

Scientists  on  Stamps     149 

Measuring  Ionization,  a  Quantum  Effect     148 

Modeling  Atoms  with  Magnets     150 

"Black  Box"  Atoms     151 

Another  Simulation  of  the  Rutherford  Atom     1 52 

Film  Loop 

Film  Loop  48 :  Rutherford  Scattering     153 

Chapter  20     Some  Ideas  from  Modern  Physical  Theories 

Activities 

Standing  Waves  on  a  Band-saw  Blade     154 

Turntable  Oscillator  Patterns  Resembling  de  Broglie  Waves     154 

Standing  Waves  in  a  Wire  Ring     154 


17 


Chapter    I  f    The  Chemical  Basis  of  Atomic  Theory 


EXPERIMENT  40  ELECTROLYSIS 

Volta  and  Davy  discovered  that  electric  cur- 
rents created  chemical  changes  never  observed 
before.  As  you  have  already  learned,  these 
scientists  were  the  first  to  use  electricity  to 
break  down  apparently  stable  compounds  and 
to  isolate  certain  chemical  elements. 

Later  Faraday  and  other  experimenters 
compared  the  amount  of  electric  charge  used 
with  the  amount  of  chemical  products  formed 
in  such  electrochemical  reactions.  Their  mea- 
surements fell  into  a  regular  pattern  that 
hinted  at  some  underlying  link  between  elec- 
tricity and  matter. 

In  this  experiment  you  will  use  an  electric 
current  just  as  they  did  to  decompose  a  com- 
pound. By  comparing  the  charge  used  with  the 
mass  of  one  of  the  products,  you  can  compute 
the  mass  and  volume  of  a  single  atom  of  the 
product. 

Theory  Behind  the  Experiment 

A  beaker  of  copper  sulfate  (CUSO4)  solution  in 
water  is  supported  under  one  arm  of  a  balance 
(Fig.  17-1).  A  negatively  charged  copper  elec- 
trode is  supported  in  the  solution  by  the  bal- 


ance arm  so  that  you  can  measure  its  mass 
without  removing  it  from  the  solution.  A  sec- 
ond, positively  charged  copper  electrode  fits 
around  the  inside  wall  of  the  beaker.  The 
beaker,  its  solution  and  the  positive  electrode 
are  not  supported  by  the  balance  arm. 

If  you  have  studied  chemistry,  you  proba- 
bly know  that  in  solution  the  copper  sulfate 
comes  apart  into  separate  charged  particles, 
called  ions,  of  copper  (Cu++)  and  sulfate  (S04=), 
which  move  about  freely  in  the  solution. 

When  a  voltage  is  applied  across  the  cop- 
per electrodes,  the  electric  field  causes  the 
S04=  ions  to  drift  to  the  positive  electrode  (or 
anode)  and  the  Cu++  ions  to  drift  to  the  nega- 
tive electrode  (or  cathode).  At  the  cathode  the 
Cu++  particles  acquire  enough  negative  charge 
to  form  neutral  copper  atoms  which  deposit 
on  the  cathode  and  add  to  its  weight.  The  mo- 
tion of  charged  particles  toward  the  electrodes 
is  a  continuation  of  the  electric  current  in  the 
wires  and  the  rate  of  transfer  of  charge  (cou- 
lombs per  second)  is  equal  to  it  in  magnitude. 
The  electric  current  is  provided  by  a  power  sup- 
ply that  converts  100- volt  alternating  current 
into  low-voltage  direct  current.  The  current 


Fig.  17-1 


Experiment  40 


127 


is  set  by  a  variable  control  on  the  power  supply 
(or  by  an  external  rheostat)  and  measured  by 
an  ammeter  in  series  with  the  electrolytic  cell 
as  shown  in  Fig.  17-1. 

With  the  help  of  a  watch  to  measure  the 
time  the  current  flows,  you  can  compute  the 
electric  charge  that  passed  through  the  cell. 
By  definition,  the  current  I  is  the  rate  of  trans- 
fer or  charge:  I  =  AQ/At.  It  follows  that  the 
charge  transferred  is  the  product  of  the  cur- 
rent and  the  time. 


AQ  =  I  X  At 

coulombs 
(coulombs  = 


sec 


X  sec) 


Since  the  amount  of  charge  carried  by  a 
single  electron  is  known  (qe  =  1.6  x  10"*^  cou- 
lombs), the  number  of  electrons  transferred, 
Ne,  is 


If  n  electrons  are  needed  to  neutralize  each 
copper  ion,  then  the  number  of  copper  atoms 
deposited,  N^u,  is 


N     =^ 
n 


'Cu 


If  the  mass  of  each  copper  atom  is  rric^,  then 
the   total  mass  of  copper  deposited,  M^u,  is 

Mcu  =  ^curricu 

Thus,  if  you  measure  I,  At  and  Mf.„,  and  you 
know  q^  and  n,  you  can  calculate  a  value  for 
nicu,  the  mass  of  a  single  copper  atom! 

Setup  and  Procedure 

Either  an  equal-arm  or  a  triple-beam  balance 
can  be  used  for  this  experiment.  First  arrange 
the  cell  and  the  balance  as  shown  in  Fig.  17-1. 
The  cathode  cylinder  must  be  supported  far 
enough  above  the  bottom  of  the  beaker  so  that 
the  balance  arm  can  move  up  and  down  freely 
when  the  cell  is  full  of  the  copper  sulfate 
solution. 

Next  connect  the  circuit  as  illustrated  in 


the  figure.  Note  that  the  electrical  connection 
from  the  negative  terminal  of  the  power  supply 
to  the  cathode  is  made  through  the  balance 
beam.  The  knife-edge  and  its  seat  must  be  by- 
passed by  a  short  piece  of  thin  flexible  wire, 
as  shown  in  Fig.  17-1  for  equal-arm  balances, 
or  in  Fig.  17-2  for  triple-beam  balances.  The 
positive  terminal  of  the  power  supply  is  con- 
nected directly  to  the  anode  in  any  convenient 
manner. 


Fig.  17-2  This  cutaway  view  shows  how  to  by-pass  the 
knife-edge  of  a  typical  balance.  The  structure  of  other 
balances  may  differ. 


Before  any  measurements  are  made,  op- 
erate the  cell  long  enough  (10  or  15  minutes)  to 
form  a  preliminary  deposit  on  the  cathode— 
unless  this  has  already  been  done.  In  any  case, 
run  the  current  long  enough  to  set  it  at  the 
value  recommended  by  your  teacher,  probably 
about  5  amperes. 

When  all  is  ready,  adjust  the  balance  and 
record  its  reading.  Pass  the  current  for  the 
length  of  time  recommended  by  your  teacher. 
Measure  and  record  the  current  I  and  the  time 
interval  At  during  which  the  current  passes. 
Check  the  ammeter  occasionally  and,  if  neces- 
sary, adjust  the  control  in  order  to  keep  the 
current  set  at  its  original  value. 

At  the  end  of  the  run,  record  the  new  read- 
ing of  the  balance,  and  find  by  subtraction  the 
increase  in  mass  of  the  cathode. 


128 


Experiment  40 


Calculating  Mass  and  Volume  of  an  Atom 

Since  the  cathode  is  buoyed  up  by  a  Uquid,  the 
masses  you  have  measured  are  not  the  true 
masses.  Because  of  the  buoyant  force  exerted 
by  the  Uquid,  the  mass  of  the  cathode  and  its 
increase  in  mass  will  both  appear  to  be  less 
than  they  would  be  in  air.  To  find  the  true  mass 
increase,  you  must  divide  the  observed  mass 
increase  by  the  factor  (1  -  DglD,.),  where  Dg  is 
the  density  of  the  solution  and  D^.  is  the  density 
of  the  copper. 

Your  teacher  will  give  you  the  values  of 
these  two  densities  if  you  cannot  find  values 
for  them  yourself.  He  will  also  explain  how  the 
correction  factor  is  derived.  The  important 
thing  for  you  to  understand  here  is  why  a  cor- 
rection factor  is  necessary. 
Ql  How  much  positive  or  negative  charge 
was  transferred  to  the  cathode? 


In  the  solution  this  positive  charge  is  car- 
ried from  anode  to  cathode  by  doubly  charged 
copper  ions,  Cu++.  At  the  cathode  the  copper 
ions  are  neutralized  by  electrons  and  neutral 
copper  atoms  are  deposited:  Cu^+  +  2e"Cu. 
Q2  How  many  electrons  were  required  to 
neutralize  the  total  charge  transferred?  (Each 
electron  carries  -1.6  x  10"'"  coulombs.) 
Q3  How  many  electrons  (single  negative 
charge)  were  required  to  neutralize  each  cop- 
per ion? 

Q4  How  many  copper  atoms  were  deposited? 
Q5  What  is  the  mass  of  each  copper  atom? 
Q6  The  mass  of  a  penny  is  about  3  grams.  If 
it  were  made  of  copper  only,  how  many  atoms 
would  it  contain?  (In  fact  modern  pennies  con- 
tain zinc  as  well  as  copper.) 
Q7  The  volume  of  a  penny  is  about  0.3  cm^ 
How  much  volume  does  each  atom  occupy? 


ACTIVITIES 


129 


DALTON'S  PUZZLE 

Once  Dalton  had  his  theory  to  work  with,  the 
job  of  figuring  out  relative  atomic  masses  and 
empirical  formulas  boiled  down  to  nothing 
more  than  working  through  a  series  of  puzzles. 
Here  is  a  very  similar  kind  of  puzzle  with 
which  you  can  challenge  your  classmates. 

Choose  three  sets  of  objects,  each  having  a 
different  mass.  Large  ball  bearing  with  masses 
of  about  70,  160,  and  200  grams  work  well.  Let 
the  smallest  one  represent  an  atom  of  hydro- 
gen, the  middle-sized  one  an  atom  of  nitrogen, 
and  the  large  one  an  atom  of  oxygen. 

From  these  "atoms"  construct  various 
"molecules."  For  example,  NHg  could  be  repre- 
sented by  three  small  objects  and  one  middle- 
sized  one,  N2O  by  two  middle-sized  ones  and 
one  large,  and  so  forth. 

Conceal  one  molecule  of  your  collection  in 
each  one  of  a  series  of  covered  Styrofoam  cups 
(or  other  hght-weight,  opaque  containers). 
Mark  on  each  container  the  symbols  (but  not 
the  formula!)  of  the  elements  contained  in  the 
compound.  Dalton  would  have  obtained  this 
information  by  quahtative  analysis. 

Give  the  covered  cups  to  other  students. 
Instruct  them  to  measure  the  "molecular" 
mass  of  each  compound  and  to  deduce  the  rela- 
tive atomic  masses  and  empirical  formulas 
from  the  set  of  masses,  making  Dalton's  as- 
sumption of  simplicity.  If  the  objects  you  have 
used  for  "atoms"  are  so  hght  that  the  mass  of 
the  styrofoam  cups  must  be  taken  into  account, 
you  can  either  supply  this  information  as  part 
of  the  data  or  leave  it  as  a  comphcation  in  the 
problem. 

If  the  assumption  of  simphcity  is  relaxed, 
what  other  atomic  masses  and  molecular 
formulas  would  be  consistent  with  the  data? 

ELECTROLYSIS  OF  WATER 

The  fact  that  electricity  can  decompose  water 
was  an  amazing  and  exciting  discovery,  yet 
the  process  is  one  that  you  can  easily  demon- 
strate with  materials  at  your  disposal.  Fig. 
17-3  provides  all  the  necessary  information. 
Set  up  an  electrolysis  apparatus  and  demon- 
strate the  process  for  your  classmates. 

In  Fig.  17-3  it  looks  as  if  about  twice  as 


Fig.  17-3 


many  bubbles  were  coming  from  one  electrode 
as  from  the  other.  Which  electrode  is  it?  Does 
this  happen  in  your  apparatus?  Would  you 
expect  it  to  happen? 

How  would  you  collect  the  two  gases  that 
bubble  off  the  electrodes?  How  could  you 
prove  their  identity? 

If  water  is  really  just  these  two  gases  "put 
together"  chemically,  you  should  be  able  to 
put  the  gases  together  again  and  get  back  the 
water  with  which  you  started.  Using  your 
knowledge  of  physics,  predict  what  must  then 
happen  to  all  the  electrical  energy  you  sent 
flowing  through  the  water  to  separate  it. 

PERIODIC  TABLE 

You  may  have  seen  one  or  two  forms  of  the 
periodic  table  in  your  classroom,  but  many 
others  have  been  devised  to  emphasize  var- 
ious relationships  among  the  elements.  Some, 
such  as  the  ones  shown  on  the  next  page,  are 
more  visually  interesting  than  others.  Check 
various  sources  in  your  library  and  prepare  an 
exhibit  of  the  various  types.  An  especially  good 
lead  is  the  article,  "Ups  and  Down  of  the  Per- 
iodic Table"  in  Chemistry,  July  1966,  which 
shows  many  different  forms  of  the  table,  in- 
cluding those  in  Fig.  17-4. 

It  is  also  interesting  to  arrange  the  ele- 
ments in  order  of  discovery  on  a  Unear  time 
chart.  Periods  of  intense  activity  caused  by 
breakthroughs  in  methods  of  extended  work  by 
a  certain  group  of  investigators  show  up  in 
groups  of  names.  A  simple  way  to  do  this  is  to 
use  a  typewriter,  letting  each  Une  represent 
one  year  (from  1600  on).  All  the  elements  then 
fit  on  six  normal  typing  pages  which  can  be 


130  Activities 


(c) 


Three  two-dimensional  spiral  forms,  (a)  Janet.  1928.  (b)  Kipp,  1942.  (c)  Sibaiua,  1941. 


Activities 


131 


fastened  together  for  mounting  on  a  wall.  A 
list  of  discovery  dates  for  all  elements  appears 
at  the  end  of  Chapter  21  in  the  Text. 

SINGLE-ELECTRODE  PLATING 

A  student  asked  if  copper  would  plate  out  from 
a  solution  of  copper  sulfate  if  only  a  negative 
electrode  were  placed  in  the  solution.  It  was 
tried  and  no  copper  was  observed  even  when 
the  electrode  was  connected  to  the  negative 
terminal  of  a  high  voltage  source  for  five 
minutes.  Another  student  suggested  that  only 
a  very  small  (invisible)  amount  of  copper  was 
deposited  since  copper  ions  should  be  attracted 
to  a  negative  electrode. 

A  more  precise  test  was  devised.  A  nickel- 
sulfate  solution  was  made  containing  several 
microcuries  of  radioactive  nickel  (no  radio- 
copper  was  available).  A  single  carbon  elec- 
trode was  immersed  in  the  solution,  and  con- 
nected to  the  negative  terminal  of  the  high 
voltage  source  again  for  five  minutes.  The 
electrode  was  removed,  dried,  and  tested  with 
a  Geiger  counter.  The  rod  was  slightly  radio- 
active. A  control  test  was  run  using  identical 
test  conditions,  except  that  no  electrical  con- 
nection was  made  to  the  electrode.  The  control 
showed  more  radioactivity. 

Repeat  these  experiments  and  see  if  the 
effect  is  true  generally.  What  explanation 
would  you   give  for  these  effects?  (Adapted 


from  Ideas  for  Science  Investigations,  N.  S.- 
T.  A.  1966). 

ACTIVITIES  FROM  SCIENTIFIC  AMERICAN 

The    following    articles    from    the    "Amateur 
Scientist"  section  of  Scientific  American  re- 
late to  Unit  5.  They  range  widely  in  difficulty. 
Accelerator,  electron,  Jan.  1959,  p.  138. 
Beta  ray  spectrometer,  Sept.  1958,  p.  197. 
Carbon  14  dating,  Feb.  1957,  p.  159. 
Cloud  chamber,  diffusion,  Sept.  1952,  p.  179. 
Cloud  chamber,  plumber's  friend,  Dec.  1956, 
p.  169. 

Cloud  chamber,  Wilson,  Apr.  1956,  p.  156. 
Cloud  chamber,  with  magnet,  June  1959, 
p.  173. 

Cyclotron,  Sept.  1953,  p.  154. 
Gas  discharge  tubes,  how  to  make,  Feb,  1958, 
p.  112. 

Geiger  counter,  how  to  make.  May  1960,  p.  189. 
Isotope  experiments.  May  1960,  p.  189. 
Magnetic  resonance  spectrometer,  Apr.  1959, 
p.  171. 

