Skip to main content

Full text of "Oceanus"

See other formats


Oceanus 


REPORTS  ON  RESEARCH  FROM  THE  WOODS  HOLE  OCEANOGRAPHIC  INSTITUTION 


Vol.  39,  No.  2  •  Fall/Winter  1996  •  ISSN  0029-8182 


Oceans  &  Climate 


^ 


The  Ocean  Conveyor  Belt 
Flows  Around  the  World 


PACIFIC 


SAMW  Subantarctic  Mode  Water 

AAIW  Antarctic  Intermediate  Water 

RSOW  Red  Sea  Overflow  Water 

AABW  Antarctic  Bottom  Water 

NPDW  North  Pacific  Deep  Water 

AAC  Antarctic  Circumpolar  Current 

CDW  Circumpolar  Deep  Water 

NADW  North  Atlantic  Deep  Water 
UPPER  IW        Upper  Intermediate  Water 

IODW  Indian  Ocean  Deep  Water 


One  of  the  keys  to  understanding  how  and  on  what  time  scales  the  vast 
volume  of  water  in  the  ocean  interacts  with  the  atmosphere  and  modifies 
Earth's  climate  is  determining  how  water  moves  from  the  surface  of  the  ocean 
into  the  interior,  how  it  returns  from  the  depths,  and  how  it  flows  between  the 
ocean  basins.  While  many  of  the  articles  in  this  issue  focus  on  the 

movement  of  water  within  the  North  Atlantic,  these 
two  recent  figures  from  WHOI  Senior 
Scientist  Bill  Schmitz  provide  a 
global  synthesis  of  the  present 
understanding  of  the  movement 
of  water  between  ocean  basins 
and  across  the  depths. 

The  numbers  in  the  top  figure 
are  flow  rates  or  transports  in 
units  called  Sverdrups  (after  Nor- 
wegian oceanographer  Harald  U. 
Sverdrup),  which  represent  flow 

at  1,000,000  cubic  meters  per  second.  The  red  arrows  show  flow 
paths  and  rates  in  the  shallow  and  intermediate  depths.  Green  and  blue 
arrows  and  numbers  show  the  paths  and  rates  for  the  deep  ocean  and  for  bot- 
tom flows,  respectively.  The  figure  at  left  provides  a  three-dimensional  perspec- 
tive, labeling  the  different  water  types  moving  along  the  pathways  and  adding 
the  color  orange  for  the  very  salty,  warm  water  that  flows  out  of  the  Red  Sea, 
along  with  the  color  purple  indicating  near-surface  circulations.  These  two 
figures  are  part  of  what  Bill  Schmitz,  who  holds  the  W.  Van  Alan  Clark,  Jr., 
Chair  for  Excellence  in  Oceanography,  calls  his  "final  report,"  a  summary  (in  a 
somewhat  speculative  vein,  he  says)  of  what  he  has  learned  over  the  past  35 
years  about  large-scale,  low-frequency  ocean  currents.  This  two  volume  work  is 
being  published  as  part  of  the  WF1O1  Technical  Report  series. 


Oceanus 


REPORTS  ON  RESEARCH  FROM  THE  WOODS  HOLE  OCEANOGRAPHIC  INSTITUTION 


Vol.  39,  No.  2  •  Fall/Winter  1996  •  ISSN  0029-8182 


Cover:  R/V  Oceania  weathers  a  North  Atlantic  storm  during  a 
March  1981  study  of  a  warm  core  ring  spawned  by  the  Gulf  Stream. 
Inset:  Researchers  aboard  R/V  Endeavor  (University  ot  Rhode  Is- 
land), an  Oceania  sister  ship,  wresde  with  a  rosette  water  sampler 
in  the  southern  Labrador  Sea  during  a  spring  1991  investigation  of 
the  origins  of  the  deep  western  boundary  current. 
Urge  photo  by  lames  McCarthy.  Harvard  University  Inset  by  Peter  Undry.  WHOI 

Oceanus  is  published  semi-annually  by  the  Woods  Hole 
Oceanographic  Institution,  Woods  Hole,  MA  02543.  508-289-3516. 
http  //www  whoi  edu/oceanus 


Oceania  and  its  logo  are  «  Registered  Trademarks  of  the  Woods 
Hole  Oceanographic  Institution,  All  Rights  Reserved. 

A  calendar-year  Oceanus  subscription  is  available  for  $15  in  the 
US,  $18  in  Canada.  The  WHOI  Publication  Package,  including 
Oceania  magazine  and  Woods  Hole  Currents  (a  quarterly  publication 
for  WHOI  Associates  and  Friends)  is  available  for  a  $25  calendar- 
year  fee  in  the  US,  $30  in  Canada.  Outside  North  America,  the 
annual  fee  for  Oceanus  magazine  only  is  $25,  and  the  Publication 

Package  costs  $40  To  receive  the  publications,  please  call 

(toll  free)  1-800-291-6458,  or  write:  WHOI  Publication  Services, 

P.O.  Box  50145,  New  Bedford,  MA  02745-0005. 


To  purchase  single  and  back-issue  copies  of  Oceanus,  please 
contact:  lane  Hopewood,  WHOI-MS#5,  Woods  Hole  MA  02543. 
Phone:  508-289-3516.  Fax:  508-457-2182 


Checks  should  be  drawn  on  a  US  bank  in  US  dollars  and  made 
payable  to  Woods  Hole  Oceanographic  Institution. 

When  sending  change  of  address,  please  include  mailing  label. 

Claims  for  missing  numbers  from  the  US  will  be  honored  within 

three  months  of  publication;  overseas,  six  months. 


Permission  to  photocopy  for  internal  or  personal  use  or  the 

internal  or  personal  use  of  specific  clients  is  granted  by  Oceanus  to 

libraries  and  other  users  registered  with  the  Copyright  Clearance 

Center  (CCC),  provided  that  the  base  fee  of  $2  per  copy  of  the 

article  is  paid  directly  to:  CCC,  222  Rosewood  Drive,  Danvers,  MA 

01923  Special  requests  should  be  addressed  to  theOcomus  editor 


Oceans  &  Climate 


Oceans  &  Climate  2 

Tlie  Ocean's  Role  in  Climate  &  Climate  Change 
By  Michael  S.  McCartney 

If  Rain  Falls  On  the  Ocean,  Does  It  Make  a  Sound?  4 

Fresh  Water's  Effect  on  Ocean  Phenomena 
By  Raymond  W.  Schmitt 

ALACE,  PALACE,  Slocum  6 

A  Dj'misrr  of  Free  Floating  Oceanographic  Instruments 
By  Raymond  W.  Schmitt 

Alpha,  Bravo,  Charlie...  9 

Ocean  Weather  Ships  1940-1980 
By  Robertson  P.  Dinsmore 

A  Century  of  N.  Atlantic  Data  Indicates  Interdecadal  Change      11 

Surface  Temperature,  Winds,  &  Ice  in  the  North  Atlantic 
By  Clara  Deser 

North  Atlantic  Oscillation  13 

By  Michael  S.  McCartney 

The  Bermuda  Station  S  14 

A  Long-Running  Oceanographic  Show 
By  Terrence  M.  loyce  and  Lynne  Talley 

Sedimentary  Record  Yields  Several  Centuries  of  Data  16 

Tlie  Little  Ice  Age  and  Medieval  Warm  Period  in  the  Sargiisso  Sea 
By  Lloyd  D.  Keigwin 

North  Atlantic's  Transformation  Pipeline  19 

It  Chills  and  Redistributes  Subtropical  Water 

By  Michael  S.  McCartney,  Ruth  G.  Curry,  and  Hugo  F.  Bezdek 

Labrador  Sea  Water  Carries  Northern  Climate  Signal  South     24 

Subpolar  Signals  Appear  Years  Later  at  Bermuda 
By  Ruth  G.  Curry  and  Michael  S.  McCartney 

Transient  Tracers  Track  Ocean  Climate  Signals  29 

By  William  I.  lenkins  and  William  M.  Smethie,  Jr. 

New  Data  on  Deep  Sea  Turbulence  33 

Shedding  Light  on  Verticil/  Mixing 
By  John  M.  Toole 

Computer  Modelers  Simulate  Real  and  Potential  Climate       36 
Combining  Equations  and  Data  Pushes  Computers'  Limits 
By  Rui  Xin  Huang  and  liayan  Yang 

The  El  Nino/Southern  Oscillation  Phenomenon  39 

Seeking  /te  "Trigger"  and  Working  Toward  Prediction 
By  Lewis  M.  Rothstein  and  Dake  Chen 


1930 


Editor:  Vicky  Cullen  •  Designer:  Jim  Canavan 
Woods  Hole  Oceanographic  Institution 

Robert  B.  Gagosian,  Director 
Frank  V.  Snyder,  Chairman  of  the  Board  of  Trustees 

lames  M.  Clark,  President  of  the  Corporation 
Robert  D.  Harrington,  Jr.,  President  of  the  Associates 


Woods  Hole  Oceanographic  Institution  is  an  Equal  Employment  Opportunity  and  Affirmative  Action  Employer 


Printed  on 
recycled  paper 


OCEANUS 


90N 


60N 


2 

1.5 

1 

0.5 

0 

-0.5 

-1 

-1.5 

-2 


90S 


120E 


Annual  surface 
temperature 
change  in  degrees 
Centigrade  for  the 
period  1975-1994 
relative  to  1955- 
1974.  This  figure, 
prepared  for  the 
1996  Intergovern- 
mental Panel  on 
Climate  Change, 
indicates  that 
Earth's  surface  has 
been,  on  average, 
warmer  (predomi- 
nating orange) 
over  the  past  20 
years  compared  to 
the  preceding  20 
years.  The  cooler 
blue  areas  show, 
however,  that  the 
warming  has  not 
been  universal. 


60E 


120E 


Our  thanks  to 

Senior  Scientist 

Robert  A.  Weller  for 

editorial  assistance 

with  this  issue. 


Oceans  &  Climate 

The  Ocean's  Role  In  Climate  &  Climate  Change 


Michael  S.  McCartney 

Senior  Scientist,  Physical  Oceanography  Department 

The  past  decade  has  brought  rapid  scientific 
progress  in  understanding  the  role  of  the 
ocean  in  climate  and  climate  change.  The 
ocean  is  involved  in  the  climate  system  primarily 
because  it  stores  heat,  water,  and  carbon  dioxide, 
moves  them  around  on  the  earth,  and  exchanges 
these  and  other  elements  with  the  atmosphere. 
Three  important  premises  of  the  oceans  and  cli- 
mate story  are: 

•  The  ocean  has  a  huge  storage  capacity  for 
heat,  water,  and  carbon  dioxide  compared  to 
the  atmosphere. 

•  Global  scale  oceanic  circulation  transports  heat, 
water,  and  carbon  dioxide  horizontally  over  large 
distances  at  rates  comparable  to  atmospheric  rates. 

•  The  ocean  and  atmosphere  exchange  as  much 
heat,  water,  and  carbon  dioxide  between  them  as 
each  transports  horizontally. 

The  ocean  and  atmosphere  are  coupled — their 
"mean  states,"  evolution,  and  variability  are 
linked.  Ocean  currents  are  primarily  a  response 
to  exchanges  of  momentum,  heat,  and  water 
vapor  between  ocean  and  atmosphere,  and  the 
resulting  ocean  circulation  stores,  redistributes, 
and  releases  these  and  other  properties.  The  at- 


mospheric part  of  this  coupled  system  exhibits 
variability  through  shifts  in  intensity  and  loca- 
tion of  pressure  centers  and  pressure  gradients, 
the  storms  that  they  spawn  and  steer,  and  the 
associated  distributions  of  temperature  and  water 
content.  Oceanic  variability  includes  anomalies 
of  sea  surface  temperature,  salinity,*  and  sea  ice, 
as  well  as  of  the  internal  distribution  of  heat  and 
salt  content,  and  changes  in  the  patterns  and 
intensities  of  oceanic  circulation.  These  coupled 
ocean-atmosphere  changes  may  impact  the  land 
through  phases  of  drought  and  deluge,  heat  and 
cold,  and  storminess. 

One  example  of  coupled  ocean-atmosphere 
variability  is  the  El  Nino/Southern  Oscillation  or 
ENSO  (see  article  on  page  39).  The  appearance 
of  warm  water  at  the  ocean's  surface  in  the  east- 
ern tropical  Pacific  off  South  America  has  a  dra- 

*Many  of  this  issue's  articles  discuss  the  physical  properties 
of  seawater.  The  density  of  seawater  changes  with  tempera- 
ture (measured  in  °C),  salinity  (measured  in  parts  per 
thousand  or  grams  of  salt  per  kilogram  of  water — typically 
given  without  units,  such  as  simply  34.9),  and  pressure. 
The  density  of  seawater  (p)  in  kilograms  per  cubic  meter  is 
close  to  and  slightly  larger  than  1,000  kilograms  per  cubic 
meter.  "Potential  density,"  (a),  is  the  value  of  the  relative 
density  if  the  seawater  is  brought  to  the  surface  without  ex- 
changing heat  on  its  way  up.  This  expression  helps  ocean- 
ographers  understand  the  water  column's  stability. 


FALL/WINTER  1996 


matic  impact  on  weather  and  seasonal-to- 
interannual  climate.  Considerable  effort  has 
been  dedicated  to  developing  the  ability  to  pre- 
dict ENSO,  including  deployment  and  mainte- 
nance of  buoys  and  other  observational  systems 
in  the  tropical  Pacific  and  sustained  attention  to 
improving  models  of  ENSO.  However,  ENSO  is 
but  one  of  the  mechanisms  by  which  the  ocean 
and  atmosphere  influence  one  another.  Such 
coupling  occurs  on  many  time  scales,  even  over 
centuries  (see  "Sedimentary  Record"  article  on 
page  16).  There  is  growing  interest  among  the 
oceanographic  community  in  developing  a  bet- 
ter understanding  of  the  ocean's  role  in  climate 
changes  on  decadal  to  centennial  time  scales, 
and  many  of  the  articles  in  this  issue  focus  on 
such  variability  in  the  North  Atlantic  Ocean. 

There  are,  as  yet,  no  continuing  observations 
dedicated,  as  the  observing  network  in  the  tropi- 
cal Pacific  is  to  ENSO,  to  monitoring,  under- 
standing, and  predicting  decadal  climate  variabil- 
ity involving  ocean-atmosphere  interaction.  Our 
challenges  are  to  learn  from  what  observations 
and  modeling  have  been  done  and  to  develop 
strategies  for  future  work. 

Sustained  observations  allow  scientists  to 
detect  climatic  spatial  patterns.  For  example,  the 
figure  opposite  shows  interdecadal  change  in 
land  and  sea  surface  temperatures.  This  figure  is 
taken  from  the  1996  Intergovernmental  Panel  on 
Climate  Change  (IPCC)  report,  a  huge  effort  of 
the  international  climate  research  community  to 
assess  Earth's  climatic  state  every  five  years.  The 
predominating  orange  indicates  that  the  earth's 
surface  has  been,  on  average,  warmer  the  past  20 
years  compared  to  the  preceding  20  years.  Signifi- 
cant blue  areas,  principally  over  the  oceans,  show 
that  the  warming  has  not  occurred  everywhere: 
Large  areas  of  the  subpolar  North  Atlantic  are 
cold,  sandwiched  between 
warm  northern  North 
America  and  northern 
Eurasia,  and  the  North  Pa- 
cific is  also  cold,  but  with  a 
subtropical  emphasis  rather 
than  a  subpolar  emphasis. 