Scintillation  counter,  Mar.  1953,  p.  104. 
Spectrograph,  astronomical,  Sept.  1956,  p.  259. 
Spectrograph,  Bunsen's,  June  1955,  p.  122. 
Spinthariscope,  Mar.  1953,  p.  104. 
SpectroheUograph,  how  to  make,  Apr.  1958, 
p.  126. 

Subatomic  particle  scattering,  simulating, 
Aug.  1965,  p.  102. 


FILM  LOOP 


FILM  LOOP  46:  PRODUCTION  OF  SODIUM 
BY  ELECTROLYSIS 

In  1807,  Humphry  Davy  produced  metallic 
sodium  by  electrolysis  of  molten  lye — sodium 
hydroxide. 

In  the  film,  sodium  hydroxide  (NaOH)  is 
placed  in  an  iron  crucible  and  heated  until  it 
melts,  at  a  temperature  of  318°C.  A  rectifier 
connected  to  a  power  transformer  supphes  a 
steady  current  through  the  Uquid  NaOH 
through  iron  rods  inserted  in  the  melt.  Sodium 
ions  are  positive  and  are  therefore  attracted 
to  the  negative  electrode;  there  they  pick  up 
electrons  and  become  metalHc  sodium,  as  in- 
dicated symbohcally  in  this  reaction: 

Na+  +  e-  =  Na. 


The  sodium  accumulates  in  a  thin,  shiny  layer 
floating  on  the  surface  of  the  molten  sodium 
hydroxide. 

Sodium  is  a  dangerous  material  which 
combines  explosively  with  water.  The  experi- 
menter in  the  film  scoops  out  a  little  of  the 
metal  and  places  it  in  water.  (Fig.  17-5.)  En- 
ergy is  released  rapidly,  as  you  can  see  from 
the  violence  of  the  reaction.  Some  of  the  so- 
dium is  vaporized  and  the  hot  vapor  emits  the 
yellow  hght  characteristic  of  the  spectrum  of 
sodium.  The  same  yellow  emission  is  easily 
seen  if  common  salt,  sodium  chloride,  or  some 
other  sodium  compound,  is  sprinkled  into  an 
open  flame. 


Fig.  17-5 


Chapter 


18 


Electrons  and  Quanta 


EXPERIMENT  41  THE  CHARGE-TO-MASS 
RATIO  FOR  AN  ELECTRON 

In  this  experiment  you  make  measurements  on 
cathode  rays.  A  set  of  similar  experiments  by 
J.  J.  Thomson  convinced  physicists  that  these 
rays  are  not  waves  but  streams  of  identical 
charged  particles,  each  with  the  same  ratio  of 
charge  to  mass.  If  you  did  experiment  38  in 
Unit  4,  "Electron-Beam  Tube,"  you  have  al- 
ready worked  with  cathode  rays  and  have  seen 
how  they  can  be  deflected  by  electric  and 
magnetic  fields. 

Thomson's  use  of  this  deflection  is  des- 
cribed on  page  36  of  the  Unit  5  Text.  Read 
that  section  of  the  text  before  beginning  this 
experiment. 


radius  R  by  a  uniform  magnetic  field  B,  the 
centripetal  force  rm/IR  on  each  electron  is 
supplied  by  the  magnetic  force  Bq^v.  Therefore 


R 


Bq,v, 


or,  rearranging  to  get  v  by  itself, 

m 

The  electrons  in  the  beam  are  accelerated 
by  a  voltage  V  which  gives  them  a  kinetic 
energy 


mv^ 


Theory  of  the  experiment 

The  basic  plan  of  the  experiment  is  to  measure 
the  bending  of  the  electron  beam  by  a  known 
magnetic  field.  From  these  measurements  and 
a  knowledge  of  the  voltage  accelerating  the 
electrons,  you  can  calculate  the  electron 
charge-to-mass  ratio.  The  reasoning  behind 
the  calculation  is  illustrated  in  Fig.  18-1.  The 
algebraic  steps  are  described  below. 


yqe- 


If  you  replace  v  in  this  equation  by  the  expres- 
sion for  V  in  the  preceding  equation,  you  get 


m  _  (BqeR 


or,  after  simpHfying, 


yqe 


m 


2V 

Bm^ 


Fig.  18-1  The  combination  of  two  relationships,  for 
centripetal  and  kinetic  energy,  with  algebraic  steps  that 
eliminate  velocity,  v,  lead  to  an  equation  for  the  charge- 
to-mass  ratio  of  an  electron. 


When  the  beam  of  electrons  (each  of  mass 
m  and  charge  <je)  is  bent  into  a  circular  arc  of 


You  can  measure  with  your  apparatus  all 
the  quantities  on  the  right-hand  side  of  this 
expression,  so  you  can  use  it  to  calculate  the 
charge-to-mass  ratio  for  an  electron. 

Preparing  the  apparatus 

You  will  need  a  tube  that  gives  a  beam  at  least 
5  cm  long.  If  you  kept  the  tube  you  made  in 
Experiment  38,  you  may  be  able  to  use  that. 
If  your  class  didn't  have  success  with  this 
experiment,  it  may  mean  that  your  vacuum 
pump  is  not  working  well  enough,  in  which 
case  you  will  have  to  use  another  method. 


134 


Experiment  41 


In  this  experiment  you  need  to  be  able  to 
adjust  the  strength  of  the  magnetic  field  until 
the  magnetic  force  on  the  charges  just  bal- 
ances the  force  due  to  the  electric  field.  To 
enable  you  to  change  the  magnetic  field,  you 
will  use  a  pair  of  coils  instead  of  permanent 
magnets.  A  current  in  a  pair  of  coils,  which 
are  separated  by  a  distance  equal  to  the  coil 
radius,  produces  a  nearly  uniform  magnetic 
field  in  the  central  region  between  the  coils. 
You  can  vary  the  magnetic  field  by  changing 
the  current  in  the  coils. 

Into  a  cardboard  tube  about  3"  in  diameter 
and  3"  long  cut  a  slot  I4"  wide.  (Fig.  18-2.) 
Your  electron-beam  tube  should  fit  into  this 
slot  as  shown  in  the  photograph  of  the  com- 
pleted set-up.  (Fig.  18-4.)  Current  in  the  pair 
of  coils  will  create  a  magnetic  field  at  right 
angles  to  the  axis  of  the  cathode  rays. 

Now  wind  the  coils,  one  on  each  side  of  the 
slot,  using  a  single  length  of  insulated  copper 
wire  (magnet  wire).  Wind  about  20  turns  of 
wire  for  each  of  the  two  coils,  one  coil  on  each 
side  of  the  slot,  leaving  10"  of  wire  free  at  both 
ends  of  the  coil.  Don't  cut  the  wire  off  the  reel 
until  you  have  found  how  much  you  will  need. 
Make  the  coils  as  neat  as  you  can  and  keep 
them  close  to  the  slot.  Wind  both  coils  in  the 
same  sense  (for  example,  make  both  clock- 
wise). 

When  you  have  made  your  set  of  coils,  you 
must  "calibrate"  it;  that  is,  you  must  find  out 
what  magnetic  field  strength  B  corresponds 
to  what  values  of  current  I  in  the  coils.  To  do 


Fig.     18-2 


Fig.     18-3 


this,  you  can  use  the  current  balance,  as  you 
did  in  Experiment  36.  Use  the  shortest  of  the 
balance  "loops"  so  that  it  will  fit  inside  the 
coils  as  shown  in  Fig.  18-3. 

Connect  the  two  leads  from  your  coils  to 
a  power  supply  capable  of  giving  up  to  5  amps 
direct  current.  There  must  be  a  varable  con- 
trol on  the  power  supply  (or  a  rheostat  in  the 
circuit)  to  control  the  current;  and  an  ammeter 
to  measure  it. 

Measure  the  force  F  for  a  current  /  in  the 
loop.  To  calculate  the  magnetic  field  due  to 
the  current  in  the  coils,  use  the  relationship 
F  =  BU  where  i  is  the  length  of  short  section  of 
the  loop.  Do  this  for  several  different  values  of 
current  in  the  coil  and  plot  a  calibration  graph 
of  magnetic  field  B  against  coil  current  I. 

Set  up  your  electron-beam  tube  as  in  Ex- 
periment 38.  Reread  the  instructions  for  oper- 
ating the  tube. 

Connect  a  shorting  wire  between  the  pins 
for  the  deflecting  plates.  This  will  insure  that 
the  two  plates  are  at  the  same  electric  poten- 
tial, so  the  electric  field  between  them  will  be 
zero.  Pump  the  tube  out  and  adjust  the  fila- 
ment current  until  you  have  an  easily  visible 
beam.  Since  there  is  no  field  between  the 
plates,  the  electron  beam  should  go  straight 
up  the  center  of  the  tube  between  the  two 
plates.  (If  it  does  not,  it  is  probably  because 
the  filament  and  the  hole  in  the  anode  are  not 
properly  aligned.) 

Turn  down  the  filament  current  and  switch 
off  the  power  supply.  Now,  without  releasing 


Experiment  41 


135 


Fig.  18-4  The  magnetic  field  is  parallel  to  the  axis  of 
the  coils;  the  electric  and  magnetic  fields  are  perpen- 
dicular to  each  other  and  to  the  electron  beam. 

the  vacuum,  mount  the  coils  around  the  tube 
as  shown  in  Fig.  18-4. 

Connect  the  coils  as  before  to  the  power 
supply.  Connect  a  voltmeter  across  the  power 
supply  terminals  that  provide  the  accelerating 
voltage  V. 

Your  apparatus  is  now  complete. 

Performing  the  experiment 

Turn  on  the  beam,  and  make  sure  it  is  travel- 
hng  in  a  straight  line.  The  electric  field  re- 
mains off  throughout  the  experiment,  and  the 
deflecting  plates  should  still  be  connected 
together. 

Turn  on  and  slowly  increase  the  current  in 
the  coils  until  the  magnetic  field  is  strong 
enough  to  deflect  the  electron  beam  noticeably. 

Record  the  current  I  in  the  coils. 

Using  the  cahbration  graph,  find  the  mag- 
netic field  B. 

Record  the  accelerating  voltage  V  between 
the  filament  and  the  anode  plate. 

Finally  you  need  to  measure  R,  the  radius 
of  the  arc  into  which  the  beam  is  bent  by  the 
magnetic  field.  The  deflected  beam  is  sUghtly 
fan-shaped  because  some  electrons  are  slowed 
by  collisions  with  air  molecules  and  are  bent 
into  a  curve  of  smaller  R.  You  need  to  know  the 
largest  value  of  R  (the  "outside"  edge  of  the 
curved  beam),  which  is  the  path  of  electrons 
that  have  made  no  collisions.  You  won't  be 
able  to  measure  R  directly,  but  you  can  find 


heary  herft  tn1b 
I  Circuhor   arc 


tane-fi 


!    peATperfCdCLihr 


f? 


.X 


d 


\ 


\ 


\ 


f?-x. 


Fig.    18-5 


it  from  measurements  that  are  easy  to  make. 
(Fig.  18-5.) 

You  can  measure  x  and  d.  It  follows  from 
Pythagoras'  theorem  that  R^  =  d^  +  (R  —  xf, 


so  R  = 


d^  +  x^ 
2x     ' 


Ql     What  is  your  calculation  of  R  on  the  basis 
of  your  measurements? 

Now  that  you  have  values  for  V,  B  and  R, 
you  can  use  the  formula  qelm=  2VIB^R'^  to  cal- 
culate your  value  for  the  charge-to-mass  ratio 
for  an  electron. 

Q2    What  is  your  value  for  Qelm,  the  charge- 
to-mass  ratio  for  an  electron? 


136 


EXPERIMENT  42  THE  MEASUREMENT  OF 
ELEMENTARY  CHARGE 

In  this  experiment,  you  will  investigate  the 
charge  of  the  electron,  a  fundamental  physical 
constant  in  electricity,  electromagnetism,  and 
nuclear  physics.  This  experiment  is  substan- 
tially the  same  as  Millikan's  famous  oil-drop 
experiment,  described  on  page  39  of  the  Unit  5 
Text.  The  following  instructions  assume  that 
you  have  read  that  description.  Like  Milhkan, 
you  are  going  to  measure  very  small  electric 
charges  to  see  if  there  is  a  limit  to  how  small 
an  electric  charge  can  be.  Try  to  answer  the 
following  three  questions  before  you  begin  to 
do  the  experiment  in  the  lab. 
Ql  What  is  the  electric  field  between  two 
parallel  plates  separated  by  a  distance  d  me- 
ters, if  the  potential  difference  between  them 
is  V  volts? 

Q2  What  is  the  electric  force  on  a  particle 
carrying  a  charge  of  q  coulombs  in  an  electric 
field  of  E  volts/meter? 

Q3  What  is  the  gravitational  force  on  a  par- 
ticle of  mass  m  in  the  earth's  gravitational 
field? 

Background 

Electric  charges  are  measured  by  measuring 
the  forces  they  experience  and  produce.  The 
extremely  small  charges  that  you  are  seeking 
require  that  you  measure  extremely  small 
forces.  Objects  on  which  such  small  forces 
can  have  a  visible  effect  must  also  in  turn  be 
very  small. 

Millikan  used  the  electrically  charged 
droplets  produced  in  a  fine  spray  of  oil.  The 
varying  size  of  the  droplets  comphcated  his 
measurements.  Fortunately  you  can  now  use 
suitable  objects  whose  sizes  are  accurately 
known.  You  use  tiny  latex  spheres  (about 
10"^  cm  diatmeter),  which  are  almost  identical 
in  size  in  any  given  sample.  In  fact,  these 
spheres,  shown  magnified  (about  5000  x)  in 
Fig.  18-6,  are  used  as  a  convenient  way  to  find 
the  magnifying  power  of  electron  microscopes. 
The  spheres  can  be  bought  in  a  water  suspen- 
sion, with  their  diameter  recorded  on  the 
bottle.  When  the  suspension  is  sprayed  into  the 
air,  the  water  quickly  evaporates  and  leaves 


Fig.  18-6  Electron  micrograph  of  latex  spheres  1.1  x 
lO'^cm,  silhouetted  against  diffracting  grating  of  28,800 
lines/inch.  What  magnification  does  this  represent? 

a  cloud  of  these  particles,  which  have  become 
charged  by  friction  during  the  spraying.  In 
the  space  between  the  plates  of  the  Millikan 
apparatus  they  appear  through  the  50-power 
microscope  as  bright  points  of  hght  against 
a  dark  background. 

You  will  find  that  an  electric  field  between 
the  plates  can  pull  some  of  the  particles  up- 
ward against  the  force  of  gravity,  so  you  will 
know  that  they  are  charged  electrically. 

In  your  experiment,  you  adjust  the  voltage 
producing  the  electric  field  until  a  particle 
hangs  motionless.  On  a  balanced  particle 
carrying  a  charge  q,  the  upward  electric  force 
Eq  and  the  downward  gravitational  force  mUg 
are  equal,  so 

mUg  =  Eq . 

The  field  E  =  VId,  where  V  is  the  voltage 
between  the  plates  (the  voltmeter  reading) 
and  d  is  the  separation  of  the  plates.  Hence 


q  = 


mOgd 
V 


Notice  that  mUgd  is  a  constant  for  all 
measurements  and  need  be  found  only  once. 
Each  value  of  q  will  be  this  constant  mUgd 
times  1/V  as  the  equation  above  shows.  That 
is,  the  value  of  q  for  a  particle  is  proportional 
to  1/V:  the  greater  the  voltage  required  to  bal- 
ance the  weight  of  the  particle,  the  smaller 
the  charge  of  the  particle  must  be. 


Experiment  42 


137 


Fig.  18-7    A  typical  set  of  apparatus.  Details  may  vary 
considerably. 


Using  the  apparatus 

If  the  apparatus  is  not  already  in  operating 
condition,  consult  your  teacher.  Study  Figs. 
18-7  and  18-8  until  you  can  identify  the  various 
parts.  Then  switch  on  the  hght  source  and 
look  through  the  microscope.  You  should  see 
a  series  of  Unes  in  clear  focus  against  a  uni- 
form gray  background. 


to  chamber 


to  vo  Hmtttr      f^\      \\' 


Confrc 


Fig.  18-8    A  typical  arrangement  of  connections  to  the 
high-voltage  reversing  switch. 

The  lens  of  the  hght  source  may  fog  up  as 
the  heat  from  the  lamp  drives  moisture  out  of 


the  hght-source  tube.  If  this  happens,  remove 
the  lens  and  wipe  it  on  a  clean  tissue.  Wait 
for  the  tube  to  warm  up  thoroughly  before 
replacing  the  lens. 

Squeeze  the  bottle  of  latex  suspension  two 
or  three  times  until  five  or  ten  particles  drift 
into  view.  You  will  see  them  as  tiny  bright 
spots  of  hght.  You  may  have  to  adjust  the  focus 
slightly  to  see  a  specific  particle  clearly.  No- 
tice how  the  particle  appears  to  move  upward. 
The  view  is  inverted  by  the  microscope— the 
particles  are  actually  falhng  in  the  earth's 
gravitational  field. 

Now  switch  on  the  high  voltage  across  the 
plates  by  turning  the  switch  up  or  down.  No- 
tice the  effect  on  the  particles  of  varying  the 
electric  field  by  means  of  the  voltage-control 
knob. 

Notice  the  effect  when  you  reverse  the 
electric  field  by  reversing  the  switch  position. 
(When  the  switch  is  in  its  mid-position,  there 
is  zero  field  between  the  plates.) 
Q4    Do  all  the  particles  move  in  the  same 
direction  when  the  field  is  on? 
Q5     How  do  you  explain  this? 
Q6    Some  particles  move  much  more  rapidly 
in  the  field  than  others.  Do  the  rapidly  moving 
particles  have  larger  or  smaller  charges  than 
the  slowly  moving  particles? 

Sometimes  a  few  particles  chng  together, 
making  a  clump  that  is  easy  to  see — the  clump 
falls  more  rapidly  than  single  particles  when 
the  electric  field  is  off.  Do  not  try  to  use  these 
for  measuring  q. 

Try  to  balance  a  particle  by  adjusting  the 
field  until  the  particle  hangs  motionless.  Ob- 
serve it  carefully  to  make  sure  it  isn't  slowly 
drifting  up  or  down.  The  smaller  the  charge, 
the  greater  the  electric  field  must  be  to  hold 
up  the  particle. 

Taking  data 

It  is  not  worth  working  at  voltages  much  below 
50  volts.  Only  highly  charged  particles  can  be 
balanced  in  these  small  fields,  and  you  are 
interested  in  obtaining  the  smallest  charge 
possible. 

Set  the  potential  difference  between  the 
plates  to  about  75  volts.  Reverse  the  field  a 


138 


Experiment  42 


few  times  so  that  the  more  quickly  moving 
particles  (those  with  greater  charge)  are  swept 
out  of  the  field  of  view.  Any  particles  that  re- 
main have  low  charges.  If  no  particles  remain, 
squeeze  in  some  more  and  look  again  for  some 
with  small  charge. 

When  you  have  isolated  one  of  these  par- 
ticles carrying  a  low  charge,  adjust  the  voltage 
carefully  until  the  particle  hangs  motionless. 
Observe  it  for  some  time  to  make  sure  that  it 
isn't  moving  up  or  down  very  slowly,  and  that 
the  adjustment  of  voltage  is  as  precise  as  pos- 
sible. (Because  of  uneven  bombardment  by 
air  molecules,  there  will  be  some  shght,  un- 
even drift  of  the  particles.) 

Read  the  voltmeter.  Then  estimate  the  pre- 
cision of  the  voltage  setting  by  seeing  how  Ht- 
tle  the  voltage  needs  to  be  changed  to  cause  the 
particle  to  start  moving  just  perceptibly.  This 
small  change  in  voltage  is  the  greatest  amount 
by  which  your  setting  of  the  balancing  voltage 
can  be  uncertain. 

When  you  have  balanced  a  particle,  make 
sure  that  the  voltage  setting  is  as  precise  as 
you  can  make  it  before  you  go  on  to  another 
particle.  The  most  useful  range  to  work  in  is 
75-150  volts,  but  try  to  find  particles  that  can 
be  brought  to  rest  in  the  200-250  volt  range 
too,  if  the  meter  can  be  used  in  that  range.  Re- 
member that  the  higher  the  balancing  field 
the  smaller  the  charge  on  the  particle. 

In  this  kind  of  an  experiment,  it  is  helpful 
to  have  large  amounts  of  data.  This  usually 
makes  it  easier  to  spot  trends  and  to  distin- 
guish main  effects  from  the  background  scat- 
tering of  data.  Thus  you  may  wish  to  contribute 
your  findings  to  a  class  data  pool.  Before  doing 
that,  however,  arrange  your  values  of  V  in  a 
vertical  column  of  increasing  magnitude. 
Q7  Do  the  numbers  seem  to  clump  together 
in  groups,  or  do  they  spread  out  more  or  less 
evenly  from  the  lowest  to  the  highest  values? 

Now  combine  your  data  with  that  collected 
by  your  classmates.  This  can  conveniently 
be  done  by  placing  your  values  of  V  on  a  class 
histogram.  When  the  histogram  is  complete, 
the  results  can  easily  be  transferred  to  a  trans- 
parent sheet  for  use  on  an  overhead  projector. 
Alternatively,  you  may  wish  to  take  a  Polaroid 


photograph   of  the   completed  histogram  for 

inclusion  in  your  laboratory  notebook. 

Q8    Does    your    histogram    suggest    that    all 

values    of   q    are   possible    and    that    electric 

charge  is  therefore  endlessly  divisible,  or  the 

converse? 

If  you  would  like  to  make  a  more  complete 
quantitative  analysis  of  the  class  results,  cal- 
culate an  average  value  for  each  of  the  high- 
est three  or  four  clumps  of  V  values  in  the  class 
histogram.  Next  change  those  to  values  of  1/V 
and  hst  them  in  order.  Since  q  is  proportional 
to  1/V,  these  values  represent  the  magnitude 
of  the  charges  on  the  particles. 

To  obtain  actual  values  for  the  charges, 
the  1/V's  must  be  multipUed  by  mttgd.  The  sepa- 
ration d  of  the  two  plates,  typically  about  5.0 
mm,  or  5.0  x  10~^m,  is  given  in  the  specifica- 
tion sheets  provided  by  the  manufacturer. 
You  should  check  this. 

The  mass  m  of  the  spheres  is  worked  out 
from  a  knowledge  of  their  volume  and  the 
densitiy  D  of  the  material  they  are  made  from. 

Mass  =  volume  x  density,  or 

The  sphere  diameter  (careful:  2)  has  been 
previously  measured  and  is  given  on  the  supply 
bottle.  The  density  D  is  1077  kg/m'  (found  by 
measuring  a  large  batch  of  latex  before  it  is 
made  into  Httle  spheres). 

Q9  What  is  the  spacing  between  the  observed 
average  values  of  1/V  and  what  is  the  differ- 
ence in  charge  that  corresponds  to  this  differ- 
ence in  1/V? 

QIO  What  is  the  smallest  value  of  1/V  that 
you  obtained?  What  is  the  corresponding  value 
of  q? 

Qll  Do  your  experimental  results  support 
the  idea  that  electric  charge  is  quantized? 
If  so,  what  is  your  value  for  the  quantum  of 
charge? 

Q12  If  you  have  already  measured  qplm  in 
Experiment  39,  compute  the  mass  of  an  elec- 
tron. Even  if  your  value  differs  the  accepted 
value  by  a  factor  of  10.  perhaps  you  will  agree 
that  its  measurement  is  a  considerable  intel- 
lectual triumph. 


139 


EXPERIMENT  43     THE  PHOTOELECTRIC 
EFFECT 

In  this  experiment  you  will  make  observations 
on  the  effect  of  light  on  a  metal  surface;  then 
you  will  compare  the  appropriateness  of  the 
wave  model  and  the  particle  model  of  hght  for 
explaining  what  you  observe. 

Before  doing  the  experiment,  read  text 
Sec.  18.4  (Unit  5)  on  the  photoelectric  effect. 

How  the  apparatus  works 

Light  that  you  shine  through  the  window  of  the 
phototube  falls  on  a  half-cylinder  of  metal 
called  the  emitter.  The  hght  drives  electrons 
from  the  emitter  surface. 

Along  the  axis  of  the  emitter  (the  center 
of  the  tube)  is  a  wire  called  the  collector.  When 
the  collector  is  made  a  few  volta  positive  with 
respect  to  the  emitter,  practically  all  the 
emitted  electrons  are  drawn  to  it,  and  will 
return  to  the  emitter  through  an  external  wire. 
Even  if  the  collector  is  made  sUghtly  negative, 
some  electrons  will  reach  it  and  there  will  be 
a  measurable  current  in  the  external  circuit. 


de'f'cc+ov- 


However  much  the  details  may  differ,  any  equipment  for 
the  photoelectric  effect  experiment  will  consist  of  these 
basic  parts. 

The  small  current  can  be  ampHfied  several 
thousand  times  and  detected  in  any  of  several 
different  ways.  One  way  is  to  use  a  small  loud- 
speaker in  which  the  ampUfied  photoelectric 
current  causes  an  audible  hum;  another  is  to 
use  a  cathode  ray  oscilloscope.  The  following 
description  assumes  that  the  output  current 
is  read  on  a  microammeter  (Fig.  18-9). 

The  voltage  control  knob  on  the  phototube 
unit  allows  you  to  vary  the  voltage  between 
emitter  and  collector.  In  its  full  counter- 
clockwise position,  the  voltage  is  zero.  As  you 
turn  the  knob  clockwise  the  "photocurrent" 
decreases.  You  are  making  the  collector  more 


detector 


ier~ 


\/t3t  /a^>»^ 


Fig.  18-9 


140 


Experiment  43 


and  more  negative  and  fewer  and  fewer  elec- 
trons get  to  it.  Finally  the  photocurrent  ceases 
altogether — all  the  electrons  are  turned  back 
before  reaching  the  collector.  The  voltage 
between  emitter  and  collector  that  just  stops 
all  the  electrons  is  called  the  "stopping  volt- 
age." The  value  of  this  voltage  indicates  the 
maximum  kinetic  energy  with  which  the  elec- 
trons leave  the  emitter.  To  find  the  value  of 
the  stopping  voltage  precisely  you  will  have  to 
be  able  to  determine  precisely  when  the  photo- 
current  is  reduced  to  zero.  Because  there  is 
some  drift  of  the  amphfier  output,  the  current 
indicated  on  the  meter  will  drift  around  the 
zero  point  even  when  the  actual  current  re- 
mains exactly  zero.  Therefore  you  will  have  to 
adjust  the  amphfier  offset  occasionally  to  be 
sure  the  zero  level  is  really  zero.  An  alternative 
is  to  ignore  the  precise  reading  of  the  current 
meter  and  adjust  the  collector  voltage  until 
turning  the  light  off  and  on  causes  no  detect- 
able change  in  the  current.  Turn  up  the  nega- 
tive collector  voltage  until  blocking  the  hght 
from  the  tube  (with  black  paper)  has  no  effect 
on  the  meter  reading— the  exact  location  of 
the  meter  pointer  isn't  important. 

The  position  of  the  voltage  control  knob  at 
the  current  cutoff  gives  you  a  rough  measure 
of  stopping  voltage.  To  measure  it  more  pre- 
cisely, connect  a  voltmeter  as  shown  in  Fig. 
18-10. 

In  the  experiment  you  will  measure  the 
stopping  voltages  as  hght  of  different  fre- 
quencies falls  on  the  phototube.  Good  colored 
filters  will  allow  light  of  only  a  certain  range  of 
frequencies  to  pass  through.  You  can  use  a 
hand  spectroscope  to  find  the  highest  fre- 
quency line  passed  by  each  filter.  The  filters 
select  frequencies  from  the  mercury  spectrum 
emitted  by  an  intense  mercury  lamp.  Useful 
frequencies  of  the  mercury  spectrum  are: 


Yellow 

5.2  X  lO'Vsec 

Green 

5.5  X  lO'^/sec 

Blue 

6.9  X  lO'^/sec 

Violet 

7.3  X  lO'^/sec 

(Ultraviolet) 

8.2  X  lO'Vsec 

DOING  THE  EXPERIMENT 

Part  I 

The  first  part  of  the  experiment  is  qualitative. 
To  see  if  there  is  time  delay  between  hght  fall- 
ing on  the  emitter  and  the  emission  of  photo- 
electrons,  cover  the  phototube  and  then  quickly 
remove  the  cover.  Adjust  the  hght  source  and 
filters  to  give  the  smallest  photocurrent  that 
you  can  conveniently  notice  on  the  meter. 
Ql  Can  you  detect  any  time  delay  between 
the  moment  that  hght  hits  the  phototube  and 
the  moment  that  motion  of  the  microamme- 
ter  pointer  (or  a  hum  in  the  loudspeaker  or 
deflection  of  the  oscilloscope  trace)  signals 
the  passage  of  photoelectrons  through  the 
phototube? 

To  see  if  the  current  in  the  phototube  de- 
pends on  the  intensity  of  incident  hght,  vary 
the  distance  of  the  hght  source. 
Q2  Does  the  number  of  photoelectrons  emit- 
ted from  the  sensitive  surface  vary  with  hght 
intensity— that  is,  does  the  output  current  of 
the  amphfier  vary  with  the  intensity  of  the 
hght? 

To  find  out  whether  the  kinetic  energy  of 
the  photoelectrons  depends  on  the  intensity  of 
the  incident  light,  measure  the  stopping  volt- 
age with  different  intensities  of  hght  falhng 
on  the  phototube. 

Q3  Does  the  kinetic  energy  of  the  photoelec- 
trons depend  on  intensity— iha.t  is,  does  the 
stopping  voltage  change? 

Finally,  determine  how  the  kinetic  energy 
of  photoelectrons  depends  on  the  frequency  of 
incident  light.  You  will  remember  (Text  Sec. 
18.5)  that  the  maximum  kinetic  energy  of  the 
photoelectrons  is  V^,gi,q^,  where  V,,op  is  the  stop- 
ping voltage  and  q^  =  1.60  x  10"'^  coulombs, 
the  charge  on  an  electron.  Measure  the  stop- 
ping voltage  with  various  filters  over  the 
window. 

Q4  How  does  the  stopping  voltage  and  hence 
the  kinetic  energy  change  as  the  light  is 
changed  from  red  through  blue  or  ultraviolet 
(no  filters)? 

Part  II 

In  the  second  part  of  the.  experiment  you  will 


Experiment  43 


141 


make  more  precise  measurements  of  stopping 
voltage.  To  do  this,  adjust  the  voltage  control 
knob  to  the  cutoff  (stopping  voltage)  position 
and  then  measure  V  with  a  voltmeter  (Fig. 
18-10.)  Connect  the  voltmeter  only  after  the 
cutoff  adjustment  is  made  so  that  the  volt- 
meter leads  will  not  pick  up  any  ac  voltage 
induced  from  other  conducting  wires  in  the 
room. 


to 
Vo/t  meter 


Fig.    18-10 

Measure  the  stopping  voltage  V^,gp  for  three 
or  four  different  hght  frequencies,  and  plot 
the  data  on  a  graph.  Along  the  vertical  axis, 
plot  electron  energy  V^ig^q^.  When  the  stopping 
voltage  V  is  in  volts,  and  q^  is  in  coulombs, 
Vqg  will  be  energy,  in  joules. 

Along  the  horizontal  axis  plot  frequency 
of  hght/. 

Interpretation  of  Results 

As  suggested  in  the  opening  paragraph,  you 
can  compare  the  wave  model  of  light  and  the 
particle  model  in  this  experiment.  Consider, 
then,  how  these  models  explain  your  obser- 
vations. 

Q5  If  the  hght  striking  your  phototube  acts 
as  waves — 

a)  Can  you  explain  why  the  stopping  voltage 
should  depend  on  the  frequency  of  hght? 

b)  Would  you  expect  the  stopping  voltage  to 
depend  on  the  intensity  of  the  light?  Why? 

c)  Would  you  expect  a  delay  between  the  time 


that  hght  first  strikes  the  emitter  and  the  emis- 
sion of  photoelectrons?  Why? 
Q6    If  the  light  is  acting  as  a  stream  of  par- 
ticles, what  would  be  the  answer  to  questions 
a,  b  and  c  above? 

If  you  drew  the  graph  suggested  in  the  Part 
II  of  the  experiment,  you  should  now  be  pre- 
pared to  interpret  the  graph.  It  is  interesting  to 
recall  that  Einstein  predicted  its  form  in  1905, 
and  by  experiments  similar  to  yours,  Milhkan 
verified  Einstein's  prediction  in  1916. 