The  figure  above  right 
puts  a  longer  time  perspec- 
tive on  the  warming  by 
showing  the  hemispheric 
and  global  average  tempera- 
ture over  the  past  135  years, 
the  rough  limit  of  useful 
sustained  measurements. 
These  curves  show  the  overall 
global  warming  beginning 
with  the  industrial  age,  but 
note  the  roughly  60  year 
oscillation  this  century,  par- 


04 

5  o.o 
i  -02 

^  -0.4 
-0.6 

04 
0.2 
00 
-02 

-04 
-0.6 


Southern  Hemisphere 


Globe 


1860         1880 


1900 


1920 


Year 


ticularly  in  the  north- 
ern hemisphere,  show- 
ing steeper  warming 
trends  1910-1940/ 
1945  and  1975-1995. 
Time  series  like  these 
lie  at  the  heart  of 
controversies  about 
global  warming  as  a 
trend  versus  as  a  phase 
of  some  mode  of 
"natural"  climate 
variability. 

Continued  sus- 
tained measurements 
of  a  broad  array  of 
climate  indicators  will 
eventually  directly 
answer  key  questions:  Is  the  steep  temperature 
rise  of  the  past  20  years  the  portent  of  a  crisis:  a 
rise  that  will  continue  through  the  next  century 
and  evolve  into  an  increasingly  major  climate 
perturbation?  Or  is  the  steep  rise  "just"  a  phase  of 
a  natural  oscillation  of  the  climate  system  super- 
imposed on  a  less  severe  warming?  Or  is  the 
entire  warming  trend  of  the  past  135  years  itself 
just  the  warming  phase  of  a  still  longer  natural 
oscillation?  There  is  a  preponderance  of  scientific 
judgement,  as  carefully  compiled  and  described 
by  the  IPCC,  that  the  answer  will  be  somewhere 
between  the  first  two  possibilities,  and  that  this  is 
caused  by  human  impact  on  the  climate  system. 

This  issue  of  Oceania  emphasizes  the  North 
Atlantic  Ocean,  but,  to  answer  these  scientific 
questions,  we  must  also  take  on  the  challenges  of 
filling  in  many  sparsely  sampled  regions,  build- 
ing on  the  ENSO  work  in  the  Pacific  and  decadal 
variability  research  in  the  North  Atlantic,  and 
working  toward  understanding  on  a  global  basis. 


1940         1960         1980     2000 

Hemispheric  and 
global  average  tem- 
perature for  the 
past  135  years. 


Scientists  aboard 
R/V  Knorr  launch  a 
rosette  water  sam- 
pler and  conduc- 
tivity/temperature/ 
depth  instrument. 
Much  of  the  data 
discussed  in  this 
issue  was  collected 
by  such  equip- 
ment. Author 
McCartney  is  the 
fellow  getting  wet 
at  top  left. 


The  Great  Salinity 
Anomaly,  a  large, 
near-surface  pool 
of  fresher-than- 
usual  water,  was 
tracked  as  it  trav- 
eled in  the  subpo- 
lar gyre  currents 
from  1968  to  1982. 


If  Rain  Falls 
On  the  Ocean 

Does  It  Make  a  Sound? 

Fresh  Water's  Effect  on  Ocean  Phenomena 


Raymond  W.  Schmitt 

Senior  Scientist,  Physical  Oceanography 

As  with  similar  questions  about  a  tree  in  the 
forest  or  a  grain  of  sand  on  the  beach,  it 
may  be  hard  to  imagine  that  a  few  inches 
of  rain  matters  to  the  deep  ocean.  After  all,  the 
ocean's  average  depth  is  around  4  kilometers  and 
only  1  to  5  centimeters  of  water  are  held  in  die 
atmosphere  at  any  one  time.  But  it  does  matter,  in 
part  because  the  ocean  is  salty.  The  effect  of  rain 
diluting  the  salts  in  the  ocean  (or  evaporation 
concentrating  them)  can  be  greater  than  the  effect 
of  heating  (or  cooling)  on  the  density  of  seawater. 


50 


40 


It  also  matters  because  rainfall  and  evapora- 
tion are  not  evenly  distributed  across  and  among 
ocean  basins — some  regions  continuously  gain 
water  while  others  continuously  lose  it.  This 
leads  to  ocean  current  systems  that  can  be  sur- 
prisingly strong.  The  processes  of  evaporation 
and  precipitation  over  the  ocean  are  a  major  part 
of  what  is  called  "the  global  water  cycle;"  indeed, 
by  all  estimates,  they  dominate  the  water  cycle 
over  land  by  factors  often  to  a  hundred.  The 
addition  of  just  one  percent  of  Atlantic  rainfall  to 
die  Mississippi  River  basin  would  more  than 
double  its  discharge  to  die  Gulf  of  Mexico. 

As  discussed  previ- 
ously in  Oceanus,  our 
knowledge  of  the  water 
cycle  over  the  ocean  is 
extremely  poor  (see  the 
Spring  1992  issue).  Yet 
we  now  realize  that  it  is 
one  of  the  most  impor- 
tant components  of  the 
climate  system.  One  of 
the  significant  pieces  of 
evidence  for  this  comes 
from  a  description  of 
die  "Great  Salinity 
Anomaly"  put  together 
by  Robert  Dickson 
(Fisheries  Laboratory, 
Suffolk,  England)  with 
other  European  ocean- 
ographers.  The  Great 
Salinity  Anomaly 
(GSA)  can  be  character- 
ized as  a  large,  near- 
surface  pool  of  fresher 
water  that  appeared  off 
the  east  coast  of 
Greenland  in  the  late 
1960s  (see  figure  at 
1  left).  It  was  carried 
5  around  Greenland  and 
10°          into  the  Labrador  Sea 


FALL/WINTER  1996 


34 


from  Lazier,  1980 
_J I 


by  the  prevailing  35 

ocean  currents,  in  the 
counterclockwise  cir- 
culation known  as  the 
subpolar  gyre.  It  hov-  Q 

ered  off  Newfound-  0^? 

land  in  1971 -72  and  £ 

was  slowly  carried  5 

back  toward  Europe  in        *° 
the  North  Atlantic 
Current,  which  is  an 
extension  of  the  Gulf 
Stream.  It  then  com- 
pleted its  cycle  and 
was  back  off  the  east 
coast  of  Greenland  by  1964  1966 

the  early  1980s, 

though  reduced  in  size  and  intensity  by  mixing 
with  surrounding  waters.  The  origin  of  the  Great 
Salinity  Anomaly  is  thought  to  lie  in  an  unusu- 
ally large  discharge  of  ice  from  the  Arctic  Ocean 
in  1967.  Its  climatic  importance  arises  from  the 
impact  it  had  on  ocean-atmosphere  interaction 
in  the  areas  it  traversed. 

The  GSA  derives  its  climate  punch  from  the 
strong  effect  of  salinity  on  seawater  density,  with 
salty  water  being  considerably  denser  than  fresh 
water.  That  is,  these  northern  waters  normally 
experience  strong  cooling  in  the  winter,  which 
causes  the  surface  water  to  sink  and  mix  with 
deeper  waters.  This  process,  called  deep  convec- 
tion (see  figure  below),  is  a  way  for  the  ocean  to 


Salinity  in  the  Labrador  Sea 

from  Ocean  Weather  Station  Bravo 


1968 


1970 


1972 


Normal  Ocean  in  Winter 


large 
heat  loss 


release  heat  to  the  atmosphere,  heat  that  then 
helps  to  maintain  a  moderate  winter  climate  for 
northern  Europe.  However,  when  the  GSA  passed 
through  a  region,  the  surface  waters  became  so 
fresh  and  light  that  even  strong  cooling  would 
not  allow  it  to  convert  into  the  deeper  waters. 
Thus,  the  deep  water  remained  isolated  from  the 
atmosphere,  which  could  not  extract  as  much 
heat  as  usual  from  the  ocean.  The  GSA  acted  as  a 
sort  of  moving  blanket,  insulating  different  parts 
of  the  deep  ocean  from  contact  with  the  atmo- 
sphere as  it  moved  around  the  gyre.  Its  impact  in 
the  Labrador  Sea  has  been  particularly  well  docu- 
mented (see  "Labrador  Sea"  article  on  page  24). 
When  the  surface  waters  were  isolated  from  deep 
waters,  they  became 


Ocean  in  Winter  with 
Fresh  Surface  Anomaly 


Deep  convection  is  a  key  component  of  the  ocean's  role  in  Earth's  climate.  Strong  win- 
ter cooling  of  surface  waters  causes  them  to  become  denser  than  water  below  them, 
which  allows  them  to  sink  and  mix  with  deeper  water.  This  process  releases  heat  from 
the  overturned  water  to  the  atmosphere  and  maintains  northern  Europe's  moderate 
winter  climate.  The  Great  Salinity  Anomaly  interrupted  this  process  as  its  pool  of 
fresher  water  prevented  convection. 


cooler.  Changing  sea 
surface  temperature 
patterns  can  affect  at- 
mospheric circulation, 
and  may  possibly  rein- 
force a  poorly  under- 
stood, decades-long 
variation  in  North  At- 
lantic meteorological 
conditions  known  as 
the  North  Atlantic  Os- 
cillation (see  box  on 
page  13).  For  it  is  the 
ocean  that  contains  the 
long-term  memory  of 
the  climate  system.  By 
comparison,  the  atmo- 
sphere has  hardly  any 
thermal  inertia.  It  is 
3  difficult  to  imagine 
|  how  the  atmosphere 
alone  could  develop  a 
regular  decadal  oscilla- 
tion, but  the  advection 
of  freshwater  anomalies 
by  the  ocean  circulation 


Salinity  as  a  func- 
tion of  time  at  10 
meters,  200  meters, 
and  1,000  meters 
depth  as  recorded 
at  Ocean  Weather 
Station  Bravo  (see 
map  on  page  10)  in 
the  Labrador  Sea. 
Deep  convection  is 
possible  when  the 
salinity  difference 
between  shallow 
and  deep  water  is 
small.  This  nor- 
mally occurs  every 
winter.  However, 
from  1968  to  1971, 
the  presence  of  the 
fresh,  shallow, 
Great  Salinity 
Anomaly  prevented 
deep  convection. 
Unfortunately, 
Weather  Station 
Bravo  is  no  longer 
maintained.  Scien- 
tists will  need  to 
use  new  technol- 
ogy like  the  PAL- 
ACE float  (see  Box 
overleaf)  in  order 
to  reestablish  such 
time  series.  Such 
data  is  essential  for 
understanding  the 
role  of  freshwater 
anomalies  in  the 
climate  system. 


OCEANUS 


could  be  an  important  key  to  this  climate  puzzle. 

Unfortunately,  we  have  no  ready  means  of 
detecting  freshwater  pulses  like  the  GSA.  While 
surface  temperature  can  be  observed  easily  from 
space,  surface  salinity,  so  far,  cannot.  The  salinity 
variations  important  for  oceanography  require 
high  precision  and  accuracy,  so  there  is  no  quick 
and  inexpensive  method  of  measurement.  We 
have  had  to  rely  on  careful  analysis  of  sparse 


historical  records  from  mostly  random  and  unre- 
lated surveys  gleaned  from  several  nations  to 
piece  the  GSA's  story  together.  But  how  many 
other  "near-great"  salinity  anomalies  have  we 
missed  because  the  signal  was  not  quite  large 
enough?  Is  there  a  systematic  way  to  monitor 
salinity  so  that  we  know  years  in  advance  of  an- 
other GSA's  approach? 

In  addition  to  variability  within  an  ocean  ba- 


ALACE,  PALACE,  Slocum 

A  Dynasty  of  Free  Floating  Oceanographic  Instruments 


Autonomous  diving  floats  have  been  developed  by 
Doug  Webb  of  Webb  Research,  Inc.  in  Falmouth, 
MA,  in  conjunction  with  Russ  Davis  of  the 
Scripps  Institution  of  Oceanography.  The  Profiling  Au- 
tonomous LAgrangian  Circulation  Explorer  (PALACE)  is 
a  free  float  that  drifts  with  the  currents  at  a  selected 
depth,  much  like  a  weather  balloon  drifts  with  the  winds. 
At  preset  time  intervals 
(typically  one  or  two 
weeks)  it  pumps  up  a 
small  bladder  with  oil 
from  an  internal  reser- 
voir, which  increases  its 
volume,  but  not  its 
mass,  and  causes  it  to 
rise  to  the  surface.  On 
the  way  up  it  records 
temperature  and  salinity 
as  a  function  of  depth. 
Once  at  the  surface  it 
transmits  the  data  to  a 
satellite  system  that  also 
determines  its  geo- 
graphical position.  The 
drift  at  depth  between 
fixes  provides  an  esti- 
mate of  the 

"Lagrangian"  velocity  at 
that  time  and  place  (as 


Doug  Webb  was  photographed  on  a  catwalk  above  a  test  tank  used  to  put 


the  Slocum  glider  through  its  paces, 
opposed  to  "Eulerian" 

measurements  of  the  velocity  past  a  fixed  point.  These 
names  derive  from  18th  century  mathematicians  who 
originated  these  ways  of  looking  at  fluid  flows). 

The  basic  technology  of  the  float  has  been  used  for 
several  years  in  the  nonprofiling  ALACE,  which  simply 
provides  velocity  information.  Hundreds  of  ALACES  have 
been  successfully  deployed  in  the  Pacific  and  Indian 
Oceans.  A  program  to  release  a  large  number  of  PALACES 
in  the  Atlantic  is  just  getting  underway. 

The  use  of  the  ALACE  as  a  platform  for  salinity  mea- 


surements is  not  without  problems.  The  slow  rising  mo- 
tion, and  low  power  available,  limit  the  type  of  sensor 
that  can  be  deployed.  The  problem  of  sensor  drift  due  to 
biological  fouling  may  be  severe  in  some  regions,  and 
methods  to  prevent  fouling  are  just  being  developed. 
However,  because  the  float  spends  most  of  its  life  in  a 
deep  and  climatically  stable  water  mass,  not  subject  to 

near-surface  atmospheric 
variations,  we  should  be 
able  to  compensate  for 
any  drifts. 

But  the  fact  that  these 
floats  move  around  is 
something  of  a  drawback 
if  the  objective  is  to 
monitor  ocean  tempera- 
ture and  salinity.  That  is, 
in  the  long  run,  we 
would  rather  that  they 
stayed  put  and  measured 
the  properties  in  one 
place.  Such  a  task  could 
be  achieved  if  the  float 
were  capable  of  gliding 
_  horizontally  and  turning 
I  as  it  rose.  The  horizontal 
~\  displacement  achieved 
-  could  be  directed  to 
maintain  one  position, 
with  each  excursion 

compensating  for  the  drift  caused  by  ocean  currents.  With 
Navy  funding,  Webb,  Davis,  and  Breck  Owens  (WHOI) 
are  currently  working  on  such  a  gliding  float  (see  photo). 
All  these  floats  depend  on  batteries  to  power  the  elec- 
tronic sensors,  the  pump  that  varies  ballast,  and  the  trans- 
mitter that  sends  data  to  the  satellite.  The  battery  life  is 
around  two  years,  depending  on  the  frequency  of  profil- 
ing and  transmitting.  One  way  to  extend  its  life  is  to  use 
the  ocean's  vertical  temperature  differences  to  run  a 
simple  heat  engine.  Doug  Webb  has  another  type  of  float 


FALLWINTER  1996 


sin,  we  would  like  to  understand  the  large  dillc-i- 
ences  in  salt  concentration  among  ocean  basins, 
(see  figure  on  next  page  )  For  example,  the  Pacific 
Ocean  is  significantly  fresher  than  the  Atlantic 
and,  because  it  is  lighter,  stands  about  halt  a 
meter  higher.  This  height  difference  drives  the 
flow  of  Pacific  water  into  the  Arctic  through  the 
Bering  Strait.  The  salinity  difference  between 
these  two  major  oceans  is  thought  to  be  caused 


by  the  transport  of  water  vapor  across  Central 
America:  The  trade  winds  evaporate  water  from 
the  surface  of  the  Atlantic,  carry  it  across  Central 
America,  and  supply  rainfall  to  the  tropical  Pa- 
cific. This  water  loss  is  the  major  cause  of  the 
Atlantic's  greater  saltiness  and  its  propensity  to 
form  deep  water.  The  extra  rainfall  on  the  Pacific 
makes  it  fresher  and  prevents  deep  convection. 
How  does  this  atmospheric  transport  vary  with 


with  such  a  propulsion  system.  It  uses  a  waxy  material 
that  expands  when  it  melts  at  around  50  degrees,  a  tem- 
perature the  float  encounters  at  several  hundred  meters 
depth  on  each  trip  to  and  from  the  surface.  This  expan- 
sion is  used  to  store  energy  to  pump  ballast  when  needed. 
Use  of  this  "free"  energy  for  propulsion  reduces  the  load 
on  the  batteries  and  extends  the  life  of  the  float.  The  ther- 
mal ballasting  engine  has  been  tested  extensively  in  the 
lab  and  recently  deployed  off  Bermuda  in  a  nongliding 
float,  where  it  performed  over  120  depth  cycles. 