Einstein's  photoelectric  equation  (Text 
Sec.  18.5)  describes  the  energy  of  the  most 
energetic  photoelectrons  (the  last  ones  to  be 
stopped  as  the  voltage  is  increased),  as 


A  9 


=  hf-W. 

This  equation  has  the  form 

y  =  kx  -  c. 

In  this  equation  -c  is  a  constant,  the  value 
of  y  at  the  point  where  the  straight  hne  cuts 
the  vertical  axis;  and  k  is  another  constant, 
namely  the  slope  of  the  line.  (See  Fig.  18-11.) 
Therefore,  the  slope  of  a  graph  oiVgig^q^  against 
/  should  be  h. 

Q7  What  is  the  value  of  the  slope  of  your 
graph?  How  well  does  this  value  compare  with 


Fig.    18-11 


142 


Experiment  43 


the  value  of  Planck's  constant,  h  =  6.6  x  10  ^* 
joule-sec?  (See  Fig.  18-12). 


hf-w 


Fig.  18-12 

With  the  equipment  you  used,  the  slope  is 
unlikely  to  agree  with  the  accepted  value  of  h 
(6.6  X  10"'^^  joule-sec)  more  closely  than  an 
order  of  magnitude.  Perhaps  you  can  give  a 


few  reasons  why  your  agreement  cannot  be 
more  approximate. 

Q8  The  lowest  frequency  at  which  any  elec- 
trons are  emitted  from  the  cathode  surface  is 
called  the  threshold  frequency,  /o-  At  this 
frequency  imTy^^j.  =  0  and  h/o  =  W,  where  W 
is  the  "work  function."  Your  experimentally 
obtained  value  of  W  is  not  likely  to  be  the  same 
as  that  found  for  very  clean  cathode  surfaces, 
more  carefully  filtered  light,  etc.  The  impor- 
tant thing  to  notice  here  is  that  there  is  a  value 
of  W,  indicating  that  there  is  a  minimum  en- 
ergy needed  to  release  photoelectrons  from  the 
emitter. 

Q9  Einstein's  equation  was  derived  from  the 
assumption  of  a  particle  (photon)  model  of 
light.  If  your  results  do  not  fully  agree  with 
Einstein's  equation,  does  this  mean  that  your 
experiment  supports  the  wave  theory? 


I 


ACTIVITIES 


WRITINGS  BY  OR  ABOUT  EINSTEIN 

In  addition  to  his  scientific  works.  Einstein 
wrote  many  perceptive  essays  on  other  areas 
of  life  which  are  easy  to  read,  and  are  still  very 
current.  The  chapter  titles  from  Out  of  My 
Later  Years  (Philosophical  Library,  N.Y.  1950) 
indicate  the  scope  of  these  essays:  Convictions 
and  Beliefs;  Science;  Pubhc  Affairs;  Science 
and  Life;  Personahties;  My  People.  This  book 
includes  his  writings  from  1934  to  1950.  The 
World  As  I  See  It  includes  material  from  1922 
to  1934.  Albert  Einstein:  Philosopher-Scien- 
tist, Vol.  I.  (Harper  Torchbook,  1959)  contains 
Einstein's  autobiographical  notes,  left-hand 
pages  in  German  and  right  hand  pages  in  En- 
ghsh,  and  essays  by  twelve  physicist  contem- 
poraries of  Einstein  about  various  aspects  of 
his  work.  See  also  the  three  articles,  "Ein- 
stein," "Outside  and  Inside  the  Elevator,"  and 
"Einstein  and  Some  Civilized  Discontents"  in 
Reader  5. 


MEASURING  q/m  FOR  THE  ELECTRON 

With  the  help  of  a  "tuning  eye"  tube  such  as 
you  may  have  seen  in  radio  sets,  you  can  mea- 
sure the  charge-to-mass  ratio  of  the  electron 
in  a  way  that  is  very  close  to  J.  J.  Thomson's 
original  method. 

Complete  instructions  appear  in  the  PSSC 
Physics  Laboratory  Guide,  Second  Edition, 
D.  C.  Heath  Company,  Experiment  IV-12, 
"The  Mass  of  the  Electron,"  pp.  79-81. 


CATHODE  RAYS  IN  A  CROOKES  TUBE 

A  Crookes  tube  having  a  metal  barrier  inside 
it  for  demonstrating  that  cathode  rays  travel 
in  straight  hnes  may  be  available  in  your  class- 
room. In  use,  the  tube  is  excited  by  a  Tesla  coil 
or  induction  coil. 

Use  a  Crookes  tube  to  demonstrate  to  the 
class  the  deflection  of  cathode  rays  in  mag- 
netic fields.  To  show  how  a  magnet  focuses 
cathode  rays,  bring  one  pole  of  a  strong  bar 
magnet  toward  the  shadow  of  the  cross-shaped 
obstacle  near  the  end  of  the  tube.  Watch  what 
happens  to  the  shadow  as  the  magnet  gets 
closer  and  closer  to  it.  What  happens  when  you 


switch  the  poles  of  the  magnet?  What  do  you 
think  would  happen  if  you  had  a  stronger 
magnet? 

Can  you  demonstrate  deflection  by  an  elec- 
tric field?  Try  using  static  charges  as  in  Ex- 
periment 34,  "Electric  Forces  I,"  to  create  a 
deflecting  field.  Then  if  you  have  an  electro- 
static generator,  such  as  a  small  Van  de  GraafF 
or  a  Wimshurst  machine,  try  deflecting  the 
rays  using  parallel  plates  connected  to  the 
generator. 


X  RAYS  FROM  A  CROOKES  TUBE 

To  demonstrate  that  x  rays  penetrate  materials 
that  stop  visible  Ught,  place  a  sheet  of  4"  x  5" 
3000-ASA-speed  Polaroid  Land  film,  still  in 
its  protective  paper  jacket,  in  contact  with  the 
end  of  the  Crookes'  tube.  (A  film  pack  cannot 
be  used,  but  any  other  photographic  film  in  a 
Ught-tight  paper  envelope  could  be  substi- 
tuted.) Support  the  film  on  books  or  the  table  so 
that  it  doesn't  move  during  the  exposure.  Fig. 
18-13  was  a  1-minute  exposure  using  a  hand- 
held Tesla  coil  to  excite  the  Crookes  tube. 


18-13 


LIGHTING  AN  ELECTRIC  LAMP 
WITH  A  MATCH 

Here  is  a  trick  with  which  you  can  challenge 
your  friends.  It  illustrates  one  of  the  many 
amusing  and  useful  apphcations  of  the  photo- 


144 


Activities 


electric  effect  in  real  life.  You  will  need  the 
phototube  from  Experiment  42,  "The  Photo- 
electric Effect,"  together  with  the  Project 
Physics  Amplifier  and  Power  Supply.  You  will 
also  need  a  1 2"V  dry  cell  or  power  supply  and 
a  6V  light  source  such  as  the  one  used  in  the 
MilHkan  Apparatus.  (If  you  use  this  light 
source,  remove  the  lens  and  cardboard  tube 
and  use  only  the  6V  lamp.)  Mount  the  lamp  on 
the  Photoelectric  Effect  apparatus  and  connect 
it  to  the  0-5V,  5  amps  variable  output  on  the 
power  supply.  Adjust  the  output  to  maximum. 
Set  the  transistor  switch  input  switch  to 
switch. 

Connect  the  Photoelectric  Effect  appa- 
ratus to  the  Amplifier  as  shown  in  Fig.  18-14. 
Notice  that  the  polarity  of  the  1.5V  cell  is  re- 
versed and  that  the  output  of  the  Amphfier 
is  connected  to  the  transistor  switch  input. 

Advance  the  gain  control  of  the  amphfier 
to  maximum,  then  adjust  the  offset  control  in 
a  positive  direction  until  the  filament  of  the 
6V  lamp  ceases  to  glow.  Ignite  a  match  near 
the  apparatus  (the  wooden  type  works  the  best) 
and  bring  it  quickly  to  the  window  of  the  photo- 
tube while  the  phosphor  of  the  match  is  still 
glowing  brightly.  The  phosphor  flare  of  the 
match  head  will  be  bright  enough  to  cause  suf- 
ficient photocurrent  to  operate  the  transistor 
switch  which  turns  the  bulb  on.  Once  the  bulb 
is  lit,  it  keeps  the  photocell  activated  by  its 
own  hght;  you  can  remove  the  match  and  the 
bulb  will  stay  lit. 

When  you  are  demonstrating  this  effect, 
tell  your  audience  that  the  bulb  is  really  a 
candle  and  that  it  shouldn't  surprise  them  that 
you  can  light  it  with  a  match.  And  of  course 
one  way  to  put  out  a  candle  is  to  moisten  your 
fingers  and  pinch  out  the  wick.  When  your 
fingers  pass  between  the  bulb  and  the  photo- 


!  Amp/if  I'er 


J 


ft>wev  Supply 


inpui 


j    O        Ch  O'S  cmp 


f  ?   rr 


/.5V  l_.__ 


^.^    4V  bulb 

If  3^  phoTo-fucK-^ 
Fig.     18-14 


-- J 


cell,  the  bulb  turns  off,  although  the  filament 
may  glow  a  httle,  just  as  the  wick  of  a  freshly 
snuffed  candle  does.  You  can  also  make  a 
"candle-snuffer"  from  a  httle  cone  of  any 
reasonable  opaque  material  and  use  this  in- 
stead of  your  fingers.  Or  you  can  "blow  out" 
the  bulb:  It  will  go  out  obediently  if  you  take 
care  to  remove  it  from  in  front  of  the  photocell 
as  you  blow  it  out. 


FILM  LOOP 


FILM  LOOP  47  THOMSON  MODEL 
OF  THE  ATOM 

Before  the  development  of  the  Bohr  theory, 
a  popular  model  for  atomic  structure  was  the 
"raisin  pudding"  model  of  J.  J.  Thomson.  Ac- 
cording to  this  model,  the  atom  was  supposed 
to  be  a  uniform  sphere  of  positive  charge  in 
which  were  embedded  small  negative  "cor- 
puscles" (electrons).  Under  certain  conditions 
the  electrons  could  be  detached  and  observed 
separately,  as  in  Thomson's  historic  experi- 
ment to  measure  the  charge/mass  ratio. 

The  Thomson  model  did  not  satisfactorily 
explain  the  stabiUty  of  the  electrons  and  es- 
pecially their  arrangement  in  "rings,"  as  sug- 
gested by  the  periodic  table  of  the  elements. 
In  1904  Thomson  performed  experiments 
which  to  him  showed  the  possibility  of  a  ring 
structure  within  the  broad  outline  of  the  raisin- 
pudding  model.  Thomson  also  made  mathe- 
matical calculations  of  the  various  arrange- 
ments of  electrons  in  his  model. 

In  the  Thomson  model  of  the  atom,  the 
cloud  of  positive  charge  created  an  electric 
field  directed  along  radii,  strongest  at  the  sur- 
face of  the  sphere  of  charge  and  decreasing  to 
zero  at  the  center.  You  are  famihar  with  a 
gravitational  example  of  such  a  field.  The 
earth's  downward  gravitational  field  is  strong- 
est at  the  surface  and  it  decreases  uniformly 
toward  the  center  of  the  earth. 

For  his  model-of-a-model  Thomson  used 
still  another  type  of  field — a  magnetic  field 
caused  by  a  strong  electromagnet  above  a  tub 
of  water.  Along  the  water  surface  the  field  is 
"radial,"  as  shown  by  the  pattern  of  iron  fihngs 
sprinkled  on  the  glass  bottom  of  the  tub.  Thom- 
son used  vertical  magnetized  steel  needles  to 
represent  the  electrons;  these  were  stuck 
through  corks  and  floated  on  the  surface  of 
the  water.  The  needles  were  oriented  with  Hke 
poles  pointing  upward;  their  mutual  repulsion 
tended  to  cause  the  magnets  to  spread  apart. 
The  outward  repulsion  was  counteracted  by 
the  radial  magnetic  field  directed  inward  to- 
ward the  center.  When  the  floating  magnets 
were  placed  in  the  tub  of  water,  they  came  to 


equiUbrium  configurations  under  the  combined 
action  of  all  the  forces.  Thomson  saw  in  this 
experiment  a  partial  verification  of  his  calcula- 
tion of  how  electrons  (raisins)  might  come  to 
equilibrium  in  a  spherical  blob  of  positive 
fluid. 

In  the  film  the  floating  magnets  are  3.8  cm 
long,  supported  by  ping  pong  balls  (Fig.  18-15). 
Equihbrium  configurations  are  shown  for  var- 
ious numbers  of  balls,  from  1  to  12.  Perhaps 
you  can  interpret  the  patterns  in  terms  of 
rings,  as  did  Thomson. 


Fig.    18-15 

Thomson  was  unable  to  make  an  exact 
correlation  with  the  facts  of  chemistry.  For 
example,  he  knew  that  the  eleventh  electron 
is  easily  removed  (corresponding  to  sodium, 
the  eleventh  atom  of  the  periodic  table),  yet 
his  floating  magnet  model  failed  to  show  this. 
Instead,  the  patterns  for  10,  11  and  12  floating 
magnets  are  rather  similar. 

Thomson's  work  with  this  apparatus  illus- 
trates how  physical  theories  may  be  tested 
with  the  aid  of  analogies.  He  was  disappointed 
by  the  failure  of  the  model  to  account  for  the 
details  of  atomic  structure.  A  few  years  later 
the  Rutherford  model  of  a  nuclear  atom  made 
the  Thomson  model  obsolete,  but  in  its  day  the 
Thomson  model  received  some  support  from 
experiments  such  as  those  shown  in  the  film. 


Chapter 


19 


The  Rutherford-Bohr  Model  of  the  Atom 


EXPERIMENT  44  SPECTROSCOPY 

In  text  Chapter  19  you  learn  of  the  immense 
importance  of  spectra  to  our  understanding  of 
nature.  You  are  about  to  observe  the  spectra 
of  a  variety  of  Ught  sources  to  see  for  yourself 
how  spectra  differ  from  each  other  and  to  learn 
how  to  measure  the  wavelengths  of  spectrum 
lines.  In  particular,  you  will  measure  the  wave- 
lengths of  the  hydrogen  spectrum  and  relate 
them  to  the  structure  of  the  hydrogen  atom. 

Before  you   begin,  review  carefully  Sec. 
19.1  of  text  Chapter  19. 


Observing  spectra 

You  can  observe  diffraction  when  you  look  at 
hght  that  is  reflected  from  a  phonegraph  rec- 
ord. Hold  the  record  so  that  hght  from  a  distant 
source  is  almost  parallel  to  the  record's  sur- 
face, as  in  the  sketch  below.  Like  a  diffraction 
grating,  the  grooved  surface  disperses  light 
into  a  spectrum. 


4 


i' 


Creating  spectra 

Materials  can  be  made  to  give  off  light  (or  be 
"excited")  in  several  diff"erent  ways:  by  heat- 
ing in  a  flame,  by  an  electric  spark  between 
electrodes  made  of  the  material,  or  by  an  elec- 
tric current  through  a  gas  at  low  pressure. 

The  hght  emitted  can  be  dispersed  into  a 
spectrum  by  either  a  prism  or  a  diff"raction 
grating. 

In  this  experiment,  you  will  use  a  diffrac- 
tion grating  to  examine  hght  from  various 
sources.  A  diff"raction  grating  consists  of  many 
very  fine  parallel  grooves  on  a  piece  of  glass  or 
plastic.  The  grooves  can  be  seen  under  a  400- 
power  microscope. 

In  experiment  33  (Young's  Experiment) 
you  saw  how  two  narrow  slits  spread  hght  of 
different  wavelengths  through  diff'erent  an- 
gles, and  you  used  the  double  sht  to  make 
approximate  measurements  of  the  wave- 
lengths of  light  of  diff'erent  colors.  The  dis- 
tance between  the  two  shts  was  about  0.2  mm. 
The  distance  between  the  lines  in  a  diffrac- 
tion grating  is  about  0.002  mm.  And  a  grating 
may  have  about  10,000  grooves  instead  of 
just  two.  Because  there  are  more  hnes  and 
they  are  closer  together,  a  grating  diffracts 
more  light  and  separates  the  different  wave- 
lengths more  than  a  double-slit,  and  can  be 
used  to  make  very  accurate  measurements 
of  wavelength. 