Doug  Webb's  dream  is  to  marry  the  thermal  engine 
with  the  glider,  and  thus  make  a  long-lived,  roving  (or 
station-keeping)  autonomous  profiler  possible.  Years  ago 
he  described  the  technical  possibilities  to  the  late  Henry 
Stommel,  who  developed  a  vision  of  how  such  an  instru- 
ment might  be  deployed  in  large  numbers  around  the 
globe  (see  Oceanus,  Winter  1989/90).  They  called  the 
instrument  Slocum,  with  the  idea  that  it  could  circumnavi- 
gate the  globe  under  its  own  power,  like  New  Englander 
Joshua  Slocum,  the  first  solo  sailor  to  perform  that  feat. 
The  Internet  could  allow  scien- 
tists to  monitor  Slocum  data 
from  their  home  laboratories 
around  the  world. 

If  we  deploy  enough 
Slocums,  their  data  should  be 
as  valuable  for  predicting 
global  climate  on  seasonal  to 
decadal  time  scales  as  satel- 
lites and  weather  balloons  are 
for  forecasting  the  daily 
weather.  Indeed,  one  of 
Slocum's  key  attractions  is 
that  it  is  inexpensive  enough 
to  deploy  in  large  numbers. 
Per-profile  costs  for  both 
temperature  and  salinity  are 
expected  to  be  $50  or  less, 
once  a  mature  system  is  oper- 
ating— vastly  cheaper  than 
anything  possible  using  ships. 
A  globe-spanning  array  of 
1,000  Slocums  would  cost  less 
than  a  new  ship,  yet  provide 
an  unprecedented  view  into 
the  internal  workings  of  the 
global  ocean.     — Ray  Schmitt 


MIT/WHOI  loint  Program  student  Steve  Jayne  holds  an  ALACE 
(Autonomous  LAgrangian  Circulation  Explorer)  float  aboard 
R/V  Knorr  during  a  1 995- 1 996  (yes,  Christmas  at  sea)  cruise  for 
the  World  Ocean  Circulation  Experiment  in  the  Indian  Ocean. 


About  1  day  at  surface 


Recording  temperature 
and  salinity  as  it  rises 


1,000m 


Drifting 
1  week 


During  a  data  collection  and  reporting  cycle,  a  PALACE  (Profiling  Autonomous  LAgrangian  Circu- 
lation Explorer)  float  drifts  with  the  current  at  a  programmed  depth,  rises  every  week  or  two  by  in- 
flating the  external  bladder  (recording  temperature  and  salinity  profiles  on  the  way  up),  spends  a 
day  at  the  surface  transmitting  data,  then  returns  to  drift  at  depth  by  deflating  the  bladder. 


The  average  surface 
salinity  distribu- 
tion in  the  global 
ocean,  as  compiled 
from  many  indi- 
vidual ship  mea- 
surements, mostly 
during  this  century. 
The  figure  also 
shows  the  approxi- 
mate coverage  ob- 
tainable with  an  ar- 
ray of  about  1,000 
Slocums  or  PAL- 
ACES. These  would 
resolve  the  large 
scale  features  of  the 
salinity  field  and 
provide  completely 
new  information 
on  its  variability 
with  time.  The  ar- 
ray would  be  an 
early  warning  sys- 
tem for  the  Great 
Salinity  Anomalies 
of  the  future. 


time?  Since  salinity  is  a  good  indicator  of  the 
history  of  evaporation  or  precipitation,  perhaps  if 
we  had  sufficient  data,  we  could  see  changes  in 
the  upper  ocean  salt  content  of  the  two  oceans 
that  reflect  variations  in  atmospheric  transports. 
How  many  years  does  it  take  for  salinity  anoma- 
lies in  the  tropical  Atlantic  to  propagate  to  high- 
latitude  convection  regions  and  affect  the  sea- 
surface  temperature  there?  What  is  the  impact  on 
the  atmospheric  circulation? 

These  and  other  climate  problems  will  con- 
tinue to  perplex  us  until  we  make  a  serious  at- 
tempt to  monitor  salinity  on  large  space  and  time 
scales.  One  approach  would  be  to  maintain  ships 
in  certain  places  to  sample  the  ocean  continually. 
A  modest  effort  along  these  lines  was  made  after 
World  War  II  when  weather  ships  were  main- 
tained at  specific  sites  by  several  nations  (see 
following  article).  The  data  they  collected  provide 
nearly  the  only  long  time-series  measurements 
available  from  deep-ocean  regions.  However,  the 
weather  ships  are  all  but  gone;  there  is  only  one 
now,  maintained  seasonally  by  the  Norwegians. 
Today's  satellites  provide  information  on  ap- 
proaching storm  systems,  but,  unfortunately,  they 
cannot  tell  us  what  we  need  to  know  about  ocean 
salinity  distributions. 

It  now  appears  that  new  technology  will  pro- 
vide the  key  to  the  salinity  monitoring  problem, 
at  a  surprisingly  modest  cost.  The  Box  on  pages  6 
and  7  describes  how  we  might  obtain  tempera- 
ture and  salinity  profiles  from  data  collected  by 
autonomous  diving  floats.  It  should  be  quite 


feasible  to  deploy  an  array  of  these  station-keep- 
ing "Slocums"  that  would  intercept  and  monitor 
the  progress  of  the  "Great  Salinity  Anomalies"  of 
the  future.  In  the  next  two  years,  a  large  number 
of  profiling  ALACE  (precursor  to  the  Slocum) 
floats  will  be  deployed  in  the  Atlantic  in  a  pre- 
liminary test  of  the  general  concept.  In  addition 
to  measuring  temperature  and  salinity,  Slocums 
might  some  day  measure  rain.  It  turns  out  that 
rain  falling  on  the  ocean  does  make  a  sound,  and 
work  is  underway  to  record  that  sound  with  hy- 
drophones and  develop  algorithms  to  convert  the 
measured  sound  level  to  rain  rates.The  remaining 
technical  obstacles  to  development  of  a  globe- 
spanning  array  of  station-keeping  Slocums  are 
small.  The  only  thing  lacking  is  a  strong  societal 
commitment  to  the  support  of  such  fundamental 
research  on  the  climate  system  of  the  earth. 

This  research  was  sponsored  by  the  National  Science 
Foundation  and  the  National  Oceanic  and  Atmospheric 
Administration's  Climate  and  Global  Change  Program. 

Mast  of  Ray  Schmitt's  career  has  been  focused  on  very  small- 
scale  processes  in  the  ocean  related  to  mixing  by  turbulence  and 
"salt  fingers."  Hoivevcr,  he  has  been  driven  toward  studies  of  the 
global-scale  hydrologic  cycle  by  a  desire  to  contribute  to  im- 
proved weather  and  climate  prediction,  so  that  he  can  better 
plan  to  take  advantage  of  the  rare  good  weather  in  Woods  Hole. 


,-. 


FALL/WINTER  1996 


Alpha,  Bravo,  Charlie 

Ocean  Weather  Ships  1940-1980 


Robertson  P.  Dinsmore 

WHOI  Marine  Operations 

The  ocean  weather  station  idea  originated  in 
the  early  days  of  radio  communications 
and  trans-oceanic  aviation.  As  early  as 
1921,  the  Director  of  the  French  Meteorological 
Service  proposed  establishing  a  stationary 
weather  observing  ship  in  the  North  Atlantic  to 
benefit  merchant  shipping  and  the  anticipated 
inauguration  of  trans-Atlantic  air  service.  Up  to 
then,  temporary  stations  had  been  set  up  for 
special  purposes  such  as  the  US  Navy  NC-4  trans- 
Atlantic  flight  in  1919  and  the  ill-fated  Amelia 
Earhart  Pacific  flight  in  1937. 

The  loss  of  a  PanAmerican  aircraft  in  1938  due 
to  weather  on  a  trans-Pacific  flight  prompted  the 
Coast  Guard  and  the  Weather  Bureau  to  begin 
tests  of  upper  air  observations  using  instru- 
mented balloons.  Their  success  resulted  in  a 
recommendation  by  Commander  E.  H.  Smith  of 
the  International  Ice  Patrol  (and  future  Director 
of  the  Woods  Hole  Oceanographic  Institution) 
for  a  network  of  ships  in  the  Atlantic  Ocean. 
World  War  II  brought  about  a  dramatic  in- 
crease in  trans-Atlantic  air  navigation,  and  in 
January  1940  President  Roosevelt  established  the 
"Atlantic  Weather  Observation  Service"  using 
Coast  Guard  cutters  and  US  Weather  Bureau  ob- 
servers. Most  flights  at  this  time  were  using  south- 


ern routes.  On  February  10,  1940,  the  327-foot 
cutters  Bibb  and  Duane  occupied  Ocean  Stations  1 
and  2 — the  forerunners  of  Stations  D  and  E  (see 
chart  on  next  page). 

With  the  US  entering  the  war,  Coast  Guard 
cutters  were  diverted  to  anti-submarine  duties, 
and  the  weather  stations  were  taken  over  by  a 
motley  assortment  of  vessels  ranging  from  con- 
verted yachts  to  derelict  freighters,  mostly  Coast 
Guard  operated.  As  trans-Atlantic  air  traffic  in- 
creased, so  did  the  number  of  weather  and  plane 
guard  stations.  The  role  of  weather  during  the 
Battle  of  Coral  Sea  and  trans-Pacific  flights  re- 
sulted in  stations  being  set  up  in  that  ocean  also. 
At  the  service's  peak,  there  were  22  Atlantic  and 
24  Pacific  stations. 

At  war's  end,  the  Navy  intended  to  discontinue 
weather  ship  operations,  but  pressure  from  sev- 
eral sources  resulted  instead  in  establishment  of  a 
permanent  peacetime  system  of  13  stations. 
These  are  shown  on  the  next  page,  with  the  posi- 
tions and  operating  nations  listed  in  the  accom- 
panying table.  Costs  of  the  program  were  shared 
by  nations  operating  transoceanic  aircraft. 

A  typical  weather  patrol  was  21  days  on-sta- 
tion.  A  "station"  was  a  210-mile  grid  of  10-mile 
squares,  each  with  alphabetic  designations.  The 
center  square,  which  the  ship  usually  occupied, 
was  "OS"  (for  "on-station").  A  radio  beacon 


Coast  Guard 
Cutter  Sebago  was 
photographed  on 
Station  A  in 
lanuary  1949. 


OCEANUS 


Ocean  Weather  Stations 
1 94O  -  1 98O 


Sta.      Position 

A  62W  W;  33~J00'  W 
56"30'N;51'00'W 
52"45'  N;  3930'  W 
WOO'  N;  4100' W 
35°00'  N;  48  00'  W 
36°00'  N;  7000'  W 
6VOO'N;15°00'W 
52*30'  N;  20*00'  W 
45=00'  W;  76  00'  W 
66*00'  N;  02*00'  E 


B 
C 
D 

E 
H 
I 
) 
K 
M 


Operator 

U.S.  &  Wet/i. 

U.S. 

U.S. 

U.S. 

U.S. 

U.S. 

U.K. 

U.K. 

France 

Norway 


PACIFIC 


Map  shows  the  13 
permanent  weather 
stations  established 
in  1946  by  the 
United  Nations 
Civil  Aviation  Orga- 
nization. Program 
costs  were  shared 
by  nations  operat- 
ing transoceanic 
aircraft.  Letters 
missing  from  the 
alphabetical  se- 
quence were  those 
used  for  stations 
occupied  during 
World  War  II  but 
not  included  in  the 
postwar  weather 
station  program. 


Weather  balloons 
were  released  from 
weather  ships  every 
six  hours  to  gather 
data  from  eleva- 
tions as  high  as 
50,000  feet. 


transmitted  the  ship's  location. 
Overflying  aircraft  would  check 
in  with  the  ship  and  receive 
position,  course  and  speed  by 
radar  tracking,  and  weather  data. 
Surface  weather  observations 
were  transmitted  every  three 
hours,  and  "upper  airs" — from 
instrumented  balloon  data — 
every  six  hours.  Using  radiosonde  transmitters  and 
radar  tracking,  balloon  observers  obtained  air 
temperature,  humidity,  pressure,  and  wind  direc- 
tion and  speed  to  elevations  of  50,000  feet. 

Oceanographic  observations  were  recom- 
mended for  weather  ships  almost  from  the  start. 
Beginning  in  1945  and  continuing  to  the  end,  US 


Sta. 

Position 

Operator 

N 

30  N;  140  W 

U.S. 

P 

50°  N;  145"  W 

Canada 

V 

34  N;  164  f 

U.S. 

ships  made  bathyther- 
mograph (B/T)  observa- 
tions that  today  consti- 
tute the  largest  B/T 
archive  in  existence. 
Many  specific,  short- 
term  programs  were 
carried  out  with  ocean- 
ographers  frequently 
riding  the  ships.  In 
addition  to  serving  as 
weather  reporters  and 
navigation  aids, 
weather  ships  occasion- 
ally rescued  downed 
aircraft  and  foundering 
ships.  Dramatic  weather 
station  rescues  include 
the  Bermuda  Sky 
Queen  in  1947  (Station 

C),  Pan-American  943  (Station  N)  in  1956,  and 
SS  Ambassador  (Station  E)  in  1964. 

By  1 970,  new  jet  aircraft  were  coming  to  rely 
less  on  fixed  ocean  stations,  and  satellites  were 
beginning  to  provide  weather  data.  In  1974,  the 
Coast  Guard  announced  plans  to  terminate  the 
US  stations,  and,  in  1977,  the  last  weather  ship 
was  replaced  by  a  newly  developed  buoy.  The 
international  program  ended  when  the  last  ship 
departed  Station  M  in  1981. 

Gipt.  Dinsmore  commanded  the  weather  ship  USCGC  Cook 
Inlet.  During  /»'»  28-year  Coast  Guard  career,  he  seri'ed  on  four 
North  Atlantic  weather  ships  and  was  weather  ship  program 
manager  before  joining  the  WHO/  Staff  in  1971.  This  article  is 
e\erpted  from  a  text  about  twice  this  length.  Interested  readers 
may  request  the  longer  account  from  the  Oceanus  office  by 
calling  508-289-3516  (email:  oceanusmag@whoi.edu). 


FALL/WINTER  1996 


A  Century 

of  North  Atlantic  Data 
Indicates  Interdecadal  Change 

Surface  Temperature,  Winds,  &  Ice  in  the  North  Atlantic 


Clara  Deser 

Research  Associate,  University  of  Colorado 

For  hundreds  of  years  mariners  have  re- 
corded the  weather  over  the  world  ocean. 
Some  100  million  marine  weather  reports 
have  accumulated  worldwide  since  1854,  when 
an  international  system  for  the  collection  of 
meteorogical  data  over  the  oceans  was  estab- 
lished. These  reports  include  measurements  of 
sea  surface  temperature,  air  temperature,  wind, 
cloudiness,  and  barometric  pressure.  In  the 
1980s,  the  National  Oceanic  and  Atmospheric 
Administration  (NOAA)  compiled  these  weather 
observations  into  a  single,  easily  accessible  digi- 
tal archive  called  the  Comprehensive  Ocean- 
Atmosphere  Data  Set.  This  important  data  set 
forms  the  basis  for  our  empirical  knowledge  of 
the  surface  climate  and  its  variability  over  the 
world's  oceans:  One  example  of  a  variable  sys- 
tem is  the  phenomenon  known  as  El  Nino  in  the 
tropical  Pacific  (see  article  on  page  39).  A  major 
challenge  in  climate  research  is  to  use  these  data 
to  document  and  understand  the  role  of  the 
oceans  in  long-term — decadal  and  centennial — 
climate  change. 

The  figure  at  right  shows  the  geographical 
distribution  of  weather  observations  over  the 
oceans  for  three  periods:  1880-1900,  1920- 
1940,  and  1960-1980.  Before  the  turn  of  the 
century,  marine  weather  reports  were  largely 
restricted  to  shipping  lanes  in  the  North  Atlantic 
and  western  South  Atlantic.  The  North  Pacific 
was  not  well  sampled  until  after  World  War  II, 
and  the  tropical  oceans  not  until  after  about 
1960;  the  southern  oceans  are  still  largely  un- 
measured. Due  to  the  irregular  sampling,  we 
focus  on  describing  climate  variations  over  the 
North  Atlantic  back  to  the  turn  of  the  century. 
Fortunately,  the  North  Atlantic  plays  an  impor- 
tant role  in  world-ocean  circulation. 

Two  parameters  are  of  key  importance  to  the 
physical  interaction  between  ocean  and  atmo- 
sphere: sea-surface  temperature  and  near-surface 
wind.  They  control  the  rates  of  heat  and  momen- 
tum transfer  between  the  two  media.  The  top 
figure  on  page  12  displays  the  long-term  average 


distributions  of  sea-surface  temperature  and 
near-surface  wind  over  the  North  Atlantic.  These 
charts  are  based  upon  all  available  observations 
since  1900.  The  prevailing  westerly  winds  or 
"westerlies"  are  a  well-known  feature  of  the  wind 
distribution.  Sea  surface  temperatures  are  gener- 
ally warmer  in  the  East  Atlantic  than  in  the  West 
Atlantic  at  the  same  latitude,  reflecting  the  mod- 
erating influence  of  the  Gulf  Stream. 

How  have  the  wind  and  temperature  distribu- 
tions changed  with  time?  A  statistical  technique 
called  empirical  orthogonal  function  analysis 
aids  in  identifying  regions  of  coherent  temporal 


Geographical  dis- 
tribution of 
weather  reports 
over  the  world. 
Colored  areas 
show  the  average 
number  of  weather 
reports  per  month 
in  each  2°  latitude 
by  2"  longitude 
square  over  the 
world  oceans  for 
each  of  the  time 
periods  indicated. 
White  areas  indi- 
cate there  are  no 
reports. 


OCEANUS 


Average  distribu- 
tions of  sea  surface 
temperature  ("C) 
(top)  and  near  sur- 
face wind  climatol- 
ogy (bottom)  over 
the  North  Atlantic 
since  1900.  The 
longest  wind  arrow 
corresponds  to  8 
meters  per  second. 


change.  The  results  of  the  statistical  analysis  point 
to  the  area  directly  south  and  east  of  Newfound- 
land as  a  site  of  pronounced  sea  surface  tempera- 
ture variability.  The  figure  below  shows  the  his- 
tory of  sea  surface  temperatures  in  this  region 
since  1900.  There  is  a  notable  tendency  for  cold 
and  warm  periods  to  be  spaced  approximately 
one  decade  apart,  as  well  as  longer-term  warming 
and  cooling  trends  that  span  several  decades. 

When  the  near-surface 
wind  field  is  analyzed 
in  a  similar  manner 
(but  independently 
from  the  sea  surface 
temperatures),  similar 
decadal-scale  oscilla- 
tions and  longer  term 
trends  are  evident.  As 
noted  in  the  box  on  the 
opposite  page,  this 
basin  scale  pattern  of 
\  variability  has  been 
|  labeled  the  North  At- 
-  lantic  Oscillation. 

What  is  the  nature 
of  these  decadal  and 
multi-decadal  fluctua- 
tions? Are  they  surface 
signatures  of  oscilla- 


tions inherent  to  the  deep  ocean  circulation?  Are 
they  global  or  confined  to  the  North  Atlantic? 
What  is  the  role  of  the  atmosphere?  There  is 
mounting  evidence  from  mathematical  models 
that  the  North  Atlantic  Ocean's  thermohaline 
(heat  and  density  driven)  circulation  may  be- 
have as  a  damped  oscillatory  system  at  decadal- 
to-multidecadal  frequencies,  with  the  atmo- 
sphere supplying  the  energy  to  maintain  the 
oscillations  against  dissipation.  In  order  to  test 
the  relevance  of  hypotheses  generated  from  the 
modeling  work,  further  description  of  the  ob- 
served climate  record  is  needed. 

A  composite  picture  of  the  decadal-scale  varia- 
tions can  be  formed  by  averaging  all  of  the  cold 
(or  warm)  periods  from  the  figure  below  left  and 
subtracting  the  long-term  mean.  The  figure  di- 
rectly below  shows  such  an  "anomaly"  compos- 
ite of  the  cold  events.  When  sea  surface  tempera- 
tures to  the  east  of  Newfoundland  are  colder 
than  normal,  the  near-surface  westerly  winds  are 
stronger  than  normal.  This  relationship  may  be 
indicative  of  positive  feedback  between  atmo- 
sphere and  ocean:  Stronger  winds  cool  the  ocean 
surface  by  enhancing  evaporation  and  heat  loss, 
while  colder  surface  temperatures  shift  the  lati- 
tude of  the  storm  track  and  prevailing  westerlies 
southward.  Thus,  the  decadal  swings  in  wind 
and  temperature  may  be  a  manifestation  of  a 
coupled  air-sea  interaction  process,  in  line  with 
recent  modeling  results.  What  determines  the 
time  scale  of  the  fluctuations,  as  well  as  their 
amplitude,  are  unsolved  issues  at  this  time. 

The  decadal  fluctuations  in  sea  surface  tem- 
perature show  an  intriguing  relation  to  the 
amount  of  sea  ice  in  the  Labrador  Sea,  as  the  top 
figure  opposite  shows.  While  information  on  sea 
ice  dates  back  only  to  1953,  it  is  evident  that 
each  of  the  decadal  swings  of  colder-than-nor- 
mal  temperatures  was  preceded  by  a  period  of 
greater-than-normal  amounts  of  sea  ice.  The 
mechanism  for  this  association  is  not  well  un- 
derstood, although  it  is  plausible  that  the  cold, 
stable  water  mass  resulting  from  melting  ice 
could  be  carried  by  ocean  currents  into  the  re- 


i  surface  temperatures  for  the  region 
ind  east  of  Newfoundland  since 
•''•p.irtures  from  normal  in  de- 

i   il  curve  is  a  low-pass  filtered 
version  I,  curve,  emphasizing  fluctua- 

tions li  i  .1  few  years. 


Composite  ot  decadal-scale  cold  events  in  the  North 
Atlantic  using  sea  surface  temperature  and  wind 
anomaly  patterns  since  1900   Blue  (red)  contours  indi- 
cate colder  (warmer)  than  normal  sea  surface  tempera- 
tures. The  longest  wind  arrow  is  1  meter  per  second. 


FALLWINTSR1996 


gion  east  of  Newfoundland.  Some  researchers 
have  hypothesized  a  complex  feedback  loop 
involving  Arctic  precipitation,  runoff,  salinity, 
and  ocean  circulation  to  explain  the  decadal- 
scale  sea  ice  variations. 

The  lack  of  understanding  of  observed,  long- 
term  climate  events  in  the  North  Atlantic  under- 
scores the  need  for  further  research,  particularly 
in  relating  the  deep  ocean  circulation  to  the 
surface  conditions.  The  work  described  by 
Michael  McCartney,  Ruth  Curry,  and  Hugo 
Bezdek  beginning  on  page  19  is  one  important 
step  in  this  direction. 

This  research  was  funded  by  a  grant  from  the  Atlantic 
Climate  Change  Program  of  the  National  Oceanic  and 
Atmospheric  Administration. 

Clara  Deser  ii'as  introduced  to  oceanography  in  l')<s  i  while  ii 
Research  .-\sststant  in  U'HO/'s  Physical  Oceanography  Depart- 
ment She  then  went  to  the  University  of  Washington  to  obtain 
a  Fh.D.  in  atmospheric  sciences,  and  hai  since  kept  her  teet 


History  of  sea  ice 
amounts  in  the  La- 
brador Sea  in  rela- 
tion to  sea  surface 
temperatures  in  the 
North  Atlantic 
since  1953. 


1954 


I960 


1970 


1980 


1990 


WIT  and  her  head  dry  at  the  Utui'ersity  of  Colorado  at  Boulder 
Currently,  she  continues  her  intellectual  pursuits  on  a  part-time 
basis  u'hile  raising  two  children  with  her  husband  lonathan. 


North  Atlantic  Oscillation 


The  top  two  panels  of  the  figure,  sea  level  pressure 
in  millibars,  show  an  example  of  regional  shifting 
climatic  patterns.  From  work  by  leff  Rogers  of 
Ohio  State  University,  they  show  the  high  (+)  and  low 
(-)  extreme  states  of  the  North  Atlantic  Oscillation 
(NAO).  The  regional  atmospheric  circulation  over  the 
North  Atlantic  is  normally  characterized  by  a  subpolar 
high  pressure  cell  centered  near  the  Azores,  and  a  subpo- 
lar low  pressure  cell  cen- 
tered near  Iceland  and 
Greenland.  Between  these 
two  centers  the  westerlies 
blow  from  North  America 
towards  Europe,  while  to 
the  north  of  the  Icelandic 
low,  and  to  the  south  of 
the  Azorian  high  the  winds 
are  easterlies.  A  characteris- 
tic oscillation  of  the 
strengths  and  positions  of 
these  pressure  centers 
occurs  interannually  and 
interdecadally.  In  the  high 
NAO  state,  the  westerlies 
are  intense,  and  the  cold 
continental  air  they  carry  off  northern  North  America  is 
warmed  by  heat  liberated  from  the  warm  ocean  waters 
they  blow  across,  and  that  warmed  air  flows  across  north- 
ern Europe  from  the  southwest.  When  the  NAO  is  in  its 
low  state,  the  Icelandic  low  pressure  center  is  displaced 
far  to  the  south,  off  Newfoundland,  and  there  is  a  high 
pressure  center  over  northern  Greenland,  causing  cold 


dry  polar  air  to  blow  across  northern  Europe,  and  then 
westward  across  the  northern  subpolar  area  towards 
North  America,  warming  on  the  way  by  the  heat  liberated 
from  the  ocean  to  the  overlying  atmosphere.  In  this  low 
NAO  state,  northern  Europe  experiences  much  cooler 
summers  and  more  severe  winters  than  in  the  high  NAO 
state,  while  Labrador  experiences  a  milder  climate.  The 
bottom  two  panels  of  the  figure  show  that  the  winter 

storm  frequency  patterns 
for  the  two  extreme  states 
of  the  NAO  are  quite  dif- 
ferent, with  the  northeast- 
ern US  experiencing  more 
Nor'easters  during  the  low 
NAO  state  than  during  the 
high. 

The  differing  winds  and 
^  the  accompanying  warmer 
=  or  cooler  periods  for 
|  northern  Europe  and 
i  northern  North  America 

'•-  that  occur  when  the  NAO 

& 

"  index  marches  from  one 
I  extreme  to  the  other  over 
~  periods  of  a  decade  or 
more  contribute  significantly  to  the  distribution  of  global 
temperature  change.  Comparison  of  the  NAO  with  a 
similar  climatic  index  known  as  the  "Pacific-North 
American"  (PNA)  index  indicates  that  on  decadal  time 
scales  there  may  be  coordinated  variations  throughout 
the  northern  hemisphere  or  even  the  whole  globe. 

— Mike  McCartne}' 


Unfortunately, 
Worthington's  ef- 
forts (photo  at 
right)  were  devoted 
to  a  period  of 
minimum  produc- 
tion of  this  water. 
In  contrast  to  this 
minimum  period 
in  1976  (denoted 
by  red  curves  of 
temperature  and 
potential  density), 
a  period  of  maxi- 
mum climatologi- 
cal  production  oc- 
curred in  1964 
(blue  curves).  In 
both  cases,  the  plot 
shows  the  annually 
averaged  properties 
for  both  calendar 
years  vs.  pressure 
to  reduce  eddy 
noise.  Note  that 
the  underlying 
thermocline  at 
pressures  of  more 
than  600  decibars 
is  similar  in  both 
periods:  Changes 
are  not  induced 
from  below. 


The  Bermuda  Station  S- 
A  Long-Running  Oceanographic  Show 

Deeper  Waters  Show  Warming  Trend 


Terrence  M.  Joyce 

Senior  Scientist,  Physical  Oceanography  Department 

Lynne  Talley 

Professor  &  Research  Oceanographer, 
Scripps  Institution  of  Oceanography 

A  time  series  of  hydrographic  measure 
merits  was  initiated  at  Bermuda  in  1954 
and  continues  to  the  present.  It  began 
under  the  banner  of  the  International  Geophysical 
Year  (1957-1958)  with  the  scientific  support  of 
Henry  Stommel  of  the  Woods  Hole  Oceanographic 
Institution  and  William  Sutcliffe,  director  of  the 
Bermuda  Biological 
Station  (BBS).  The  scien- 
tists and  personnel  of 
the  originating  institu- 
tions have  been  the 
most  active  participants 
over  the  years,  but  the 
data  have  been  widely 
used  by  the  interna- 
tional Oceanographic 
community.  While  other 
long  time  series  of  mea- 
surements in  the  North 
Atlantic  began  in  asso- 
ciation with  weather 
ships,  (see  "Alpha, 
Bravo,  Charlie"  on  page 
9)  only  the  Bermuda 
measurements  have  a  strong  Oceanographic  focus. 