Use  a  real  diff"raction  grating  to  see  spec- 
tra simply  by  holding  the  grating  close  to  your 
eye  with  the  hnes  of  the  grating  parallel  to  a 
distant  hght  source.  Better  yet,  arrange  a  sht 
about  25  cm  in  front  of  the  grating,  as  shown 
below,  or  use  a  pocket  spectroscope. 

'  ■  •    .--Source 


d  \'re.zT 


Look  through  the  pocket  spectroscope  at  a 
fluorescent  light,  at  an  ordinary  (incandescent) 
light  bulb,  at  mercury-vapor  and  sodium-vapor 
street  lamps,  at  neon  signs,  at  hght  from  the 
sky  (but  don't  look  directly  at  the  sun),  and  at 
a  flame  into  which  various  compounds  are  in- 
troduced (such  as  salts  of  sodium,  potassium, 
strontium,  barium,  and  calcium). 
Ql  Which  color  does  the  grating  diff'ract  into 
the  widest  angle  and  which  into  the  narrow- 
est? Are  the  long  wavelengths  diffracted  at  a 


Experiment  44 


147 


wider  angle  than  the  short  wavelengths,  or 
vice-versa? 

Q2  The  spectra  discussed  in  the  Text  are  (a) 
either  emission  or  absorption,  and  (b)  either 
hne  or  continuous.  What  different  kinds  of 
spectra  have  you  observed?  Make  a  table  show- 
ing the  type  of  spectrums  produced  by  each 
of  the  hght  sources  you  observed.  Do  you  detect 
any  relationship  between  the  nature  of  the 
source  and  the  kind  of  spectra  it  produces? 

Photographing  the  spectrum 

A  photograph  of  a  spectrum  has  several  ad- 
vantages over  visual  observation.  A  photo- 
graph reveals  a  greater  range  of  wavelengths; 
also  it  allows  greater  convenience  for  your 
measurement  of  wavelengths. 

When  you  hold  the  grating  up  to  your  eye, 
the  lens  of  your  eye  focuses  the  diffracted  rays 
to  form  a  series  of  colored  images  on  the  retina. 
If  you  put  the  grating  in  front  of  the  camera 
lens  (focused  on  the  source),  the  lens  will 
produce  sharp  images  on  the  film. 

The  spectrum  of  hydrogen  is  particularly 
interesting  to  measure  because  hydrogen  is  the 
simplest  atom  and  its  spectrum  is  fairly  easily 
related  to  a  model  of  its  structure.  In  this  ex- 
periment, hydrogen  gas  in  a  glass  tube  is 
excited  by  an  electric  current.  The  electric 
discharge  separates  most  of  the  H2  molecules 
into  single  hydrogen  atoms.) 

Set  up  a  meter  stick  just  behind  the  tube 
(Fig.  19-1).  This  is  a  scale  against  which  to 
observe  and  measure  the  position  of  the  spec- 
trum hnes.  The  tube  should  be  placed  at  about 
the  70-cm  mark  since  the  spectrum  viewed 
through  the  grating  will  appear  nearly  70  cm 
long. 

From  the  camera  position,  look  through 
the  grating  at  the  glowing  tube  to  locate  the 
positions  of  the  visible  spectral  hnes  against 
the  meter  stick.  Then,  with  the  grating  fas- 
tened over  the  camera  lens,  set  up  the  camera 
with  its  lens  in  the  same  position  your  eye  was. 
The  lens  should  be  aimed  perpendicularly  at 
the  50  cm  mark,  and  the  grating  hnes  must  be 
parallel  to  the  source. 

Now  take  a  photograph  that  shows  both 
the  scale  on  the  meter  stick  and  the  spectral 


Fig.    19-1 

hnes.  You  may  be  able  to  take  a  single  exposure 
for  both,  or  you  may  have  to  make  a  double 
exposure— first  the  spectrum,  and  then,  with 
more  hght  in  the  room,  the  scale.  It  depends 
on  the  amount  of  hght  in  the  room.  Consult 
your  teacher. 

Analyzing  the  spectrum 

Count  the  number  of  spectral  hnes  on  the 
photograph,  using  a  magnifier  to  help  pick 
out  the  faint  ones. 

Q3  Are  there  more  hnes  than  you  can  see 
when  you  hold  the  grating  up  to  your  eye?  If 
you  do  see  additional  hnes,  are  they  located 
in  the  visible  part  of  the  spectrum  (between 
red  and  violet)  or  in  the  infrared  or  ultraviolet 
part? 

The  angle  d  through  which  hght  is  diffrac- 
ted by  a  grating  depends  on  the  wavelength 
X  of  the  hght  and  the  distance  d  between  hnes 
on  the  grating.  The  formula  is  a  simple  one: 

X  =  d  sin  6. 

To  find  6,  you  need  to  find  tan  6  =  xll  as 
shown  in  Fig.  19-2.  Here  x  is  the  distance  of 
the  spectral  hne  along  the  meter  stick  from  the 
source,  and  t  is  the  distance  from  the  source 
to  the  grating.  Use  a  magnifier  to  read  x  from 
your  photograph.  Calculate  tan  9,  and  then 
look  up  the  corresponding  values  of  6  and  sin  6 
in  trigonometric  tables. 

To  find  d,  remember  that  the  grating  space 
is  probably  given  as  hnes  per  inch.  You  must 
convert  this  to  the  distance  between  hnes  in 
meters.  One  inch  is  2.54  x  10"^  meters,  so  if 
there  are  13,400  hnes  per  inch,  then  d  is 


148  Experiment  44 


(^     ^Oiyy.^A    by 


t 


X  -^ 


o^  red  (('(^lit  Soured 

Fig.  19-2  Different  images  of  the  source  are  formed  on 
of  diffracted  light.  The  angle  of  diffraction  is  equal  to  the 
ment  angle  of  the  source  in  the  photograph  so 

(2.54  X  10-2)  /  (1.34  X  10^)  -  1.89  x  10-«  meters. 

Calculate  the  values  of  A.  for  the  various 
spectral  hnes  you  have  measured. 
014    How  many  of  these  lines  are  visible  to  the 
eye? 

QS    What  would  you  say  is  the  shortest  wave 
length  to  which  your  eye  is  sensitive? 
QQ    What  is  the  shortest  wavelength  that  you 
can  measure  on  the  photograph? 

Compare  your  values  for  the  wavelengths 
with  those  given  in  the  text,  or  in  a  more  com- 
plete list  (for  instance,  in  the  Handbook  of 
Chemistry  and  Physics).  The  differences  be- 
tween your  values  and  the  pubUshed  ones 
should  be  less  than  the  experimental  uncer- 
tainty of  your  measurement.  Are  they? 

This  is  not  all  that  you  can  do  with  the  re- 
sults of  this  experiment.  You  could,  for  ex- 
ample, work  out  a  value  for  the  Rydberg 
constant  for  hydrogen  (mentioned  in  Text 
Sec.  19.2). 

More  interesting  perhaps  is  to  calculate 
some  of  the  energy  levels  for  the  excited  hydro- 
gen atom.  Using  Planck's  constant  (h  =  6.6  x 
10-3"),  the  speed  of  hght  in  vacuum  (c  =  3.0 
X  10»  m/sec),  and  your  measured  value  of  the 
wavelength  A  of  the  separate  hnes.  you  can 
calculate  the  energy  of  photons'  various  wave- 
lengths, E  =  hf=hclK  emitted  when  hydrogen 
atoms  change  from  one  state  to  another.  The 
energy  of  the  emitted  photon  is  the  difference 
in  energy  between  the  initial  and  final  states 


the  film  by  different  colors 
apparent  angular  displace- 


tan  d=- 


f. 


n  =5 
n  -4 

n  ^3 


A- 

'ground  staiC  'f°'^    _) 


*o  ^2 


o-f  hydrogen   octom 


i. 


nsf 


Fig.     19-3 
of  the  atom. 

Make  the  assumption  (which  is  correct) 
that  for  all  hnes  of  the  series  you  have  observed 
the  final  energy  state  is  the  same.  The  energies 
that  you  have  calculated  represent  the  energy 
of  various  excited  states  above  this  final  level. 

Draw  an  energy-level  diagram  something 
hke  the  one  shown  here  (Fig.  19-3.).  Show  on  it 
the  energy  of  the  photon  emitted  in  transition 
from  each  of  the  excited  states  to  the  final 
state. 

Q7    How  much  energy  does  an  excited  hydro- 
gen atom  lose  when  it  emits  red  hght? 


ACTIVITIES 


SCIENTISTS  ON  STAMPS 

As  shown  here,  scientists  are  pictured  on  the 
stamps  of  many  countries,  often  being  honored 
by  other  than  their  homeland.  You  may  want 
to  visit  a  stamp  shop  and  assemble  a  display 
for  your  classroom. 

See  also  "Science  and  the  Artist,"  in  the 
Unit  4  Handbook. 


MEASURING  IONIZATION, 
A  QUANTUM  EFFECT 

With  an  inexpensive  thyratron  885  tube,  you 
can  demonstrate  an  effect  that  is  closely  re- 
lated to  the  famous  Franck-Hertz  effect. 

Theory 

According  to  the  Rutherford-Bohr  model,  an 
atom  can  absorb  and  emit  energy  only  in  cer- 
tain amounts  that  correspond  to  permitted 
"jumps"  between  states. 

If  you  keep  adding  energy  in  larger  and 
larger  "packages,"  you  will  finally  reach  an 
amount  large  enough  to  separate  an  electron 
entirely  from  its  atom— that  is,  to  ionize  the 
atom.  The  energy  needed  to  do  this  is  called 
the  ionization  energy. 

Now  imagine  a  beam  of  electrons  being 
accelerated  by  an  electric  field  through  a  re- 
gion of  space  filled  with  argon  atoms.  This  is 
the  situation  in  a  thyratron  884  tube  with  its 
grid  and  anode  both  connected  to  a  source 
of  variable  voltage,  as  shown  schematically 
in  Fig.  19-4). 

+  X)0 


cathodt-' 


fi'lomtut 


Fig.     19-4 


In  the  form  of  its  kinetic  energy  each  elec- 
tron in  the  beam  carries  energy  in  a  single 
"package."  The  electrons  in  the  beam  colhde 
with  argon  atoms.  As  you  increase  the  acceler- 
ating voltage,  the  electrons  eventually  become 
energetic  enough  to  excite  the  atoms,  as  in  the 
Franck-Hertz  effect.  However,  your  equipment 
is  not  sensitive  enough  to  detect  the  resulting 
small  energy  absorptions.  So  nothing  seems  to 
happen.  The  electron  current  from  cathode  to 
anode  appears  to  increase  quite  linearly  with 
the  voltage,   as  you  would  expect— until  the 


150 


Activities 


electrons  get  up  to  the  ionization  energy  of 
argon.  This  happens  at  the  ionization  poten- 
tial V,,  which  is  related  to  the  ionization  en- 
ergy E,  and  to  the  charge  q^  on  the  electron 
as  follows: 

£,  =  q^V; 

As  soon  as  electrons  begin  to  ionize  argon 
atoms,  the  current  increases  sharply.  The 
argon  is  now  in  a  different  state,  called  an  ion- 
ized state,  in  which  it  conducts  electric  cur- 
rent much  more  easily  than  before.  Because 
of  this  sudden  decrease  in  electrical  resistance, 
we  may  use  the  thyratron  tube  as  an  "elec- 
tronic switch"  in  such  devices  as  stroboscopes. 
(A  similar  process  ionizes  the  air  so  that  it  can 
conduct  Ughtning.)  As  argon  ions  recapture 
electrons,  they  emit  photons  of  ultraviolet  and 
of  visible  violet  hght.  When  you  see  this  violet 
glow,  the  argon  gas  is  being  ionized. 

For  theoretical  purposes,  the  important 
point  is  that  ionization  takes  place  in  any  gas 
at  a  particular  energy  that  is  characteristic 
of  that  gas.  This  is  easily  observed  evidence  of 
one  special  case  of  Bohr's  postulated  discrete 
energy  states. 

Equipment 

Thyratron  884  tube 

Octal  socket  to  hold  the  tube  (not  essential 
but  convenient) 
Voltmeter  (0-30  volts  dc) 
Ammeter  (0-100  milhamperes) 
Potentiometer   (10,000   ohm,   2   watts  or 
larger)    or    variable    transformer,    0-120 
volts  ac 

Power  supply,  capable  of  dehvering  50-60 
mA  at  200  volts  dc 

Connect  the  apparatus  as  shown  schemat- 
ically in  Fig.  19-7. 

Procedure 

With  the  potentiometer  set  for  the  lowest  avail- 
able anode  voltage,  turn  on  the  power  and  wait 
a  few  seconds  for  the  filament  to  heat.  Now  in- 
crease the  voltage  by  small  steps.  At  each  new 
voltage,  call  out  to  your  partner  the  voltmeter 
reading.  Pause  only  long  enough  to  permit  your 
partner  to  read  the  ammeter  and  to  note  both 


readings  in  your  data  table.  Take  data  as  rapid- 
ly as  accuracy  permits:  Your  potentiometer 
will  heat  up  quickly,  especially  at  high  cur- 
rents. If  it  gets  too  hot  to  touch,  turn  the  power 
off  and  wait  for  it  to  cool  before  beginning 
again. 

Watch  for  the  onset  of  the  violet  glow. 
Note  in  your  data  table  the  voltage  at  which 
you  first  observe  the  glow,  and  then  note  what 
happens  to  the  glow  at  higher  voltages. 

Plot  current  versus  voltage,  and  mark  the 
point  on  your  graph  where  the  glow  first  ap- 
peared. From  your  graph,  determine  the  first 
ionization  potential  of  argon.  Compare  your 
experimental  value  with  pubhshed  values, 
such  as  the  one  in  the  Handbook  of  Chemistry 
and  Physics. 

What  is  the  energy  an  electron  must  have 
in  order  to  ionize  an  argon  atom? 

MODELING  ATOMS  WITH  MAGNETS 

Here  is  one  easy  way  to  demonstrate  some  of 
the  important  differences  between  the  Thom- 
son "raisin  pudding"  atom  model  and  the 
Rutherford  nuclear  model. 

To  show  how  alpha  aprticles  would  be 
expected  to  behave  in  colhsions  with  a  Thom- 
son atom,  represent  the  spread-out  "pudding" 
of  positive  charge  by  a  roughly  circular  ar- 
rangement of  small  disc  magnets,  spaced  four 
or  five  inches  apart,  under  the  center  of  a 
smooth  tray,  as  shown  in  Fig.  19-5.  Use  tape 


Fig.  19-5    The  arrangement  of  the  mag  nets  for  a   Thom- 
son atom". 


or  putty  to  fasten  the  magnets  to  the  under- 
side of  the  tray.  Put  the  large  magnet  (repre- 
senting the  alpha  particle)  down  on  top  of  the 
tray  in  such  a  way  that  the  large  magnet  is 
repelled  by  the  small  magnets  and  sprinkle 
onto  the  tray  enough  tiny  plastic  beads  to  make 
the  large  magnet  shde  freely.  Now  push  the 
"alpha  particle"  from  the  edge  of  the  tray 
toward  the  "atom."  As  long  as  the  "alpha  par- 
ticle" has  enough  momentum  to  reach  the 
other  side,  its  deflection  by  the  small  mag- 
nets under  the  tray  will  be  quite  small — never 
more  than  a  few  degrees. 

For  the  Rutherford  model,  on  the  other 
hand,  gather  all  the  small  magnets  into  a  ver- 
tical stack  under  the  center  of  the  tray,  as 
shown  in  Fig.  19-6.  Turn  the  stack  so  that  it 


Activities 
nuclecLS 


151 


Fig.  19-6  The  arrangement  of  the  magnets  for  a  "Ruth- 
erford atom." 

repels  "alpha  particles"  as  before.  This  "nu- 
cleus of  positive  charge"  now  has  a  much 
greater  effect  on  the  path  of  the  "alpha  par- 
ticle." 

Have  a  partner  tape  an  unknown  array  of 
magnets  to  the  bottom  of  the  tray — can  you 
determine  what  it  is  hke  just  by  scattering  the 
large  magnet? 