In  recent  years,  large  international  programs 
including  the  World  Ocean  Circulation  Experi- 
ment (WOCE)  and  the  loint  Global  Ocean  Flux 
Study  (IGOFS)  have  pro- 
vided scientific  justification 
for  continuation  and  expan- 
sion of  the  multidisciplinary 
Bermuda  measurements.  As 
these  programs  begin  to 
wind  down,  it  is  important 
to  recognize  the  significance 
of  this  time  series  study  to 
understanding  of  climatic 
change:  These  measurements 
are  the  principal  source  of 
information  about  subsur- 
face changes  in  the  Sargasso 
Sea  and  the  subtropical  gyre 
over  the  past  four  decades. 
In  early  years,  the  time  series 


During  the  cold  winter  of  1977,  Val  Worthington  ventured 
out  aboard  the  National  Oceanic  and  Atmospheric  Admin- 
istration's Researcher  to  study  the  formation  of  18°  Water, 
one  of  the  principal  North  Atlantic  water  masses,  in  the 
northern  Sargasso  Sea.  In  the  photo,  Worthington  is  leaving 
the  "hero  platform"  after  launching  a  Nansen  bottle  cast. 


was  denoted  by  the  attribution  "Panulirus,"  the 
name  of  the  small  BBS  research  vessel  used  to 
carry  out  the  sampling.  However,  over  the  years, 
several  other  research  vessels  have  serviced  the  site, 
and  the  present  practice  is  to  call  the  time  series 
Station  S,  in  keeping  with  the  convention  for 
many  weather  ship  sites. 

In  the  field  of  physical  oceanography,  Station  S 
is  not  optimally  located:  There  are  no  major  water 
masses  formed  there,  and  the  circulation  of  the 
subtropical  gyre  and  deep  boundary  currents  only 
peripherally  affect  the  island.  However,  its  location 
in  proximity  to  these 
major  North  Atlantic 
circulation  features 
make  it  ideal  for  deter- 
mining larger-scale 
changes  in  the  basin  as 
they  pass  by.  To  make  an 
analogy  with  the  study 
of  automobiles,  a  re- 
searcher might  visit 
various  factories  to  study 
manufacturing  and 
design  changes  or  just  sit 
by  a  busy  highway  and 
observe  the  passing 
traffic — Station  S  fits  the 
latter  category  very  well. 

In  the  short  space 

available,  we  wish  to  discuss  some  of  the  changes 
observed  at  Bermuda,  what  they  import  for  the 
North  Atlantic,  and  possible  reasons  for  their 
occurrence.  We  will  only  look  at  two  "layers"  in 
the  water  column,  the  Eighteen-Degree  Water  and 
the  North  Atlantic  Deep  Water. 

Eighteen-Degree  Water  was  first  described  by 
WFIOI  physical  oceanographer  Valentine 
Worthington  in  1959  as  one  of  the  major  water 
masses  formed  in  the  northern  Sargasso  Sea  in  late 
winter:  It  is  the  principal  type  of  subtropical  water 
found  in  the  North  Atlantic.  It  occurs  at  depths  of 
a  few  hundred  meters  at  Station  S  and  is  character- 
ized by  a  layer  of  nearly  constant  density  having  a 
temperature  of  about  1 8°  C.  While  this  layer  does 
not  form  at  the  surface  near  Bermuda,  it  occurs  just 
below  the  depths  of  late  winter  mixed  layers  near 
the  island  and  is  closely  coupled  to  surface  layers 
found  there.  The  Station  S  time  series  has  been  one 
of  the  main  barometers  (or,  more  correctly,  f/icr- 
mometers)  of  changes  in  this  water  mass  as  it  flows 


FALL/WINTER  1996 


hy  the  island  from  source  regions  to  the  northeast. 

The  thickness  of  Eighteen-Degree  Water  varies 
by  a  factor  of  two  over  the  course  of  the  current  42- 
year  time  series.  Temperature  and  salinity  (density) 
changes  also  occur  over  time,  and  the  layer  tem- 
perature is  only  approximately  equal  to  18°C.  In 
years  when  this  Eighteen-Degree  Water  is  produced 
in  large  quantities,  the  surface  salinity  (but  not 
necessarily  temperature)  at  Station  S  is  high  In 
poor  production  years,  the  surface  salinity  is  low. 
Thus,  long-term  changes  in  this  water  mass  seem 
to  be  closely  linked  with  processes  that  affect  the 
surface  salinity  High  production  periods  seem  to 
be  recur  at  intervals  of  approximately  12  to  14 
years.  Many  of  us  recall  when  Worthington  con- 
vinced the  funding  agencies  to  mount  a  field  study 
of  Eighteen-Degree  Water,  only  to  find  that  none 
had  formed  that  year.  We  can  now  see  that  the 
climatological  minimum  of  the  signal  at  Bermuda 
occurred  during  the  mid  1970s  when  he  went  to 
sea!  Our  study  of  the  processes  controlling  this 
variability  has  not  provided  a  conclusive  answer  as 
to  why  this  periodicity  occurs  and  how  it  is  linked 
to  surface  salinity — but  not  temperature — changes, 
though  we  believe  it  must  have  some  connection 
with  changes  in  atmospheric  circulation  and  pre- 
cipitation over  the  subtropical  Atlantic  Ocean. 

At  depths  of  1,500  to  2,500  meters  at  Station  S, 
we  find  another  clear  signal  that  is  not  connected 
with  atmospheric  changes  over  the  subtropical 
gyre.  This  is  a  slow  increase  in  temperature  over 
time  with  a  trend  that  is  apparent  in  records  that 
date  to  the  early  1920s  when  deep-water  oceano- 
graphic  measurements  were  first  made  near  Ber- 
muda with  accurate  reversing  thermometers.  The 
long-term  trend  of  this  temperature  change  is  at  a 
rate  of  approximately  0.5°  C 
per  century.  It  is  one  of  the 
clearest  examples  of  a  long- 
term  increase  of  ocean  tem- 
perature in  the  subsurface 
ocean.  That  is  not  to  say  that 
the  annually  averaged  tem- 
perature in  this  layer  always 
increases.  In  fact  there  are 
decadal  time-scale  changes  at 
this  depth  too,  with  1993 
appearing  to  be  the  coldest  in 
20  years.  Since  waters  at  this 
depth  do  not  communicate 
with  the  surface  anywhere  in 
the  subtropical  gyre,  it  is  the 
subpolar  gyre  to  the  north 
that  is  the  most  likely  cause  for  the  variability,  if 
not  the  trend.  Our  present  studies  indicate  that 
long-term  changes  in  the  production  of  Labrador 
Sea  Water  are  associated  with  the  decadal  variabil- 
ity in  the  deep  water  at  Station  S.  Since  the  Station 
S  bottom  is  at  about  3,000  meters,  this  deep  layer 


800 

1000 
1200 
1400 

1600 
u 

1800 
2000 
2200 
2400 


10 


Temperature  :C 


65  OO'W  64  45'W  64  30'W  64  15'W 

The  location  of  Station  S  is  a  short  steam 
southeast  from  St.  Georges'  harbor.  The 
smoothed  bathymetry  is  plotted  at  one  kilo- 
meter depth  increments. 


is  the  deepest  avail- 
able in  the  time  se- 
ries. There  appears  to 
be  a  lag  of  5  to  6 
years  until  the  Labra- 
dor Sea  Water  signal 
appears  at  Bermuda 
(see  "Labrador  Sea" 
article  on  page  24). 
Many  of  the  cli- 
mate studies  that  can 
be  undertaken  using 
the  time  series  at 
Bermuda  are  ongoing 
and  of  greater  value 
as  time  goes  on.  This  is  related  simply  to  the  fact 
that  phenomena  with  time  scales  of  a  decade  must 
be  studied  with  time  series  that  span  several 
"events"  in  order  to  make  any  statistically  signifi- 
cant statements.  The  examples  given  above  are 
marginal  in  the  statistical  sense  because  even  the 
42-year  data  set  is  not  long  enough — we  are  al- 
ways left  looking  at  the  most  recent  years  of  data 
and  wondering  what  will  come  next!  For  this 
reason  it  is  essential  that  the  observations  con- 
tinue beyond  the  lives  of  some  of  the  large  field 
programs  like  WOCE  and  JGOFS,  which  will  end 
in  a  few  years.  Though  it  is  possible  that  techno- 
logical advances  may  enable  additional  measure- 
ments or  more  cost-effective  methods,  the  Ber- 
muda time  series  currently  offers  our  principal 
window  into  the  climate  of  the  subtropical  gyre  of 
the  North  Atlantic. 

The  National  Science  Foundation  has  funded  the  Station  S 
work  over  the  years. 

Terr}'  loyce  was  a  member  of  the  first  class  aiimitied  to  the  MIT/ 
WHO/  loin!  Program  in  Physical 
Oceanography  1968.  His  first  cniise, 
with  Henry  Stommcl  aboard 
Atlantis  II  that  same  year,  made  a 
port  cat!  in  Bermuda,  so  his  Ber- 
muda connection  extends  far  beyond 
this  article.  Since  completing  his 
doctorate  in  1972,  he  has  conducted 
research  at  WHO/  except  tor  ,i  six- 
month  stint  in  Germany.  Though  his 
research  interests  have  changed  over 
the  years,  there  has  always  been  a 
strong  thread  ot  seagoing  work  and 
data  analysis/interpretation.  For  the 
past  eight  years,  he  has  senvd  as 
director  of  the  World  Ocean  Circula- 
tion Experiment  Hydrographic 
Program  Office  based  at  WHO/. 

Lynne  Taliey  was  also  ii  WHOI/MIT 
joint  Program  student,  having  begun 


her  graduate  career  as  a  large-scale  observational  oceanograplier 
working  on  items  like  Eighteen-Degree  Water  using  the  Bermuda 
record  and  Labrador  Sea  Water.  She  moved  into  theoretical 
studies  o/  unstable  currents  for  her  degree,  completed  in  1 982,  but 
several  years  after  moving  on  through  a  postdoc  and  beginnings  of 
her  research  career  she  found  herself  squarely  back  in  the  interme- 
diate and  mode  waters  of  the  world,  both  literally  and  in  print. 


K  35.2         35.4 

Salinity  %. 

Deep  water 
changes  at  Ber- 
muda are  illus- 
trated by  two  con- 
trasting years:  1959 
(red  curves)  and 
1987  (blue  curves). 
During  the  inter- 
vening period,  the 
deep  water  warmed 
up  and  became 
saltier  between 
about  1,200 
decibars  (or  ap- 
proximately 1,200 
meters)  and  the 
bottom.  At  shal- 
lower pressures, 
eddy  variability 
(denoted  by  the 
dotted  lines  on  ei- 
ther side  of  the  an- 
nual means)  ob- 
scures any 
differences.  The 
long-term  warming 
trend  at  Bermuda 
can  be  traced  back 
to  1922,  when  ac- 
curate deep  water 
temperature  mea- 
surements were 
made  by  the  Dan- 
ish ship  Diimi  //. 


OCEANUS 


Average  summer- 
time temperature 
over  six  centuries  in 
the  northern  hemi- 
sphere. Note  gener- 
ally cooler  tem- 
peratures between 
1550  and  1900,  the 
period  known  as 
the  Little  Ice  Age. 
(Data  courtesy  of 
Raymond  S.  Brad- 
ley, University  of 
Massachusetts). 


Sedimentary  Record  Yields 
Several  Centuries  of  Data 

Tlie  Little  Ice  Age  and  Medieval  Warm  Period  in  the  Sargasso  Sea 


-0.1 


U1 

e_ 

t 
o 


Lloyd  D.  Keigwin 

Senior  Scientist,  Geology  &  Geophysics  Department 

New  Englanders  claim  a  birthright  to 
complain  about  the  weather.  As  we  note 
that  the  summer  of  1996  was  coolest 
and  wettest  in  recent  memory,  most  of  us  have 
already  forgotten  that  summer  1995  was  unusu- 
ally warm  and  dry.  Such  variability  in  weather  is 
normal,  yet  in  historical  times  there  have  been 
truly  exceptional  events.  For  example,  1816  is 
known  as  the  "Year  Without  a  Summer."*  During 
that  year,  there  were  killing  frosts  all  over  New 
England  in  May,  lune,  and  August,  (uly  1816  was 
the  coldest  July  in  American  history,  and  frosts 
came  again  in  September.  Crop  failure  led  to  food 
shortages  through- 
out the  region. 
Although  the  im- 
mediate cause  of 
cooling  has  been 
ascribed  to  the 
volcanic  eruption 
of  Tambora  in 
Indonesia  the  year 
before,  the  Year 
Without  a  Summer 
occurred  during  a 
time  when  weather 
was  generally  more  harsh  than  today.  Persistently 
harsher  weather  suggests  a  change  in  climate,  and 
the  late  16th  through  the  19th  centuries  have 
become  known  as  the  "Little  Ice  Age." 

The  Little  Ice  Age,  and  several  preceding  cen- 
turies, which  are  often  called  the  "Medieval 
Warm  Period,"  are  the  subject  of  controversy. 
Neither  epoch  is  recognized  at  all  locations 
around  the  globe,  and  indeed  at  some  locations 
there  is  clear  evidence  of  warming  while  others 
show  distinct  cooling.  One  author  titled  a  paper: 
"Was  there  a  Medieval  Warm  Period,  and  if  so 
when  and  where?"  Nevertheless,  when  data  from 
all  Northern  Hemisphere  locations  are  consid- 
ered, the  annual  average  summer  temperature 
proves  to  be  a  few  tenths  of  a  degree  lower  dur- 
ing  the  coldest  part  of  the  Little  Ice  Age  in  the 
late  1500s  and  early  1600s.  Various  forcing 


-0.3 


1-0.4 


-0.5 


1800's 


100 


200 
Years 


mechanisms  have  been  proposed  for  such 
changes,  including  variation  in  the  sun's  energy 
output,  volcanic  eruptions,  and  mysterious  inter- 
nal oscillations  in  Earth's  climate  system,  but 
none  satisfy  all  of  the  data. 

Natural  climate  changes  like  the  Little  Ice  Age 
and  the  Medieval  Warm  Period  are  of  interest  for 
a  few  reasons.  First,  they  occur  on  decade  to 
century  time  scales,  a  gray  zone  in  the  spectrum 
of  climate  change.  Accurate  instrumental  data  do 
not  extend  back  far  enough  to  document  the 
beginning  of  these  events,  and  historical  data  are 
often  of  questionable  accuracy  and  are  not  wide- 
spread geographically.  Geological  data  clearly 
document  globally  coherent  climate  change  on 

thousand-,  ten  thou- 
sand-, and  hundred 
thousand-year  time 
scales,  so  why  is  the 
record  so  confusing 
over  just  the  past 
1,000  years?  Second, 
as  humanity  contin- 
l  ues  to  expand  and 
-  make  more  de- 
'-  mands  on  our 
planet,  annual  aver- 
age temperature 
changes  of  a  degree  could  have  considerable 
social  and  economic  impacts.  Third,  as  there  is 
widespread  agreement  among  climatologists  that 
changes  due  to  human  impacts  on  atmospheric 
chemistry  will  eventually  lead  to  global  warming 
of  about  two  degrees  over  the  next  century,  it  is 
important  to  understand  the  natural  variability 
in  climate  on  the  century  time  scale.  Will  the 
human  effects  occur  during  a  time  of  natural 
warming  or  cooling? 

Of  several  approaches  to  studying  climate  on 
decadal  to  century  time  scales,  here  I  will  touch 
on  the  study  of  long  series  of  measurements 
made  at  sea  and  the  study  of  deep  sea  sediments. 