With  this  magnet  analogue  you  can  do 
some  quantitative  work  with  the  scattering 
relationships  that  Rutherford  investigated. 
(See  text  Sec.  19.3  and  Film  Loop  48,  "Ruther- 
ford Scattering"  at  the  end  of  this  Handbook 
chapter.)  Try  again  with  different  sizes  of 
magnets.  Devise  a  launcher  so  that  you  can 
control  the  velocity  of  your  projectile  magnets 
and  the  distance  of  closest  approach. 


..-J^ 


(XlplT<X 

porticle 


Fig.    19-7 


1)  Keep  the  initial  projectile  velocity  v  con- 
stant and  vary  the  distance  b  (see  Fig.  19-7); 
then  plot  the  scattering  angle  (/>  versus  b. 

2)  Hold  b  constant  and  carry  the  speed  of  the 
projectile,  then  plot  </>  versus  v. 

3)  Try  scattering  hard,  nonmagnetized  discs 
off  each  other.  Plot  4>  versus  b  and  (/>  versus 
V  as  before.  Contrast  the  two  kinds  of  scatter- 
ing-angle distributions. 


"BLACK  BOX"  ATOMS 

Place  two  or  three  different  objects,  such  as  a 
battery,  a  small  block  of  wood,  a  bar  magnet, 
or  a  ball  bearing,  in  a  small  box.  Seal  the  box, 
and  have  one  of  your  fellow  students  try  to  tell 
you  as  much  about  the  contents  as  possible, 
without  opening  the  box.  For  example,  sizes 
might  be  determined  by  tilting  the  box,  rela- 
tive masses  by  balancing  the  box  on  a  support, 
or  whether  or  not  the  contents  are  magnetic 
by  checking  with  a  compass. 

The  object  of  all  this  is  to  get  a  feeling  for 
what  you  can  or  cannot  infer  about  the  struc- 
ture of  an  atom  purely  on  the  basis  of  sec- 
ondary evidence.  It  may  help  you  to  write  a  re- 
port on  your  investigation  in  the  form  you  may 
have  used  for  writing  a  proof  in  plane  geome- 
try, with  the  property  of  the  box  in  one  column 
and  your  reason  for  asserting  that  the  property 
is  present  in  the  other  column.  The  analogy 
can  be  made  even  better  if  you  are  exception- 
ally brave:  Don't  let  the  guesser  open  the  box, 
ever,  to  find  out  what  is  really  inside. 


152 


Activities 


ANOTHER  SIMULATION 

OF  THE  RUTHERFORD  ATOM 

A  hard  rubber  "potential-energy  hill"  is  avail- 
able from  Stark  Electronics  Instruments,  Ltd., 
Box  670,  Ajax,  Ontario,  Canada.  When  you  roll 
steel  balls  onto  this  hill,  they  are  deflected  in 


somewhat  the  same  way  as  alpha  particles 
are  deflected  away  from  a  nucleus.  The  poten- 
tial-energy hill  is  very  good  for  quantitative 
work  such  as  that  suggested  for  the  magnet 
analogue  in  the  activity  "Modehng  atoms  with 
magnets." 


FILM  LOOPS 


FILM  LOOP  48:         RUTHERFORD 
SCATTERING 

This  film  simulates  the  scattering  of  alpha  par- 
ticles by  a  heavy  nucleus,  such  as  gold,  as  in 
Ernest  Rutherford's  famous  experiment.  The 
film  was  made  wdth  a  digital  computer. 

The  computer  program  was  a  sHght  modi- 
fication of  that  used  in  film  loops  13  and  14, 
on  program  orbits,  concerned  with  planetary 
orbits.  The  only  difference  is  that  the  operator 
selected  an  inverse-square  law  of  repulsion 
instead  of  a  law  of  attraction  such  as  that  of 
gravity.  The  results  of  the  computer  calcula- 
tion were  displayed  on  a  cathode-ray  tube  and 
then  photographed.  Points  are  shown  at  equal 
time  intervals.  Verify  the  law  of  areas  for  the 
motion  of  the  alpha  particles  by  projecting  the 
film  for  measurements.  Why  would  you  expect 
equal  areas  to  be  swept  out  in  equal  times? 

All  the  scattering  particles  shown  are  near 
a  nucleus.  If  the  image  from  your  projector  is 
1  foot  high,  the  nearest  adjacent  nucleus  would 
be  about  500  feet  above  the  nucleus  shown. 
Any  alpha  particles  moving  through  this  large 
area  between  nuclei  would  show  no  appre- 
ciable deflection. 

We  use  the  computer  and  a  mathematical 
model  to  tell  us  what  the  result  will  be  if  we 
shoot  particles  at  a  nucleus.  The  computer 
does  not  "know"  about  Rutherford  scattering. 
What  it  does  is  determined  by  a  program  placed 
in  the  computer's  memory,  written  in  this 
particular  instance  in  a  language  called  For- 
tran. The  programmer  has  used  Newton's  laws 
of  motion  and  has  assumed  an  inverse-square 
repulsive  force.  It  would  be  easy  to  change 
the  program  to  test  another  force  law,  for  ex- 


ample F  =  Klr^.  The  scattering  would  be  com- 
puted and  displayed;  the  angle  of  deflection 
for  the  same  distance  of  closest  approach 
would  be  different  than  for  inverse-square 
force. 

Working  backward  from  the  observed 
scattering  data,  Rutherford  deduced  that  the 
inverse-square  Coulomb  force  law  is  correct 
for  all  motions  taking  place  at  distances 
greater  than  about  10~'*m  from  the  scattering 
center,  but  he  found  deviations  from  Cou- 
lomb's law  for  closer  distances.  This  suggested 
a  new  type  of  force,  called  nuclear  force. 
Rutherford's  scattering  experiment  showed 
the  size  of  the  nucleus  (supposedly  the  same  as 
the  range  of  the  nuclear  forces)  to  be  about 
10"^^m,  which  is  about  1/10,000  the  distance 
between  the  nuclei  in  soUd  bodies. 


Chapter 


20 


Some  Ideas  from  Modern  Physical  Theories 


ACTIVITIES 


STANDING  WAVES  ON  A  BAND-SAW 
BLADE 

Standing  waves  on  a  ring  can  be  shown  by 
shaking  a  band-saw  blade  with  your  hand. 
Wrap  tape  around  the  blade  for  about  six 
inches  to  protect  your  hand.  Then  gently  shake 
the  blade  up  and  down  until  you  have  a  feehng 
for  the  lowest  vibration  rate  that  produces  re- 
inforcement of  the  vibration.  Then  double  the 
rate  of  shaking,  and  continue  to  increase  the 
rate  of  shaking,  watching  for  standing  waves. 
You  should  be  able  to  maintain  five  or  six 
nodes. 

TURNTABLE  OSCILLATOR  PATTERNS 
RESEMBLING  DE  BROGLIE  WAVES 

If  you  set  up  two  turntable  oscillators  and  a 
Variac  as  shown  in  Fig.  20-1,  you  can  draw 
pictures  resembhng  de  Broglie  waves,  Hke 
those  shown  in  Chapter  20  of  your  text. 

Place  a  paper  disc  on  the  turntable.  Set 
both  turntables  at  their  lowest  speeds.  Before 
starting  to  draw,  check  the  back-and-forth 
motion  of  the  second  turntable  to  be  sure  the 
pen  stays  on  the  paper.  Turn  both  turntables 
on  and  use  the  Variac  as  a  precise  speed  con- 
trol on  the  second  turntable.  Your  goal  is  to 
get  the  pen  to  follow  exactly  the  same  path 
each  time  the  paper  disc  goes  around.  Try 
higher  frequencies  of  back-and-forth  motion 
to  get  more  wavelengths   around  the  circle. 


For  each  stationary  pattern  that  you  get,  check 
whether  the  back-and-forth  frequency  is  an 
integral  multiple  of  the  circular  frequency. 

STANDING  WAVES  IN  A  WIRE  RING 

With  the  apparatus  described  below,  you  can 
set  up  circular  waves  that  somewhat  resemble 
the  de  Broghe  wave  models  of  certain  electron 
orbits.  You  will  need  a  strong  magnet,  a  fairly 
stiff  wire  loop,  a  low-frequency  oscillator,  and 
a  power  supply  with  a  transistor  chopping 
switch. 

The  output  current  of  the  oscillator  is 
much  too  small  to  interact  with  the  magnetic 
field  enough  to  set  up  visible  standing  waves 
in  the  wire  ring.  However,  the  oscillator  cur- 
rent can  operate  the  transistor  switch  to  con- 
trol ("chop")  a  much  larger  current  from  the 
power  supply  (see  Fig.  20-2). 


OSc'iHcdjr 


Fig.    20-1 


Fig.  20-2  The  signal  from  the  oscillator  controls  the 
transistor  switch,  causing  it  to  turn  the  current  from  the 
power  supply  on  and  off.  The  "chopped"  current  in 
the  wire  ring  interacts  with  the  magnetic  field  to  pro- 
duce a  pulsating  force  on  the  wire. 


The  wire  ring  must  be  of  non-magnetic 
metal.  Insulated  copper  magnet  wire  works 
well:  Twist  the  ends  together  and  support  the 


Activities 


155 


ring  at  the  twisted  portion  by  means  of  a  bind- 
ing post,  Fahnestock  clip,  thumbtack,  or  ring- 
stand  clamp.  Remove  a  httle  insulation  from 
each  end  for  electrical  connections. 

A  ring  4  to  6  inches  in  diameter  made  of 
22-guage  enameled  copper  wire  has  its  lowest 
rate  of  vibration  at  about  20  cycles/sec.  Stiffer 
wire  or  a  smaller  ring  will  have  higher  charac- 
teristic vibrations  that  are  more  difficult  to  see. 

Position  the  ring  as  shown,  with  a  section 
of  the  wire  passing  between  the  poles  of  the 
magnet.  When  the  pulsed  current  passes 
through  the  ring,  the  current  interacts  with 
the  magnetic  field,  producing  alternating 
forces  which  cause  the  wire  to  vibrate.  In 
Fig.  20-2,  the  magnetic  field  is  vertical,  and  the 
vibrations  are  in  the  plane  of  the  ring.  You 
can  turn  the  magnet  so  that  the  vibrations  are 
perpendicular  to  the  ring. 

Because  the  ring  is  clamped  at  one  point, 
it  can  support  standing  waves  that  have  any 
integral  number  of  half  wavelengths.  In  this 
respect  they  are  different  from  waves  on  a  free 
wire  ring,  which  are  restricted  to  integral 
numbers  of  whole  wavelengths.  Such  waves 
are  more  appropriate  for  comparison  to  an 
atom. 

When  you  are  looking  for  a  certain  mode  of 
vibration,  position  the  magnet  between  ex- 
pected nodes  (at  antinodes).  The  first  "charac- 
teristic, or  state"  "mode  of  vibration,"  that  the 
ring  can  support  in  its  plane  is  the  first  har- 
monic, having  two  nodes:  the  one  at  the  point 


of  support  and  the  other  opposite  it.  In  the  sec- 
ond mode,  three  nodes  are  spaced  evenly 
around  the  loop,  and  the  best  position  for  the 
magnet  is  directly  opposite  the  support,  as 
shown  in  Fig.  20-3. 


Fig.  20-3 

You  can  demonstrate  the  various  modes 
of  vibration  to  the  class  by  setting  up  the  mag- 
net, ring,  and  support  on  the  platform  of  an 
overhead  projector.  Be  careful  not  to  break 
the  glass  with  the  magnet,  especially  if  the 
frame  of  the  projector  happens  to  be  made  of 
a  magnetic  material. 

The  Project  Physics  Film  Loop  "Vibrations 
of  a  Wire,"  also  shows  this. 


INDEX 


I 


INDEX/TEXT  SECTION 


Actinide  series,  24 
Alchemy,  6-7 
Alkaline  earth  family,  19 
Alpha  particle,  66-67,  68 
Anode,  34 
Argon,  85 
Aristotle,  4-5,  7 
Atom,  3,  11-14,  29 

compound,  13 

hydrogen,  72,  74 

levels,  83-85 

mass,  14-15,  17,  28,  33 

mercury,  79 

model,   12,    13,  66,   71,   75,   78, 
107 

number,  24-25,  55 

stationary  states  of,  72 

structure,     33-35,     54-55,     83 

theory  of,  4,  8 
Atomic  bomb,  45 
Atomic  mass  unit,  40 
Atomic  number,  24-25 
Atomic  physics,  113 
Atomic  theory,  86,  88-89 
Atomic-volume,   of   elements,   21 
Atomism,  3,  5,  16 

Balmer,  Johann  Jakob,  63,  77,  78, 

83 
Barium  platinocyanide,  48 
Battery,  25-26 

Bohr,  Niels,  34,  70,  71-75,  76,  106, 
117 

inadequacy  of  theory,  86,  88-89 

model,  55,  58,  83 

periodic  table,  86 

quantization  rule,  73 

theory,  75,  77-79,  82 
Born,  Max,  104,  106,  113 
Boyle,  Robert,  7,  116 
Brownian  motion,  45 
Bunsen,  Robert  W.,  61 

California  Institute  of  Technology, 

40,47 
Cambridge  Electron  Accelerator, 

98 
Cambridge  University,  35,  104 
Cathode,  34 

rays,  34,  36-37,  40 
Cavendish,  Henry,  7 
Cavendish  Laboratory,  35 
Charge,  nuclear,  69-71 

total,  28 
Chemical  formula,  16 
Chemistry,  7 
Colhsion,  elastic,  79 
Columbia  University,  40,  47 


Compounds,  8,  29 
Compton,  Arthur  H.,  100 
Conductors,  25 
Coulomb,  28,  35,  58 
Crookes,  Sir  William,  34 
tube,  34 

Dalton,  John,  13 

atomic  theory,  8,  11-14,  25 

compounds,  29 

element  symbols,  10 

model,  12 

A  New  System  of  Chemical  Phi- 
losophy, 11 
Davisson,  C.  J.,  102 
Davy,  Humphrey,  26 
De  Broglie,  Louis,   101,   102,   103, 
105 

waves,  101,  102,  103,  109 
Delphi,  shrine  of,  2 
Democritus,  3,  4,  116 
Deterministic,  114 
Diffraction,  106 

grating,  50 

pattern,  94,  102,  111 

X-ray,  51 
Dirac,  P.  A.  M.,  105,  106 
Dobereiner,  Johann  Wolfgang,  18 
Dublin  Institute  for  Advanced 

Studies,  105 
Dynamite,  80 

Einstein,  Albert,  43,  45,  95,  96,  98, 
114 

photoelectric  effect,  43-44,  46- 
47 
Electricity 

and  matter,  25-26,  28-29 
Electrodes,  26 
Electrolysis,  25,  26,  28 
Electromagnetic    theory,    of   light, 

42 
Electromagnetic  wave,  35 
Electron,  37,  100 

charge  of,  37-38 

kinetic  energy  of,  41 

momentum  of,  101 

orbits  of,  82-86 

shells,  84 

subshells,  85 

velocity,  109 

volts  (eV),  79,  82 
Electroscope,  50 
Elements,  4 

atomic  mass  of,  14-15 

atomic-volume,  21 

combining  capacity,  17 

family  of,  18-19 


159 


four  basic,  5 

known  by  1872  (table),  16 

melting    and    boiling    tempera- 
tures of  (table),  31 

noble,  24 

order  among,  18-19 

properties  of,  16-18 

rare  earth,  24 

transition,  24 

triads,  18 
Elements  of  Chemistry  (Lavoisier), 

7-8 
Empedocles,  4 
Energy,  kinetic,  41,  79,  98 

levels,  85 

potential,  42,  98 
Epicurus,  5 
Escher,  M.  C,  115 
Esterman,  I.,  102 

Faraday,  Michael,  26,  28,  29 
Fluoresce,  48 
Fluorescent  lights,  34 
Formula,  chemical,  16 
Franck,  James,  79,  82 
Franck-Hertz  experiment,  79,  82 
Fraunhofer,  Joseph  von,  61,  62 
Frequency,  41,  72 
threshold,  41 