•This  phenomenon  is  described  in  Volcano  Weather- 
The  Story  of  1816,  the  Year  Without  a  Summer  by  Henry 
Stommel  and  Elizabeth  Stommel  (Seven  Seas  Press, 
Newport,  RI,  1983) 


1600's 


300  400 

Before  Present 


500 


600 


FALL/WINTER  1996 


Ordinarily,  there  is  little  overlap  between  these 
two  approaches.  Reliable  and  continuous  hydro- 
graphic  observations  rarely  extend  back  beyond 
several  decades,  and  deep  sea  sediments  usually 
accumulate  too  slowly  to  resolve  brief  climate 
changes.  I  lowever,  the  northern  Sargasso  Sea  is  a 
region  where  we  have  five  decades  of  nearly  con- 
tinuous biweekly  hydrographic  data  (see  preced- 
ing article),  a  long  history  of  sediment  trap  col- 
lections to  document  the  rain  of  particles  from 
the  sea  surface  to  the  seafloor,  and  exceptional 
deep  sea  cores  of  sediment.  The  co-occurrence  of 
these  three  elements  has  led  to  one  of  the  first 
reconstructions  of  sea  surface  temperature  for 
recent  centuries  in  the  open  ocean. 

Oceanographically,  Station  S  in  the  western 
Sargasso  Sea  is  important  because  tempera- 
ture and  salinity  change  there  is  typical 
of  a  large  part  of  the  western  North 
Atlantic,  and  it  is  exclusively  western 
North  Atlantic  water  that  is  trans- 
ported northward  and  eastward  by 
the  Gulf  Stream.  These  are  the  wa- 
ters that  eventually  cool  and  sink 
in  the  Norwegian  and  Greenland 
Seas,  flowing  southward  to  com- 
plete a  large-scale  convection  cell 
that  plays  a  fundamental  role  in 
regulating  Earth's  climate. 

In  addition  to  long  time  series  of  hydro- 
graphic  data  from  Station  S,  the  site  is  remark- 
able for  the  long  series  of  sediment  trap  data 
collected  by  WHOI's  Werner  Deuser,  beginning 
in  the  1970s.  Those  traps  have  recovered  nearly 
continuous  samples  of  the  seasonally  changing 
rain  of  particles  that  settle  from  surface  waters  to 
the  seafloor.  An  important  component  of  those 
particles  is  the  calcium  carbonate  shells  of 
planktonic  protozoans  known  as  foraminifera. 
There  are  about  30  species  of  foraminifera,  or 
"forams,"  and  Deuser's  investigations  have  es- 
tablished the  seasonal  change  in  species  abun- 
dance and  their  stable  isotope  composition.  We 
now  know  from  these  studies  that  only  one 
species  of  planktonic  foram,  Globigerinoides 
ruber,  lives  year-round  at  the  surface  of  the  Sar- 
gasso Sea,  and  it  happens  to  deposit  its  calcium 
carbonate  close  to  oxygen  isotopic  equilibrium 
with  seawater.  This  means  that  G.  ruber  is  ideal 
for  reconstructing  past  changes  in  the  tempera- 
ture and  salinity  of  Sargasso  Sea  surface  waters, 
as  the  figure  at  right  illustrates. 

Note  that  the  average  sea  surface  temperature 
and  salinity  from  near  Bermuda  display  some 
systematic  variability  on  an  annual  average  basis 
since  1955.  These  changes  reflect  a  decade-long 
variability  in  the  North  Atlantic  climate  regime 
that  is  known  as  the  North  Atlantic  Oscillation 
(see  box  on  page  13).  In  this  time  series,  the  most 


severe  climate  occurred 
in  the  1960s  when 
annual  average  sea 
surface  temperatures 
were  depressed  about 
half  a  degree  by  extreme 
storminess  in  the  west- 
ern North  Atlantic. 
Cold,  dry  winds  during 
winter  storms  also 
probably  raised  surface 
ocean  salinity  in  the 
1 960s  by  promoting 
increased  evaporation. 
If  we  had  "annual  aver- 
age forams"  from  the 

1960s,  their  oxygen  isotope  ratio  would  look 
like  the  time  series  shown  in  purple.  The 
biggest  climate  change  of  the  past  five 
decades  could  indeed  be  recorded  by 
the  forams. 

Long  before  I  knew  that  G.  ruber 
was  the  best  possible  foram  for 
reconstructing  sea  surface  tempera- 
tures, I  selected  that  species  for  my 
stable  isotope  studies  because  of  its 
consistent  abundance  on  the  Ber- 
muda Rise,  in  the  northern  Sargasso 
Sea  to  the  east  of  Station  S.  At  the  time 
(the  early  1980s),  the  Bermuda  Rise  was  under 
consideration  as  a  possible  site  for  burial  of  low 
level  nuclear  waste,  and  it  was  necessary  to  know 
how  rapidly  and  continuously  the  sediment  accu- 
mulates. It  turns  out  that  because  of  the  action  of 
deep  ocean  currents,  fine-grained  clay  and  silt 
particles  are  selectively  deposited  there,  resulting 
in  very  high  rates  of  sedimentation.  And  whether 
samples  are  of  modern  or  glacial  age,  G.  ruber  is 
consistently  present.  Much  of  my  work  over  the 
past  decade  has  documented  the  climate  changes 
that  occur  on  thousand  year  time  scales  and  are 
preserved  in  foram  isotope  ratios  and  other  data 
from  Bermuda  Rise  sediments. 

Until  recently,  the  available  data  from  the 
Bermuda  Rise  showed  evidence  of  century-  to 
thousand-year  climate  change  continuing  right 
up  to  about  a  thousand  years  ago,  the  age  of  the 


36.25 


36.50- 
36.75- 

£•  37.00- 

c 

5  37.25- 
37.50 

38.75- 


Salinity 


1950 


1 

1960 


--0.75 


—  -Q 

--0.50  £  5 

h 


Shells  of  plank- 
tonic animals 
called  formainifera 
record  climatic 
conditions  as  they 
are  formed.  This 
one,  Globigerinoides 
niber,  lives  year- 
round  at  the  sur- 
face of  the  Sar- 
gasso Sea.  The 
form  of  the  live 
animal  is  shown 
above,  and  its 
shell,  which  is  ac- 
tually about  the 
size  of  a  fine  grain 
of  sand,  at  left. 


1970  1980  1990 

Year 


2000 


Bermuda  Station  S 
hydrography 
shows  the  oxygen 
isotope  ratio  that  a 
foram  would  have 
if  it  deposited  its 
shell  in  equilib- 
rium with  the  an- 
nual average  sea 
surface  tempera- 
ture and  salinity 
observed  since 
1954  at  Stations 
near  Bermuda.  The 
large  decrease  in 
sea  surface  tem- 
perature and  in- 
crease in  salinity  in 
the  late  1960s  was 
caused  by  unusu- 
ally unpleasant 
weather  those 
years.  (Tempera- 
ture and  salinity 
data  provided  by 
Terry  loyce.) 


OCEANUS 


Since  1978,  Scien- 
tist Emeritus 
Werner  Deuser  has 
collected  a  nearly 
continuous  suite  of 
deep  sediment  trap 
samples  at  the 
Ocean  Flux  Pro- 
gram site  near  Sta- 
tion S.  The  Ocean 
Flux  Program  traps 
are  shown  follow- 
ing recovery 
aboard  the  Ber- 
muda Biological 
Station  vessel 
Weatherbird  11.  The 
traps  were  de- 
ployed along  a  bot- 
tom tethered 
mooring  at  500, 
1,500  and  3,200 
meters  depths  to 
intercept  particles 
sinking  through 
the  water  column. 
Deuser  recently 
passed  the  leader- 
ship of  the  Ber- 
muda time-series 
program  on  to  As- 
sistant Scientist 
Maureen  Conte. 


sediment  at  the  tops  of  our  cores.  Because  these 
samples  were  recovered  with  large,  heavy  tools 
that  free  fall  into  the  seafloor,  I  suspected  that 
they  might  have  pushed  away  sediments  of  the 
last  millennium  without  actually  coring  them.  As 
a  test  of  this  idea,  we  acquired  a  box  core  from 
the  Bermuda  Rise  (box  cores  penetrate  the  sea- 
floor  slowly  and  disturb  surface  sediments  little) 
and  radiocarbon  dated  its  surface  sediment  at  the 
National  Ocean  Sciences  Accelerator  Mass  Spec- 
trometry  Facility  located  at  WHOI.  Results 
showed  that  the  sediment  was  modern,  and  addi- 
tional dates  were  used  to  construct  a  detailed 
chronology  of  the  past  few  millennia.  When  tem- 
peratures were  calculated  from  oxygen  isotope 
results  on  G.  ruber  from  the  box  core,  and  when 
data  were  averaged  over  50  year  intervals,  I  found 
a  consistent  pattern  of  sea  surface  temperature 
change  (see  figure  below).  The  core-top  data  indi- 
cate temperatures  of  nearly  23  degrees,  very  close 
to  the  average  temperature  at  Station  S  over  the 
past  50  years.  However,  during  the  Little  Ice  Age 
of  about  300  years  ago  sea  surface  temperatures 
were  at  least  a  full  degree  lower  than  today,  and 
there  was  an  earlier  cool  event  26 

centered  on  1,700  years  ago.  Events 
warmer  than  today  occurred  about 
500  and  1,000  years  ago,  during 
the  Medieval  Warm  Period,  and  it 
was  even  warmer  than  that  prior  to 
about  2,500  years  ago. 

These  results  are  exciting  for  a 
few  reasons.  First,  events  as  young 
and  as  brief  as  the  Little  Ice  Age 
and  the  Medieval  Warm  Period 
have  never  before  been  resolved  in 
deep  sea  sediments  from  the  open 
ocean.  Because  the  Sargasso  Sea 
has  a  rather  uniform  temperature 
and  salinity  distribution  near  the 


25- 


24- 


23- 


22- 


21 


surface,  it  seems  that  these  events  must  have  had 
widespread  climatic  significance.  The  Sargasso 
Sea  data  indicate  that  the  Medieval  Warm  Period 
may  have  actually  been  two  events  separated  by 
500  years,  perhaps  explaining  why  its  timing  and 
extent  have  been  so  controversial.  Second,  it  is 
evident  that  the  climate  system  has  been  warm- 
ing for  a  few  hundred  years,  and  that  it  warmed 
even  more  from  1,700  years  ago  to  1,000  years 
ago.  There  is  considerable  discussion  in  the  scien- 
tific literature  and  the  popular  press  about  the 
cause  of  warming  during  the  present  century. 
Warming  of  about  half  a  degree  this  century  has 
been  attributed  to  the  human-induced  "green- 
house effect."  Although  this  is  not  universally 
accepted,  it  is  widely  accepted  that  eventually 
changes  to  Earth's  atmosphere  will  cause  climate 
warming.  The  message  from  the  Bermuda  Rise  is 
that  human-induced  warming  may  be  occurring 
at  the  same  time  as  natural  warming — not  an 
ideal  situation.  Finally,  building  on  the  studies  of 
physical  oceanographers  and  climatologists, 
marine  geologists  and  paleoclimatologists  may 
use  the  North  Atlantic  Oscillation  as  a  model  for 
understanding  North  Atlantic  climate  change  on 
longer,  century  and  millennial  time  scales. 

This  work  was  funded  by  the  National  Oceanic  &  Atmo- 
spheric Administration's  Atlantic  Climate  Change  Program. 

We  encourage  Oceanus  authors  to  include  a  bit  of  humor  in  the 
short  biographies  we  request.  Lloyd  Keigwin  claimed  to  be"a 
humorless  scientist"  who  doesn't  like  writing  bios,  so  we  asked 
Eben  Franks,  a  research  assistant  in  Lloyd's  lab,  to  provide  some 
information.  Here's  what  Eben  wrote:  In  addition  to  running  a 
demanding  research  program,  Lloyd  Keigu'in  is  also  a  Com- 
mander in  the  Nav}'  Reserve.  Despite  nearly  30  years  of  sea- 
going experience  he  still  finds  himself  subject  to  seasickness. 
/Editor's  note:  This  is  not  unusual  among  oceanographers!] 
Lloyd  has  been  deeply  affected  by  episodes  of  the  popular  PBS 
series  "This  Old  House"  and  has  spent  14  years  (and  counting) 
demolishing  two  perfectly  adijuate  houses  in  the  name  of  reno- 
vation. His  limited  spare  time  is  consumed  with  multifarious 
projects  ranging  from  attempting  to  convince  the  Naiy  to 
convert  a  nuclear  sub  for  oceanographic  research  to  casting 
longing  looks  at  the  antique  German  and  British  spans  cars 
collecting  dust  in  his  burn. 


Medieval 
Warm  Period 


I | I I I I | I I I I | I I I FT    r 
500  1000  1500  2000 

Years  Before  Present 


2500 


3000 


Estimated  sea  surface  temperature  from  Station  S  annual  averages 
and  from  Globigerinoides  ruber  shell  oxygen  isotopes  averaged  at 
50-year  intervals.  Note  that  the  range  of  sea  surface  temperature 
variability  on  longer  time  scales  is  much  larger  than  what  has 
been  observed  since  1954  at  Station  S. 


FALL/WINTER  1996 


North  Atlantic's 

Transformation  Pipeline  Chills 

and  Redistributes  Subtropical  Water 

But  It's  Not  A  Smooth  Process  And  It  Mightily  Affects  Climate 


Michael  S.  McCartney 

Senior  Scientist,  Physical  Oceanography  Department 

Ruth  G.  Curry 

Research  Associate,  Physical  Oceanography  Department 

Hugo  F.  Bezdek 

Director,  Atlantic  Oceanographic  and  Meteorological 

Laboratory,  National  Oceanic  and  Atmospheric 

Administration 

Warm  and  salty  waters  from  the  upper 
part  of  the  South  Atlantic  flow  north- 
ward across  the  equator  and  then 
progress  through  the  tropical  and  subtropical 
North  Atlantic  to  reach  high  latitudes.  Beginning 
with  the  intense  northward  flow  of  the  Gulf 
Stream  off  the  East  Coast  of  the  United  States, 
these  waters  are  exposed  to  vigorous  cooling, 


liberating  considerable  oceanic  heat  to  the  atmo- 
sphere. This  is  the  first  stage  of  "warm  water  trans- 
formation" within  the  North  Atlantic,  a  process 
that  culminates  in  the  high  latitude  production  of 
cold  and  fresh  waters  that  return  to  the  South 
Atlantic  in  deep  reaching  currents  beneath  the 
warm  waters  of  the  subtropics  and  tropics. 

This  article  focuses  on  the  part  of  this  warm 
water  transformation  that  occurs  northwards  of 
about  45°  N,  the  subpolar  realm  of  the  North 
Atlantic.  Here  the  warm  waters  brought  to  the 
area  by  the  Gulf  Stream  flow  eastward  across  the 
basin  and  then  sweep  northwards  in  the  eastern 
Atlantic,  continuing  to  cool,  and  freshening  as 
precipitation  and  continental  runoff  exceed 
evaporation.  This  transformation  occurs  along 


The  pathways  asso- 
ciated with  the 
transformation  of 
warm  subtropical 
waters  into  colder 
subpolar  and  polar 
waters  in  the 
northern  North  At- 
lantic. Along  the 
subpolar  gyre 
pathway  the  red  to 
yellow  transition 
indicates  the  cool- 
ing to  Labrador 
Sea  Water,  which 
flows  back  to  the 
subtropical  gyre  in 
the  west  as  an  in- 
termediate depth 
current  (yellow). 
In  the  Norwegian 
and  Greenland 
Seas  the  red  to 
blue/purple  transi- 
tions indicate  the 
transformation  to 
a  variety  of  colder 
waters  that  spill 
southwards  across 
the  shallow  ridge 
system  connecting 
northern  Europe, 
Iceland,  Green- 
land, and  northern 
North  America. 
These  overflows 
form  up  into  a 
deep  current  also 
flowing  back  to 
the  subtropics 
(purple),  but  be- 
neath the  Labrador 
Sea  Water.  The 
green  pathway  also 
indicates  cold  wa- 
ters— but  so  influ- 
enced by  continen- 
tal runnoff  as  to 
remain  light  and 
near  the  sea  sur- 
face on  the  conti- 
nental shelf. 