Galileo,  116 
Gases,  25,  50 

noble,  19,  24 

spectra  of,  59-63 
Gassendi,  Pierre,  7 
Geiger,  Hans,  66,  67,  68,  69 
Geiger  counter,  69 
Geissler,  Heinrich,  34 

tubes,  34 
Gell-Mann,  Murray,  38 
Generator,  high  voltage,  52 

Van  de  Graaf,  52 
Germer,  L.  H.,  102 
Goldstein,  Eugene,  34 
Gravitational  constant  (G),  7 
Greeks 

and  order,  2 
Guericke,  34 

Halogens,  19 

Heisenberg,  Werner,  105,  106 

Herschel,  John,  61 

Hertz,  Heinrich,  40 

Hertz,  Gustav,  79,  82 

Hittorf,  Johann,  34 

Hydrogen 

atom,  72,  83 

spectral  series  of,  75,  77-79 

spectrum,  63-65 


Ionized  gas,  50 

Ions,  26 

Institute   of   Advanced    Studies, 

Princeton,  45 
Integers,  89 

Joule,  78 

K-shell,  83 

Kinetic  energy,  41,  79,  98 
King  William  IV,  13 
Kirchhoff,  Gustave  R.,  61,  62 

Lavoisier,  Antoine,  7 

Elements  of  Chemistry,  IS 
Law  of  conservation 

of  definite  proportions,  12 

of  mass,  12 

of  multiple  proportions,  13 
Leucippus,  3,  4,  116 
Light  wave,  scattered,  100 
Lithium  atom,  83 
Lord  Rayleigh,  24 
L-shell,  83 
Lucretius,  3,  5 

On  the  Nature  of  Things,  3 

Magnus,  Albert,  17 
Manchester  University,  66 
Marsden,  67,  68,  69 
Mass 

atomic,  14-15,  28,  33 

equivalent,  98 

law  of  conservation  of,  12 

relativistic,  96 
Matter,  and  electricity,  25-26,  28- 
29 

model  of,  4 

nature  of,  1 

theory  of,  5-6 
Maxwell,  James  C.,  106 
McGill  University,  Montreal,  66 
Melville,  Thomas,  59 
Mendeleev,  Dmitri,  19,  20,  70,  116 

periodic  table,  19-23 
Mercury  atom,  79 
Metals,  alkali,  18 
Metaphysics  (Aristotle),  5 
Meteorology,  12 
Meyer,  Lothar,  21 
Microscope 

electron,  109 

field-ion,  109 

light,  109 
Millikan,  Robert  A.,  38,  40,  47 

oil  drop  experiment,  38,  39 
Model  of  atom 

Bohr,  55,  58,  71-75 


160 


mathematical,  78,  107 

physical,  107 

Rutherford,  66-69 

Thomson,  55,  56 
Momentum,  99 
Monolith,  1 
M-shell,  84 

Neutron,  50 
Newlands,  J.A.R.,  18 
A  New  System  of  Chemical  Phi- 
losophy (Dalton),  11 
Newton,  Isaac,  7,  116 
Nobel,  Alfred  B.,  80 
Nobel  Prize,  80,  100 

physics  in,  40,  43,  49,  81 
Noble  elements,  24 

gases,  24 
N-shell,  85 
Nuclear 

atom,  68 

charge,  69-71 

size,   69—71 

Oberlin  College,  47 
Orbits,  of  electrons,  82-86 
Owens    College,    Manchester,    En- 
gland, 35 

Particles,  charged,  35 

wave-Uke,  101-103,  106 
Paschen,  F.,  64 
Periodicity,  33 
Periodic  properties,  21 
Periodic   table,   19-23,   23-25,  33, 

70,  82-86 
Photoelectric  current,  41 

effect,    40,    41,    43-44,    46-47 
Photon,  43,  101,  111 

momentum  of,  100 
Planck,  Max,  47 

constant,  43,  46,  47,  72 
Pliicker,  Julius,  34 
Potential  energy,  42,  98 
Probability  interpretation,  111-114 
Pro  ton -neutron  theory,  105 
Pupin,  Michael,  48 

q/m  value,  35,  37,  38,  54 
Quanta,  41,  46,  47,  55,  111 
Quantum,  43 

light,  99 

mechanics,   95,   106,    107,    113, 
114,  117 

numbers,  89 


physics,  47 

theory,  41,  88,  100,  112 
Quarks,  38 

Rontgen,  Wilhelm  K.,  48,  50,  51 

On   a  New  Kind  of  Rays,   48 

rays  (Xrays),  50 
Radar,  108 
Radiation,  duahsm  of,  101 

particle-like,  99 
Ramsay,  William,  24 
Rare-earth  element,  24 
Relativistic  mass,  96 
Relativity  Theory,  95-99 
Rutherford,  Ernest,  35,  66,  67,  117 

Bohr  model,  71,  82 
Rydberg,  J.  R.,  64 

constant,  77 

Scattering  experiment,  66 
Schrodinger,  Erwin,  105,  106,  107, 

111 
Scientific  Revolution,  7 
Shells,  84 

Smith,  Frederick,  48 
Spectra,  59-63 
Spectroscope,  61 
Spectrum  analysis,  61 
Stationary  states,  72 
Sub-shells,  85 

Thomson,  J.  J.,  32,  35,  37,  40,  50, 
96,  117 

atom  model,  55 

q/m  experiment,  36 
Transition  elements,  24 
Triads 

of  elements,  18 

Ultraviolet  light,  51 
Uncertainty  principle,  110-111 
University  of  Chicago,  40 

Van  de  Graaf  generator,  52 
Velocity,  109 

electron,  109 
Volta,  Allessandro,  25 
Voltage,  stopping,  42 

Wollaston,  William,  61 
Wavelengths,  50,  51 

X  ray,  48,  50,  51,  53,  54,  99,  100, 
102 
diffraction,  51 


161 


INDEX/HANDBOOK  SECTION 


Accelerator,  electron.  Scientific  American 
(January  1959),  131 

Activities : 

activities  from  Scientific  American,  131 
"black  box"  atoms,  151-152 
cathode  rays  in  a  Crooke's  tube,  143 
Dalton's  Puzzle,  129 
electrolysis  of  water,  129 
lighting  an  electric  lamp  with  a  match,  144 
measurement  of  ionization,  149-150 
measuring  q/m  for  the  electron,  143 
modeling  atoms  with  magnets,  150-151 
periodic  table(s),  129-131 
scientists  on  stamps,  305 
single-electrode  plating,  131 
standing  waves  on  a  band-saw  blade,  154 
standing  waves  in  a  wire  ring,  154-155 
Thomson  model  of  the  atom,  145 
turntable  oscillator  patterns  resembling 

de  Broglie  laws,  154 
X-rays  from  a  Crooke's  tube,  143 

Alpha  particles,  scattering  of,  153 

Argon,  ionization  energy  of,  149-150 

Atom(s),  "black  box"  (activity),  151-152 
copper,  calculating  mass  and  volume  of,  128 
modeling  with  magnets  (activity),  150-151 
Rutherford-Bohr  model  of,  146-148 
Thomson  model  of  (activity),  145 
see  also  nucleus 

Atomic  masses,  relative  (activity),  129 

Balanced  particle,  electric  force  on,  136 
Band-saw  blade,  standing  waves  on   (activity), 

154 
Beta  ray  spectrometer,  Scientific  American 

(September  1958),  131 
"Black  box"  atoms  (activity),  151-152 
de  Broglie  waves,  154 

Calibration,  of  coils,  134 

Carbon  14  dating.  Scientific  American  (February 

1957),  131 
Cathode  ray(s),  and  charge-to-mass  ratio,  133- 
135 
in  a  Crooke's  tube  (activity),  143 
Charge-to-mass  ratio,  of  electron,  143 
equation  for,  133 
(experiment),  138-135 
Chemical  change,  and  electric  currents,  126—128 
Cloud    chamber,    diffusion.    Scientific    American 
(September  1952),  131 
plumber's  friend.  Scientific  American  (Decem- 
ber 1956),  131 
Wilson,  Scientific  American  (April  1956),  131 
with  magnet.  Scientific  American  (June  1959), 
131 
Copper  atom,  calculating  mass  of,  127-128 
Coulomb's  force  law,  153 


Crooke's  tube,  cathode  rays  in  (activity),  143 

x-rays  from  (activity),  143 
Current  balance,  in  calibrating  coils,  134 
Cyclotron,  Scientific  American  (September 
1953),  131 

Dalton's  Puzzle  (activity),  129 
Davy,  Humphry,  and  electrochemical  reactions, 
126 
and  sodium  production  by  electrolysis,  132 
Diffraction  angle,  of  light,  formula  for,  147-148 
Diffraction  grating,  146-147 

Einstein,  Albert 

Albert  Einstein:  Philosopher-Scientist,  143 

Out  of  My  Later  Years,  143 

photoelectric  equation  of,  141 

The  World  As  I  See  It,  143 
Electric  charge,  computation  of,  127 

measurement  of  (experiment),  136-138 
Electric  currents,  and  chemical  change,  126-128 
Electric  force,  on  balanced  particle,  136 
Electric  lamp,  lighting  with  a  match  (activity), 

144 
Electrolysis,  (experiment),  126-128 

sodium  production  by  (film  loop),  132 

of  water  (activity),  129 
Electron,  charge  of,  141;  (experiment)  136-138 

charge-to-mass  ratio  for  (experiment), 
133-135 

measuring  q/m  for  (activity),  143 
Electron  micrograph,  of  latex  spheres,  136 
Elementary    charge,    measurement    of    (experi- 
ment), 136-138 
Experiments : 

charge-to-mass  ratio  for  an  electron,  133-135 

electrolysis,  126-128 

measurements  of  elementary  charge,  136—138 

photoelectric  effect,  the,  139-142 

spectroscopy,  146-148 

Faraday,  and  electrochemical  reactions,  126 
Film  loops: 

Production  of  sodium  by  electrolysis,  132 

"Rutherford  scattering,"  151,  153 
Fortran,  153 
Franck-Hertz  effect,  149 

Gas    discharge    tubes,    how   to   make,   Scientific 

American  (February  1958),  131 
Geiger  counter,  how  to  make.  Scientific  American 

(May  1969),  131 

Handbook  of  Chemistry  and  Physics,  148,  150 
High  voltage  reversing  switch,  137 
Hydrogen,  Rydberg  constant  for,  148 
Hydrogen  spectrum,  measuring  wavelengths  of 
(experiment),  146-148 


163 


Ionization,  measurement  of  (activity),  149-150 
Ionization  energy,  149 
Ionization  potential,  150 

Isotopic  experiments,  Scientific  American  (May 
1960),  131 

Latex  spheres,  electron  micrograph  of,  136 
Light,  calculation  of  diffraction  angle,  147-148 
dispersion  into  a  spectrum,  146-147 
effect  on  metal  surface  (experiment), 

139-142 
wave  vs.  particle  models  of,  139,  141-142 
Linear   time    chart   of   element  discovery   dates, 
129,  131 

Magnetic  resonance  spectrometer.  Scientific 

American  (April  1959),  131 
Magnets,  modeling  atoms  with  (activity), 

150-151 
"Mass  of  the  Electron,  The,"  Physics  Laboratory 

Guide,  143 
Measurement  of  elementary  charge  (experiment), 

136-138 
Mercury  spectrum,  frequencies  of,  140 
Milliken,  oil  drop  experiment,  136 
Modeling  atoms  with  magnets  (activity), 

150-151 

Newton,  laws  of  motion,  153 
Nuclear  force,  153 
Nucleus,  size  of,  153 
see  also  Atom 

Out  of  My  Later  Years  (Albert  Einstein),  143 

Particle  model,  of  hght,  139,  141-142 

Periodic  Table(s),  exhibit  of  (activity),  129-131 

Photoelectric  effect,  144 

(experiment),  139-142 
Photoelectric  equation,  Einstein's,  141 
Physics    Laboratory    Guide,    "The    Mass    of    the 

Electron,"  143 
Planck's  constant,  142,  148 
Potential-energy  hill,  152 
Pythagoras'  theorem,  135 

"Raisin  pudding"  model  of  atom,  145,  150 
Rutherford  nuclear  atom  model,  150-151 
Rutherford-Bohr  model  of  atom,  146-148,  149 


Rutherford  scattering  (film  loop),  151, 
Rydberg  constant,  for  hydrogen,  148 


153 


Scientific  American,  activities  from,  131 

Scintillation  counter.  Scientific  American  (March 
1953), 131 

Single-electrode  plating  (activity),  131 

Sodium,  production  by  electrolysis   (film  loop), 
132 

Spectra,  creation  of,  146 
observation  of,  146-147 

Spectrograph,  astronomical.  Scientific  American 
(September  1956),  131 

Spectrograph,  Bunsen's  Scientific  American  (June 
1955),  131 

Spectroheliograph,  how  to  make,  Scientific  Amer- 
ican (April  1958),  131 

Spectroscopy  (experiment),  146—148 

Spectrum,  analysis  of,  147-148 
photographing  of,  147 

Spectrum  lines,  measuring  wavelengths  of  (ex- 
periment), 146-148 

Spinthariscope,  Scientific  American  (March 
1953),  131 

Stamps,  scientists  depicted  on  (activity),  147 

Standing  waves,  on  a  band  saw  (activity),  154 
in  a  wire  ring  (activity),  154—155 

Subatomic  particle  scattering,  simulating.  Scien- 
tific American  (August  1955),  131 

Thomson,  J.  J.,  and  cathode  rays,  133 

"raisin  pudding"  model  of  atom,  145,  150 
Thratron  884  tube,  in  ionization 
measurement  activity,  149 
Threshold  frequency,  142 
Turntable  oscillator  patterns  (activity),  154 

"Ups  and  Downs  of  the  Periodic  Table,"  129 

Vibration,  modes  of,  155 

Volta,  and  electrochemical  reactions,  126 

Wave(s),  de  Broglie,  154 

model,  of  light,  139,  141-142 

standing,  154—155 
Water,  electrolysis  of  (activity),  129 
World  As  I  See  It,  The  (Albert  Einstein),  143 

X-rays  from  a  Crooke's  tube  (activity),  143 


164 


Answers  to  End-of-Section  Questions 


Chapter  17 

Q1     The  atoms  of  any  one  element  are  identical  and 

unchanging. 

Q2     Conservation  of  matter;  the  constant  ratio  of 

combining  weights  of  elements.  These  successes  lend 

strength  to  the  atomic  theory  of  matter  and  to  the 

hypothesis  that  chemical  elements  differ  from  one 

another  because  they  are  composed  of  different 

kinds  of  atoms. 

Q3     No. 

Q4     It  was  the  lightest  known  element — and  others 

were  rough  multiples. 

Q5     Relative  mass;  and  combining  number,  or 

"valence." 

Q6     2,4,5,1,2. 

Q7     Density,  melting  point,  chemical  activity, 

"valence." 

Q8     Because  when  the  elements  are  arranged  as  they 

were  in  his  table,  there  is  a  periodic  recurrence  of 

elements  with  similar  properties;  that  is,  elements 

with  similar  properties  tend  to  fall  in  the  same  column 

of  the  table. 

Q9     increasing  atomic  mass. 

Q10    When  he  found  that  the  chemical  properties  of  the 

next  heaviest  element  clearly  indicated  that  it  did  not 

belong  in  the  next  column  but  in  one  further  to  the  right. 

Q11     He  was  able  to  predict  in  considerable  detail  the 

properties  of  missing  elements,  and  these  predictions 

proved  to  be  extremely  accurate,  once  the  missing 

elements  were  discovered  and  studied. 

Q12     Its  position  in  the  periodic  table,  determined  by 

many  properties  but  usually  increasing  regularly  with 

atomic  mass.  Some  examples  are:  hydrogen,  1;  oxygen, 

8;  uranium,  92. 

Q13     Water,  which  had  always  been  considered  a  basic 

element,  and  had  resisted  all  efforts  at  decomposition, 

was  easily  decomposed. 

Q14     New  metals  were  separated  from  substances 

which  had  never  been  decomposed  before. 

Q15     The  amount  of  charge  transferred  by  the  current, 

the  valence  of  the  elements,  and  the  atomic  mass  of 

the  element. 

Q16     First,  when  two  elements  combine,  the  ratio  of 

their  combining  masses  is  equal  to  the  ratio  of  their 

values  for  A/v.  Secondly,  A/v  is  a  measure  of  the  amount 

of  the  material  which  will  be  deposited  in  electrolysis. 

Chapter  18 

Q1     They  could  be  deflected  by  magnetic  and  electric 

fields. 

Q2    The  mass  of  an  electron  is  about  1800  times  smaller 

than  the  mass  of  a  hydrogen  ion. 

Q3     (1)  identical  electrons  were  emitted  by  a  variety 

of  materials;  and  (2)  the  mass  of  an  electron  was  much 

smaller  than  that  of  an  atom. 