OCEANUS 


1 


Temperature  ( "C) 
of  the  deep  mixed 
layer  near  the  end 
of  winter  in  the  ar- 
eas where  that 
depth  exceeds  200 
meters.  It  is  based 
on  hydrographic 
survey  data  re- 
corded from  1957 
to  1967. 


Examples  of  the 
variation  of  tem- 
perature with 
depth  from  the 
stations  used  for 
the  figure  above. 


two  distinct  pathways.  The  Norwegian  Current 
carries  part  of  the  warm  water  flowing  northward 
past  Ireland  into  the  Norwegian  and  Greenland 
(Nordic)  Seas,  while  the  subpolar  gyre  carries  the 
rest  westwards  towards  and  past  Greenland  to  the 
Labrador  Basin. 

The  transformation  from  warm  to  colder  water 
is  a  multi-year  process:  Wintertime  winds  cool 
the  surface  waters,  causing  them  to  convect  or 
vertically  overturn  and  mix  progressively  more 
deeply  into  the  cooler  waters  beneath.  This  sea- 
sonal overturning  creates  large  volumes  of  verti- 
cally homogenized  water,  called  mode  waters.  In 
summer,  the  sun  heats  the  surface  waters,  form- 
ing a  cap  of  warmer  water  that  effectively  isolates 
the  mode  water  from  contact  with  the  atmo- 
sphere. Surface  cooling  in  the  following  winter 
removes  the  cap,  and  reexposes  the  mode  water, 
which  then  undergoes  another  round  of  winter 

Temperature  (°C) 

2     3     4      5     6     7     8     9    10   1 1    1 2    1 3    14    15 


Labrador 
Basin  < 


500 


D 
m 
•a 


1000 


3 
ro 


1500 


2000 


cooling  in  which  it  gains  more  thickness.  The 
mode  water  thus  cools  and  thickens  progressively 
through  consecutive  annual  reexposures  to  the 
atmosphere  as  it  simultaneously  flows  counter- 
clockwise around  the  subpolar  gyre. 

In  winter  data  the  seasonally  exposed  mode 
waters  form  a  smoothly  varying  ring  of  progres- 
sively colder  and  more  deeply  convecting  waters 
(see  figure  at  left).  Estimates  of  flow  speeds  in  the 
subpolar  gyre  suggest  a  transit  time  of  about  a 
decade  for  a  parcel  of  water  that  enters  the  trans- 
formation pipeline  east  of  Newfoundland  with  a 
temperature  of  12°  to  14°C,  travels  counterclock- 
wise, and  emerges  from  the  pipeline  in  the  Labra- 
dor Basin  at  temperatures  colder  than  4°  C.  The 
figure  below  left  illustrates  the  great  thickness  of 
the  regional  convection  in  this  mode  water  ring 
with  temperature/depth  profiles  from  the  New- 
foundland Basin  where  the  ring  begins,  west  of 
Ireland  in  the  northward  flow  of  the  eastern  sub- 
polar gyre,  in  the  Irminger  Basin  east  of 
Greenland  where  flow  is  westward,  and  in  the 
Labrador  Basin  where  the  mode  water  ring  ends. 

The  process  of  heat  liberation  from  ocean  to 
atmosphere  by  water  flowing  along  the  pipeline 
acts  as  a  regional  radiator,  particularly  for  north- 
ern Europe  where  the  westerlies  carry  the  heat 
extracted  from  the  ocean.  There  is  considerable 
evidence  for  interdecadal  variability  in  this  cli- 
mate process.  Our  first  evidence  comes  from  the 
Labrador  Sea  where  deep  wintertime  convective 
overturning  constitutes  the  last  stage  of  cooling 
along  the  transformation  pathway  and  vertically 
homogenizes  the  water  column  to  depths  some- 
times exceeding  2,000  meters,  creating  the  so- 
called  Labrador  Sea  Water  (coolest  profile  in 
figure  below  left). 

The  following  article  discusses  the  history  of 
Labrador  Sea  Water  (LSW).  The  figure  on  the 
opposite  page  shows  the  LSW  temperature  record 
overplotted  with  a  smoothed  version  of  the 
North  Atlantic  Oscillation  index,  an  expression  of 
the  relative  strength  of  the  atmospheric  westerlies 
(see  box  on  page  13).  The  LSW  temperature  his- 
tory shows  a  long  period  of  warming  from  the 
1930s  to  1971  followed  by  cooling  from  1972  to 
1993,  culminating  in  the  1990s  with  the  coldest, 
freshest,  and  thickest  LSW  ever  observed.  This 
cooling  trend,  however,  was  interrupted  by  a  brief 
warming  in  the  late  1970s  and  early  1980s.  Thus 
the  first  piece  of  evidence  for  climate  change  in 
the  warm  water  transformation  system  is  that  the 
system's  end  product,  LSW,  which  in  the  figure 
opposite  maps  in  winter  1962  as  3.3°C,  shows 
interdecadal  variability  with  temperatures  as 
warm  as  3.5°C  in  1970,  compared  to  as  cold  as 
3.1°C  forty  years  earlier  and  as  cold  as  2.7°C 
twenty-three  years  later. 

Several  agents  interact  to  produce  LSW.  The 


FALL/WINTER  1996 


LSW  within  the  Labrador  Basin  has  a  "residence 
time."  The  total  volume  of  LSW  is  only  partially 
replaced  each  year  through  the  addition  of  trans- 
formed warm  water  and  export  of  LSW  from  the 
Labrador  Basin  to  the  rest  of  the  North  Atlantic. 
To  visualize  this,  imagine  the  Labrador  Basin  as  a 
Jacuzzi.  Water  is  supplied  to  the  tub  at  some 
temperature  and  at  some  rate,  is  well  stirred  and 
mixed  in  the  tub,  and  that  mix  is  drained  away  at 
a  rate  that  matches  the  supply  rate.  The  residence 
time  is  the  volume  of  the  tub  divided  by  the  rate 
of  supply  and  is  a  measure  of  the  time  a  given 
parcel  of  water  spends  in  the  tub  before  going 
down  the  drain,  and  thus  is  the  average  time  the 
water  in  the  tub  loses  heat  to  the  overlying  colder 
air.  The  temperature  of  the  mixed  water  in  the 
tub  depends  on  the  temperature  history  of  the 
warm  water  being  fed  into  the  tub,  the  rate  of 
that  water  supply,  the  residence  time,  and  the  rate 
of  heat  loss  from  the  tub  to  the  air  above  it.  LSW 
temperature  similarly  depends  first  on  the  tem- 
perature history  of  the  product  emerging  trom 
the  warm  water  transformation  pipeline  into  the 
Labrador  Basin  plus  the  rate  of  that  flow,  and 
second  on  the  history  of  the  heat  exchange  be- 
tween the  Labrador  Basin  waters  and  the  overly- 
ing atmosphere  and  surrounding  ocean  (sort  of 
an  uninsulated  tub!).  Other  factors  can  influence 
such  histories:  Ice  atop  the  Jacuzzi  alters  how  its 
tub  temperature  evolves;  similarly,  sea  ice  and 
upper  ocean  freshwater  circulation  can  modify 
convective  cooling  in  the  Labrador  Basin. 

The  second  kind  of  evidence  for  climate 
change  signals  in  the  northern  North  Atlantic 
comes  from  sea  surface  temperature  (SST)  data. 
While  such  data  ultimately  represents  just  the 
skin  of  the  ocean,  it  has  the  advantage  of  much 
higher  data  density  in  space  and  time  than  sub- 
surface data,  for  it  can  be  collected  from  ships  of 
opportunity  without  requiring  a  specialized  re- 
search vessel  and  in  recent  decades  can  be  re- 
motely sensed  from  satellites.  SST  is  a  quantity  of 
great  importance  in  heat  exchange  with  the  over- 
lying atmosphere.  With  some  care  in  interpreta- 
tion, it  can  be  linked  to  conditions  beneath  the 
sea  surface.  Clara  Deser's  article  on  page  11  dis- 
cusses the  overall  regional  northern  North  Atlan- 
tic SST  history.  Subpolar  and  northern  subtropi- 
cal SSTs  were  overall  anomalously  warm  in  the 
1950s  and  1960s.  Winter  SST  data  reveals  that 
this  general  warmth  involved  warm  SST  anoma- 
lies propagating  with  the  oceanic  circulation.  The 
figure  on  the  next  page  documents  the  birth  in 
1951  of  a  warm  winter  SST  anomaly  east  of  New- 
foundland (red  patch)  and  traces  its  progression 
around  the  subpolar  gyre  in  subsequent  winters 
through  1968.  In  the  late  1960s,  a  cold  SST  area 
develops  in  the  subtropics  near  40°  N,  propagates 
into  the  subpolar  gyre  with  maximal  extent  in  the 


mid  1970s,  fades  around  1980,  and  reestablishes 
in  the  mid  1980s.  We  call  these  SST  "anomalies" 
because  they  represent  departures  from  the  long- 
term  local  temperatures. 

Comparison  of  this  interdecadal  progression 
of  warm  and  cold  SST  anomalies  to  the  LSW 
temperature  history  shows  provocative  parallels: 
The  post-World  War  II  LSW  warming  trend  occurs 
while  the  warm  SST  anomaly  travels  along  the 
transformation  pathway.  The  LSW  cooling  period 
beginning  in  1971  coincides  with  substantial  cold 
SST  areas  in  the  subpolar  gyre — but  that  general 
coldness  is  interrupted  by  warmer  SST  anomalies 


1940      1950 


1960      1970 


1980 


1990 


in  1980-81,  when  the  LSW  cooling  trend  was 
also  interrupted.  Thus  the  second  piece  of  evi- 
dence for  climate  change  in  the  warm  water 
transformation  system:  There  are  significant 
interdecadal  winter  SST  anomalies  in  the  subpo- 
lar gyre  as  a  whole  and  in  particular  moving 
along  the  warm  water  transformation  pathway. 
Periods  of  relatively  warm  SST  anomalies  along 
the  transformation  pipeline  correspond  to  peri- 
ods when  LSW  warmed,  and  periods  of  relatively 
cold  SST  anomalies  to  periods  when  LSW  cooled. 
These  observations  suggest  that  the  supply  of 
transforming  warm  water  to  our  Labrador  Basin 
Jacuzzi  was  running  warmer  in  the  1950s  and 
1960s  compared  to  the  succeeding  decades.  To 
further  that  idea,  a  link  between  SST  anomalies 
and  the  subsurface  water  undergoing  transforma- 
tion is  needed. 

The  third  kind  of  evidence  for  variability  in 
the  northern  North  Atlantic  climate  system 
comes  from  subsurface  hydrographic  data,  which 
allow  us  to  link  warm  and  cold  winter  SST  phases 
to  variability  in  the  mode  water  distribution.  The 
figure  on  page  23  shows  temperature  and  salinity 
difference  fields  at  400  meters  in  the  northern 
North  Atlantic.  This  depth  falls  in  the  winter 
convection  range  of  the  mode  water  along  the 
warm  water  transformation  pathway  (but  below 
the  depth  of  the  seasonal  warming  cap),  and  thus 
periods  of  warm  or  cold  400-meter  temperature 
correspond  to  the  array  of  isotherms  of  the  top 
figure  on  page  20  shifting  counterclockwise  or 
clockwise,  respectively.  The  fields  in  the  figure  on 
page  23  are  constructed  by  subtracting  the  tem- 
peratures for  one  time  period  from  the  preceding 
time  period  and  mapping  the  difference.  Our 


The  history  of  tem- 
perature (circles 
and  black  line)  of 
the  Labrador  Sea 
Water  convecting 
in  the  central  La- 
brador Basin  to 
depths  sometimes 
exceeding  2,000 
meters.  This  is 
compared  to  the 
interdecadal  march 
of  an  index  of  the 
North  Atlantic  Os- 
cillation, with  the 
reds  indicating  the 
high  index  periods 
of  strong  wester- 
lies, and  the  blues 
the  low  index  peri- 
ods of  weak  wester- 
lies. See  the  North 
Atlantic  Oscillation 
box  on  page  1 3  for 
an  explanation  of 
the  high  and  low 
states  of  the 
"NAO".TheNAO 
index  plotted  here 
is  formed  from  the 
sea-level  pressure 
difference  between 
the  subtropical 
Azores  high  pres- 
sure center  and  the 
subarctic  low  pres- 
sure center  near 
Iceland.  The  au- 
thors thank  lames 
Hurrell  (National 
Center  for  Atmo- 
spheric Research ) 
for  the  latest  up- 
dates of  NAO  in- 
dex data. 


OCEANUS 


A  time  series  of 
maps  derived  from 
winter  SST  mea- 
surements from 
1949  to  1996.  Pro- 
nounced warm 
SST  anomalies  are 
indicated  by  red 
and  pronounced 
cold  SST  anoma- 
lies by  blue. 


search  for  systematic  warming/cooling  is  reason- 
ably successful  for  the  period  after  World  War  II, 
but  breaks  down  in  more  recent  years  because 
less  data  was  collected  in  the  1980s.  Although 
more  measurements  have  been  made  in  the 
1990s,  data  generally  takes  several  years  to  be- 
come available  from  data  archive  centers.  The 
figure  on  page  23  shows  that  almost  the  entire 
subpolar  region  at  400  meters  warmed  in  1958- 
65  compared  to  1950-57.  Thus  as  the  warm  SST 
anomaly  traversed  the  transformation  pipeline, 
the  deep  winter  convection  temperatures  were 
warmer  than  usual. 

This  warming  trend  continued  in  the  more 
northern  area  through  1966-72,  but  reversed  to 
cooling  in  the  upstream  part  of  the  transforma- 
tion pipeline  between  Newfoundland  and  Ire- 
land. We  attribute  this  to  an  influx  into  the  pipe- 


line  of  abnormally  cold  subtropical  waters  — 
which  are  visible  in  SST  in  1969  to  1972  —  and  to 
winter  convection  beginning  to  run  abnormally 
cold.  In  the  final  panel  of  the  figure  on  the  oppo- 
site page  we  see  that  cooling  has  taken  over  al- 
most everywhere,  consistent  with  the  continued 
propagation  of  the  cold  SST  disturbance  along 
the  pipeline.  Thus  we  find  evidence  that  the  sup- 
ply pipeline  of  our  oceanic  Jacuzzi  has  slow 
interdecadal  variations  of  temperature,  and  that 
its  waiming  and  cooling  trends  are  in  phase  with 
the  trends  of  the  Jacuzzi  tub  water,  the  LSW. 
So  far,  what  we  have  described  are  several 
aspects  of  interdecadal  variability  of  upper  ocean 
signals  and  their  linked  behavior.  Our  fourth 
kind  of  evidence  for  changes  in  the  northern 
North  Atlantic  returns  to  the  LSW  temperature 
history  and  the  overplotted  NAO  index  recorded 


FALL/WINTER  1996 


60°N 


50 


40' 


60'w 


50'N 


40'w 


in  the  figure  on  the  previous  page. 
This  index  registered  high  values 
around  1950,  declined  to  very  low 
values  in  the  late  1960s,  then  rose  to 
very  high  values  again  in  the  early 
1990s,  indicating  an  interdecadal  cycle 
of  the  basic  index  of  the  North  Atlan- 
tic atmosphere's  mid-latitude  climatic 
state.  Dan  Cayan  of  the  Scripps  Insti- 
tution of  Oceanography  has  shown 
that  a  high  NAO  index  tends  to  en- 
hance liberation  of  heat  from  ocean  to 
atmosphere  over  the  Labrador  Basin 
while  a  low  NAO  index  tends  to  di- 
minish it.  Extrapolating  his  results  to 
this  interdecadal  time  scale,  the  1950s 
and  1960s,  with  NAO  index  declining, 
correspond  to  progressively  reduced 
cooling  over  the  Labrador  Basin.  Thus 
not  only  was  our  Jacuzzi  supply  pipe- 
line running  warm,  but  less  heat  was 
escaping  from  the  Jacuzzi  tub,  that  is, 
from  the  Labrador  Basin  to  the  overly- 
ing atmosphere.  Thus  the  weakening 
westerlies  reinforced  the  warming 
trend  for  LSW.  Conversely,  in  the  pe- 
riod of  strengthening  westerlies  in  the  1970s  and 
1980s,  not  only  was  the  transformation  pipeline 
running  cold,  but  the  loss  of  heat  from  the  ocean 
to  the  atmosphere  over  the  Labrador  Basin  was 
progressively  enhanced,  reinforcing  the  LSW 
cooling  trend. 