Q4    All  other  values  of  charge  he  found  were  multiples 

of  that  lowest  value. 

Q5     Fewer  electrons  are  emitted,  but  with  the  same 

average  energy  as  before. 


Q6    The  average  kinetic  energy  of  the  emitted  electrons 

decreases  until,  below  some  frequency  value,  none 

are  emitted  at  all.  •  ■    i  . 

Qj  ^  Light  source 


Evacuated  tube 


Q8     The  energy  of  the  quantum  is  proportional  to  the 

frequency  of  the  wave,  E  —  hf. 

Q9     The  electron  loses  some  kinetic  energy  in  escaping 

from  the  surface. 

QIC     The  maximum  kinetic  energy  of  emitted  electrons 

is  2.0  eV. 

Q11     When  x  rays  passed  through  material,  say  air, 

they  caused  electrons  to  be  ejected  from  molecules, 

and  so  produced  +  ions. 

Q12     (1)  Not  deflected  by  magnetic  field;  (2)  show 

diffraction  patterns  when  passing  through  crystals; 

(3)  produced  a  pronounced  photoelectric  effect. 

Q13     (1)  Diffraction  pattern  formed  by  "slits"  with 

atomic  spacing  (that  is,  crystals);  (2)  energy  of  quantum 

in  photoelectric  effect;  (3)  their  great  penetrating  power. 

Q14     For  atoms  to  be  electrically  neutral,  they  must 

contain  enough  positive  charge  to  balance  the  negative 

charge  of  the  electrons  they  contain;  but  electrons  are 

thousands  of  times  lighter  than  atoms. 

Q15     There  are  at  least  two  reasons:  First,  the  facts 

never  are  all  in,  so  models  cannot  wait  that  long. 

Secondly,  it  is  one  of  the  main  functions  of  a  model  to 

suggest  what  some  of  the  facts  (as  yet  undiscovered) 

might  be. 

Chapter  19 

Q1     The  source  emits  light  of  only  certain  frequencies, 

and  is  therefore  probably  an  excited  gas. 

Q2     The  source  is  probably  made  up  of  two  parts:  an 

inside  part  that  produces  a  continuous  spectrum;  and 

an  outer  layer  that  absorbs  only  certain  frequencies. 

Q3     Light  from  very  distant  stars  produces  spectra 

which  are  identical  with  those  produced  by  elements 

and  compounds  here  on  earth. 

Q4     None  (he  predicted  that  they  would  exist  because 

the  mathematics  was  so  neat). 

Q5     Careful  measurement  and  tabulation  of  data  on 

spectral  lines,  together  with  a  liking  for  mathematical 

games. 


165 


Q6    At  this  point  In  the  development  of  the  book,  one 
cannot  say  what  specifically  accounts  for  the  correct- 
ness of  Balmer's  formula  (the  explanation  requires 
atomic  theory  which  is  yet  to  come).  But  the  success  of 
the  formula  does  indicate  that  there  must  be  something 
about  the  structure  of  the  atom  which  makes  it  emit 
only  discrete  frequencies  of  light. 
Q7    They  have  a  positive  electric  charge  and  are 
repelled  by  the  positive  electric  charge  in  atoms.  The 
angle  of  scattering  is  usually  small  because  the  nuclei 
are  so  tiny  that  the  alpha  particle  rarely  gets  near 
enough  to  be  deflected  much.  However,  once  in  a  while 
there  is  a  close  approach,  and  then  the  forces  of 
repulsion  are  great  enough  to  deflect  the  alpha  particle 
through  a  large  angle. 

Q8     Rutherford's  model  located  the  positively  charged 
bulk  of  the  atom  in  a  tiny  nucleus — in  Thomson's  model 
the  positive  bulk  filled  the  entire  atom. 
Q9     It  is  the  number,  Z,  of  positive  units  of  charge  found 
in  the  nucleus,  or  the  number  of  electrons  around  the 
nucleus. 

Q10    3  positive  units  of  charge  (when  ail  3  electrons 
were  removed). 

Q11     Atoms  of  a  gas  emit  light  of  only  certain  fre- 
quencies, which  implies  that  each  atom's  energy  can 
change  only  by  certain  amounts. 
Q12     None.  (He  assumed  that  electron  orbits  could 
have  only  certain  values  of  angular  momentum,  which 
Implied  only  certain  energy  states.) 
Q13    All  hydrogen  atoms  have  the  same  size  because 
in  all  unexcited  atoms  the  electron  is  in  the  innermost 
allowable  orbit. 

Q14    The  quantization  of  the  orbits  prevents  them 
from  having  other  arbitrary  sizes. 
Q15     Bohr  derived  his  prediction  from  a  physical  model, 
from  which  other  predictions  could  be  made.  Balmer 
only  followed  out  a  mathematical  analogy. 
Q16     According  to  Bohr's  model,  an  absorption  line 
would  result  from  a  transition  within  the  atom  from  a 
lower  to  a  higher  energy  state  (the  energy  being  ab- 
sorbed from  the  radiation  passing  through  the  material). 
Q17     (a)  4.0  eV  (b)  0.1  eV  (c)  2.1  eV. 
Q18    The  electron  arrangements  in  noble  gases  are 
very  stable.  When  an  additional  nuclear  charge  and  an 
additional  electron  are  added,  the  added  electron  Is 
bound  very  weakly  to  the  atom. 

Q19     Period  I  contains  the  elements  with  electrons  in 
the  K  shell  only.  Since  only  two  electrons  can  exist  in  the 


K  shell.  Period  I  will  contain  only  the  two  elements  with 
one  electron  and  two  electrons  respectively.  Period  II 
elements  have  electrons  in  the  K  (full)  and  L  shells.  The  L 
shell  can  accommodate  8  electrons,  so  those  elements 
with  only  one  through  eight  electrons  in  the  L  shell 
will  be  in  Period  II.  And  so  forth. 
Q20     It  predicted  some  results  that  disagreed  with 
experiment;  and  it  predicted  others  which  could  not  be 
tested  in  any  known  way.  It  did,  however,  give  a  satis- 
factory explanation  of  the  observed  frequency  of  the 
hydrogen  spectral  lines,  and  it  provided  a  first  physical 
picture  of  the  quantum  states  of  atoms. 


Chapter  20 

Q1     It  increases,  without  limit. 

Q2     It  increases,  approaching  ever  nearer  to  a  limiting 

value,  the  speed  of  light. 

Q3     Photon  momentum  is  directly  proportional  to  the 

frequency  of  the  associated  wave. 

Q4    The  Compton  effect  is  the  scattering  of  light  (or 

x-ray)  photons  from  electrons  in  such  a  way  that  the 

photons  transfer  a  part  of  their  energy  and  momentum 

to  the  electrons,  and  thus  emerge  as  lower  frequency 

radiation.  It  demonstrated  that  photons  resemble 

material  particles  in  possessing  momentum  as  well  as 

energy;  both  energy  and  momentum  are  conserved  in 

collisions  involving  photons  and  electrons. 

Q5     By  analogy  with  the  same  relation  for  photons. 

Q6     The  regular  spacing  of  atoms  in  crystals  is  about 

the  same  as  the  wavelength  of  low-energy  electrons. 

Q7     Bohr  invented  his  postulate  just  for  the  purpose. 

Schrodinger's  equation  was  derived  from  the  wave 

nature  of  electrons  and  explained  many  phenomena 

other  than  hydrogen  spectra. 

Q8     It  is  almost  entirely  mathematical — no  physical 

picture  or  models  can  be  made  of  it. 

Q9     It  can.  But  less  energetic  photons  have  longer 

associated  wavelengths,  so  that  the  location  of  the 

particle  becomes  less  precise. 

Q10     It  can.  But  the  more  energetic  photons  will 

disturb  the  particle  more  and  make  measurement  of 

velocity  less  precise. 

Q11     They  are  regions  where  there  is  a  high  probability 

of  quanta  arriving. 

Q12     As  with  all  probability  laws,  the  average  behavior 

of  a  large  collection  of  particles  can  be  predicted 

with  great  precision. 


166 


staff  and  Consultants  (continued) 


Sidney  Rosen,  University  of  Illinois,  Urbana 
John  J.  Rosenbaum,  Livermore  High  School, 

Calif. 
William  Rosenfeld,  Smith  College,  Northampton, 

Mass. 
Arthur  Rothman,  State  University  of  New  York, 

Buffalo 
Daniel  Rufolo,  Clairemont  High  School,  San 

Diego,  Calif. 
Bernhard  A.  Sachs.  Brooklyn  Technical  High 

School,  N.Y. 
Morton  L.  Schagrin,  Denison  University,  Granville, 

Ohio 
Rudolph  Schiller,  Valley  High  School,  Las  Vegas, 

Nev. 
Myron  O.  Schneiderwent,  Interlochen  Arts 

Academy,  Mich. 
Guenter  Schwarz,  Florida  State  University, 

Tallahassee 
Sherman  D.  Sheppard,  Oak  Ridge  High  School, 

Tenn. 
William  E.  Shortall,  Lansdowne  High  School, 

Baltimore,  Md. 
Devon  Showley,  Cypress  Junior  College,  Calif. 
William  Shurcliff.  Cambridge  Electron 

Accelerator,  Mass. 
Katherine  J.  Sopka,  Harvard  University 
George  I.  Squibb,  Harvard  University 
Sister  M.  Suzanne  Kelley,  O.S.B.,  Monte  Casino 

High  School,  Tulsa,  Okla. 
Sister  Mary  Christine  Martens,  Convent  of  the 

Visitation,  St.  Paul,  Minn. 


Sister  M.  Helen  St.  Paul,  O.S.F.,  The  Catholic 

High  School  of  Baltimore,  Md. 
M.  Daniel  Smith,  Earlham  College,  Richmond, 

Ind. 
Sam  Standring,  Santa  Fe  High  School,  Santa  Fe 

Springs,  Calif. 
Albert  B.  Stewart,  Antioch  College,  Yellow 

Springs,  Ohio 
Robert  T.  Sullivan,  Burnt  Hills-Ballston  Lake 

Central  School,  N.Y. 
Loyd  S.  Swenson,  University  of  Houston,  Texas 
Thomas  E.  Thorpe,  West  High  School,  Phoenix, 

Ariz. 
June  Goodfield  Toulmin,  Nuffield  Foundation, 

London,  England 
Stephen  E.  Toulmin,  Nuffield  Foundation,  London, 

England 
Emily  H.  Van  Zee,  Harvard  University 
Ann  Venable,  Arthur  D.  Little,  Inc.,  Cambridge, 

Mass. 
W.  O.  Viens,  Nova  High  School,  Fort  Lauderdale, 

Fla. 
Herbert  J.  Walberg,  Harvard  University 
Eleanor  Webster,  Wellesley  College,  Mass. 
Wayne  W.  Welch,  University  of  Wisconsin, 

Madison 
Richard  Weller,  Harvard  University 
Arthur  Western,  Melbourne  High  School,  Fla. 
Haven  Whiteside,  University  of  Maryland,  College 

Park 
R.  Brady  Williamson,  Massachusetts  Institute  of 

Technology,  Cambridge 
Stephen  S.  Winter,  State  University  of  New  York, 

Buffalo 


167 


Brief  Answers  to  Study  Guide  Questions 


Chapter  17 

17.1  Information 

17.2  80.3%  zinc;  19.7%  oxygen 

17.3  47.9%  zinc 

17.4  13.9  times  mass  of  H  atom;  same 

17.5  986  grams  nitrogen;  214  grams  hydrogen 

17.6  9.23  times  mass  of  H  atom 

17.7  (a)  14.1 

(b)  28.2 

(c)  7.0 

17.8  Derivation 

17.9  Na;1     Al;3     P;5    Ca;  2     Sn;4 

17.10  (a)  Ar— K;  Co— Ni;  Te— I;  Th— Pa;  U— Np; 

Es — Fm;  IVId — No 
(b)  Discussion 

17.11  Graph 

17.12  Graph;  discussion 

17.13  8.0  grams;  0.895  gram 

17.14  (a)  0.05  gram  Zn 

(b)  0.30  gram  Zn 

(c)  1.2  gram  Zn 

17.15  (a)  0.88  gram  CI 

(b)  3.14  grams  I 

(c)  Discussion 

(d)  Discussion 

17.16  Discussion 

17.17  Discussion 

17.18  Discussion 

17.19  35.45  grams 

17.20  Discussion 

17.21  Discussion 

17.22  1,3,5 
2,4 


Chapter  18 

18.1  Information 

18.2  (a)  2.0  X  10"  m/sec 
(b)  1.8  X  lOiicoul/kg 

18.3  Proof 

18.4  Discussion 

18.5  Discussion 

18.6  2000  A;  ultraviolet 

18.7  4  X  10-19  joule;  4  X  10-18  joule 

18.8  2.6  X  10-19;  1.6  eV 

18.9  4.9  X  lOiVsec 

18.10  (a)  6  X  lOiVsec 

(b)  4  X  10-19  joule 

(c)  2.5  X  10-0  photons 

(d)  2.5  photons/sec 

(e)  0.4  sec 

(f)  2.5  X  10-10  photon 

(g)  6.25  X  1017  electrons/sec;  0.1  amp 

18.11  1.3  X  101' photons 

18.12  (a)  6.0  X  1023  electrons 

(b)  84  X  10-'  copper  atoms/cm-^ 

(c)  1.2  X  lO-"cm» 

(d)  2.3  X  10-''  cm 

18.13  (a)  2x  =  n\ 

(b)  2x  =  any  odd  number  of  half  wavelengths 

(c)  cos  e  -  2d/\  for  first  order 

18.14  1.2xi0i9/sec 

18.15  Discussion 

18.16  1.2  X  105  volts;  1.9  X  10-"  joule;  1.2  X  10^  eV 

18.17  Glossary 

18.18  Discussion 


Chapter  19 

19.1  Information 

19.2  Discussion 

19.3  Five  listed  in  Text,  but  theoretically  an  Infinite 
number. 

Four  lines  in  visible  region. 

19.4  /7  =  8;  \  =  3880  A 
n  =  10;\  =  3790A 
n  =  12;  \  =  3740  A 

19.5  (a)  Yes 

(b)  n,  =  oo 

(c)  Lyman  series  910  A;  Balmer  series  3650  A; 
Paschen  series  8200  A 

(d)  21.8  X  10-19  joule,  13.6  eV 

19.6  Discussion 

19.7  Discussion 

19.8  2.6  X  10-14  m 

19.9  (a)  Discussion 
(b)  10-V1 

19.10  3.5  m 

19.11  Derivation 

19.12  Discussion 

19.13  List 

19.14  Diagram 

19.15  Discussion 

19.16  Discussion 

19.17  Discussion 

19.18  Discussion 

19.19  Discussion 

19.20  Discussion 

19.21  Discussion 

19.22  Essay 

19.23  Discussion 


Chapter  20 

20.1  Information 

20.2  0.14cor4.2  X  10' m/sec 

20.3  3.7  X  10-14  newtons 

20.4  p  =  m.,v  and  KE  =  m„vV2 

20.5  (a)  Changes  are  too  small 
(b)  1.1  X  10-12  kg 

20.6  (a)  2.7  X  10^3  joules 

(b)  3.0  X  101G  kg 

(c)  5  X  10-"% 

(d)  Rest  mass 

20.7  (a)  1.2  X  10-22  kg  m/sec 

(b)  1.1  X  10-22  kgm/sec 

(c)  2.4  X  10-22  kg  m/sec 

(d)  1.1  X  10-22  kgm/sec 

20.8  p  =  1.7  X  10-27  kg  m/sec;  i/  =  1.9  x  10^  m/sec 

20.9  Discussion 

20.10  Diagram 

20.11  6.6  X  10-5  m/sec 

20.12  3.3  X  10-33  m 

20.13  \  becomes  larger 

20.14  Discussion 

20.15  3  X  10-31  m 

20.16  Discussion 

20.17  (a)  3.3  X  10-25  m/sec 

(b)  5.0  X  10-8  m/sec 

(c)  3.3  X  10-G  m/sec 

(d)  3.3  X  10G  m/sec 

20.18  Discussion 

20.19  Discussion 

20.20  Discussion 

20.21  Discussion 

20.22  Discussion 

20.23  Discussion 

20.24  Discussion 


HOLT,  RINEHART  AND  WINSTON, INC.