What  we  have  described  is  a  first  step  in  this 
sort  of  climate  change  work:  establishing  interde- 
pendence in  the  climate  system's  physical  proper- 
ties. The  second  phase  is  even  more  difficult  and 
challenging  than  the  first:  How  are  those  proper- 
ties linked  or  physically  coupled?  Which  "are  in 
charge,"  so  to  speak?  Is  the  ocean  a  passive  par- 
ticipant, merely  responding  to  atmospheric  forc- 
ing changes,  or  do  feedbacks  from  the  ocean  to 
the  atmosphere  force  evolution  of  the  atmo- 
spheric system's  climatic  state?  Certainly,  there  is  a 
large  heat  release  to  the  atmosphere  involved  in 
maintaining  the  mean  state  of  climate  in  the 
North  Atlantic  region,  and  it  makes  sense  that 
changes  in  the  warm  water  transformation  system 
ought  to  lead  to  changes  in  the  overlying  atmo- 
sphere. But  the  link  between  midlatitude  SST 
anomalies  and  their  potential  forcing  of  climate 
change  signals  has  been  surprisingly  elusive  to 
theory  and  modeling  efforts,  and  is,  in  fact,  one  of 
the  primary  unresolved  issues  in  climate  change 
research.  You  might  say  it  is  our  "missing  link." 

Continued  measurements  are  essential  to 
further  progress  in  unraveling  the  signals  and 
their  underlying  physics,  both  to  monitor  the 
evolution  of  the  system  and  to  sharpen  under- 
standing of  the  physics  of  specific  elements.  Two 


-2  0  -10          -02    02  10  2.0 

Temperature  Difference  at  400db  (  C) 


-0  50       -0".  10  -0  02   0 

Salinity  Difference 


02         010 
at  400db  ( 


of  us  (McCartney  and  Curry)  took  the 
Institution's  Research  Vessel  Knorr  to  the  subpolar 
gyre  in  fall  1996  to  begin  a  concentrated  two-year 
international  effort  directed  at  understanding  the 
seasonal  cycle  of  the  warm  water  transformation 
pipeline  of  the  eastern  subpolar  gyre. 

Mike  McCartney  came  to  the  Institution  in  1 973  with  a  me- 
chanical engineering/  fluid  mechanics  degree,  and  the  late  Val 
Worthington  supeivised  his  on-the-job  training  in  oceanography. 
His  interests  in  the  formation  of  higher  latitude  water  masses 
and  their  subsequent  circulation  have  taken  him  all  over  the 
world's  oceans,  and  he  has  been  doing  it  long  enough  to  person- 
ally observe  the  climatic  evolution  of  his  favorite  water  masses. 
His  goal  is  to  plan  longer  and  more  frequent  crimes  on  his  sail 
boat,  where,  unlike  on  research  vessels,  fancy  equipment  can  be 
restricted  to  GPS  navigation. 

Ruth  Curry  came  to  Woods  Hole  in  1 980  as  a  volunteer  at-sea 
walchstander  and  "mud  washer."  Sixteen  years — and  many 
cruises  — later,  the  sea  still  holds  its  allure,  but  Ruth  now  sails  as 
a  Chief  Scientist,  measuring  change!'  in  ivater  mass  properties 
and  ocean  circulation  in  search  of  pieces  to  the  global  climate 
pu^le.  Parenthood  has  changed  her  perspective  somewhat — 
(joins  '"  sea  lised  '"  mean  "°  sleep,  long  hours,  lousy  food,  and 
unpleasant  working  conditions,  she  says.  Now  it  means  getting 
more  sleep  than  normal  having  to  think  only  about  work,  and 
getting  all  your  meals  cooked  for  you— a  vacation! 
Hugo  Be^iek  joined  the  Scripps  Institution  of  Oceanography  in 
1970  following  completion  of  a  degree  in  physics.  After  four 
years  of  projects  in  undenvater  acoustics  and  air-sea  transfer,  he 
became  a  program  manager  at  the  Office  of  Naval  Research 
(and  ceased,  he  says,  doing  the  actual  work).  In  1980,  he 
moved  to  another  administrative  position  as  Director  of  AOML, 
a  NOAA  research  lab  that  focuses  on  climate-related  oceanogra- 
phy and  meteorolog}'.  "During  the  last  few  years, "  he  reports,  "I 
have  realized  the  error  of  my  ways  and  have  been  struggling 
mightily  to  redress  past  sins  and  return  to  honest  work  once 
again.    With  the  help  of  generous  people  such  as  my  co-authors, 
perhaps  such  an  eventuality  is  possible." 


Maps  of  tempera- 
ture and  salinity 
changes  for  succes- 
sive time  periods  at 
a  depth  near  400 
meters  in  the 
northern  North  At- 
lantic. Reds  and 
yellows  indicate 
that  the  later  of  the 
two  periods  is 
warmer  or  saltier, 
while  blues  and 
greens  indicate  the 
later  period  is 
cooler  or  fresher. 
Gray  areas  indicate 
small  salinity  and 
temperature 
change.  White  ar- 
eas indicate  a  lack 
of  data. 


Labrador  Sea  Water  Carries 
Northern  Climate  Signal  South 

Subpolar  Signals  Appear  Years  Later  at  Bermuda 


Ruth  G.  Curry 

Research  Associate,  Physical  Oceanography  Department 

Michael  S.  McCartney 

Senior  Scientist,  Physical  Oceanography  Department 

Changes  in  wind  strength,  humidity,  and 
temperature  over  the  ocean  affect  rates  of 
evaporation,  precipitation,  and  heat 
transfer  between  ocean  and  air.  Long-term  atmo- 


Density  Range  of 
Labrador  Sea  Water 


spheric  climate  change  signals  are  imprinted  onto 
the  sea  surface  layer — a  thin  skin  atop  an  enor- 
mous reservoir — and  subsequently  communi- 
cated to  the  deeper  water  masses.  Labrador  Sea 
Water  is  a  subpolar  water  mass  shaped  by  air-sea 
exchanges  in  the  North  Atlantic.  It  is  a  major 
contributor  to  the  deep  water  of  the  Atlantic,  and 
changes  of  conditions  in  its  formation  area  can 
be  read  several  years 
later  at  mid-depths  in 
the  subtropics.  Map- 
ping these  changes 
through  time  is  helping 
us  to  understand  the 
causes  of  significant 
warming  and  cooling 
patterns  we  have  ob- 
served at  these  depths 
in  the  North  Atlantic 
and  links  the  subtropi- 
cal deep  signals  back  to 
the  subpolar  sea  surface 
conditions. 

Labrador  Sea  Water 
(LSW)  is  the  end-prod- 
uct of  the  transforma- 
tion process,  described 
in  the  preceding  article, 
that  modifies  warm  and 
saline  waters  through 
heat  and  freshwater 
exchanges  with  the 
atmosphere.  In  the  last 
stage  of  this  transfor- 
mation, deep  winter- 
time convection  occurs 
in  the  Labrador  Basin 
between  North  America 
and  Greenland  where 


A:  Structure  of  the  North  Atlantic  basins  is  defined  by  bathymetric  contours  (2,000,  3,000,  and  4,000  meters) 
progressively  shaded  gray.  The  Mid-Atlantic  Ridge  separates  the  western  and  eastern  basins.  The  dominant  cur- 
rents at  mid  depths  (near  1,500  meters)  are  approximated  by  the  pink  lines  (bringing  relatively  warm  water 
northward)  and  the  blue  lines  (transporting  cold  waters  southward). 
B:  Thickness  in  meters  of  the  density  layer  corresponding  to  the  Labrador  Sea  Water  (LSW). 
C:  Temperature,  in  degrees  centigrade,  of  a  density  surface  in  the  middle  of  the  LSW  layer.  These  maps  high- 
light the  geographical  distribution  of  the  Labrador  Sea  Water  and  the  Mediterranean  Outflow  water  masses 
that  mix  to  produce  Upper  North  Atlantic  Deep  Water. 

D:  Temperature  profiles  from  the  Labrador  Basin  and  the  Mediterranean  Outflow  region  (locations  are  shown 
by  x  in  panel  A).  Green  shading  denotes  the  density  layer  that  is  mapped  in  panel  B  for  each  profile. 


strong  westerly  winds 
cool  the  surface  waters, 
making  them  denser 
than  the  underlying 
deep  water.  Convection 
occurs  when  the  denser 
surface  waters  sink  and 
mix  with  the  deep  water 


FALL/WINTER  1996 


to  produce  the  cold,  thick,  and  homogeneous 
LSW  water  mass. 

In  the  figure  at  left  (top  right  panel)  we  use  the 
thickness  of  the  LSW  mass  to  indicate  its  source 
region  and  the  pathways  along  which  it  spreads. 
In  the  North  Atlantic,  the  LSW  is  very  thick,  but  as 
the  circulation  carries  it  away  from  the  Labrador 
Basin,  mixing  with  other  thinner  water  masses 
progressively  erodes  its  thickness.  The  intense 
pink  colors  of  the  two  righthand  figures  opposite 
indicate  thicknesses  exceeding  2,000  meters  in  the 
LSW  formation  area.  Purple  colors  show  some- 
what thinner  LSW  spreading  northeast  into  the 
Irminger  Basin  east  of  Greenland,  eastward  via  the 
North  Atlantic  Current  into  the  Iceland  Basin,  and 
southwestward  along  the  western  boundary  into 
the  subtropical  basin. 

In  this,  the  cold,  fresh  and  thick  LSW  contrasts 
sharply  to  its  neighboring  North  Atlantic  water 
masses  of  the  same  density.  A  cool,  salty,  but  very 
thin  layer  of  Iceland-Scotland  Overflow  Water 
occupies  the  northeast  corner  of  the  map  as  the 
yellow-white  colors  on  both  thickness  and  tem- 
perature maps.  The  warm,  very  salty,  and  thin 
Mediterranean  Overflow  Water  is  represented  by 
the  green  (thickness)  and  yellow-red  (tempera- 
ture) tongues  extending  across  the  North  Atlantic 
from  its  source  region  at  the  Strait  of  Gibraltar. 
Recirculations  associated  with  the  Deep  Western 
Boundary  Current,  the  Gulf  Stream,  and  the 
North  Atlantic  Current  (an  extension  of  the  Gulf 
Stream)  mix  the  LSW  and  Mediterranean  waters, 
creating  the  intermediate  thicknesses  (blue  col- 
ors) and  temperatures  (green  colors)  between  the 
two  sources.  The  water  mass  that  results  from  the 
mixing  of  LSW  and  Mediterranean  Overflow 
Water  is  called  the  LIpper  North  Atlantic  Deep 
Water;  it  represents  one  of  the  major  elements 
exported  into  the  South  Atlantic  as  part  of  the 
global  conveyor  belt  (see  inside  front  cover). 

The  500-meter  thickness  contour  roughly 
separates  the  areas  where  LSW  strongly  influences 
this  layer  from  regions  where  Mediterranean 
Overflow  characteristics  predominate  and  shows 
that  these  LSW  influences  can  be  traced  south- 
wards along  the  western  boundary  all  the  way  to 
the  tropics.  Although  the  top  of  the  LSW  layer  is 
at  the  sea  surface  in  its  subpolar  source  region,  in 
the  subtropics  it  is  isolated  from  contact  with  the 
atmosphere  and  occupies  depths  between  1,200 
and  2,200  meters. 

LSW  properties — temperature,  salinity,  and 
thickness — have  changed  significantly  through 
time,  and  continued  measurements  in  the  Labra- 
dor Basin  since  the  1950s  enable  us  to  create  the 
time  series  shown  in  the  figure  above  right.  This 
record  shows  a  general  warming  from  the  1930s 
to  1971,  followed  by  a  cooling  trend  that  persists 
to  the  present.  Thickness  of  the  LSW  layer  is  di- 


Salinity  Anomalies  Occupy 
Labrador  Basin 


o.  30- 
E 


•  DOI« 


1930      1940      1950      1960      1970      19 


1930      1940      1950      1960      1970      1980      1990 


-C 
I- 


1940      1950      1960      1970      1980      1990 


rectly  related  to  the  intensity  of  wintertime  con- 
vection, with  strong  convection  producing  a  thick 
layer  and  weak  convection  resulting  in  a  relatively 
thin  layer.  Thick  conditions  in  the  1930s,  1950s, 
1970s,  and  1990s  indicate  periods  of  strong  con- 
vection and  loosely  correlate  to  cooler,  fresher 
LSW  conditions.  Note  the  abrupt  end  to  the 
1950s  and  1960s  warming,  increasing  salinity, 
and  thinning  with  the  onset  of  strong  convection 
in  1972.  This  cooling,  freshening,  and  thickening 
event,  however,  is  interrupted  by  a  period  of  weak 
convection  in  the  early  1980s.  Then,  by  1987,  the 
return  of  strong  convection  culminates  in  the 
coldest,  freshest,  and  thickest  conditions  ever 
measured.  The  figure  at  far  right  contrasts  vertical 
temperature  profiles  from  the  warm  and  cold 
phases  of  LSW  to  emphasize  the  extraordinary 
cooling  of  the  Labrador  Basin's  water  column 
over  the  past  25  years.  Note  that  this  cooling  has 
chilled  the  LSW  beyond  its  previous  cool  state 
more  than  60  years  ago,  in  the  1920s  and  1930s. 
Two  factors  principally  determine  LSW  prop- 
erty history:  the  strength  of  the  winds  and  the 
periodic  appearance  of  freshwater  anomalies  at 
the  sea  surface.  The  westerlies,  which  blow  cold, 
dry  air  from  Canada  across  the  Labrador  Basin, 
are  a  significant  factor  in  determining  the  depth 
of  Labrador  Basin  wintertime  convection.  An 
increase  in  wind  strength  removes  more  heat 
from  the  surface  waters  and  deepens  the  extent  of 
the  convection.  This  also  results  in  a  cooler  over- 
all LSW,  since  the  increased  heat  loss  at  the  sea 
surface  is  distributed  downward  as  the  water 
column  converts.  The  relative  strength  of  the 
westerlies  is  represented  by  the  North  Atlantic 
Oscillation  (NAO)  index.  (See  NAO  Box  on  page 
13.  The  NAO  index  is  defined  in  the  figure  cap- 
tion on  page  21.)  Overplotted  on  the  thickness 
axis  in  the  figure  above,  the  NAO  index  (shaded 


peralure  CO 


Left,  time  series  of 
Labrador  Sea  Water 
properties  in  its 
source  region. 
Thickness  is  the 
vertical  distance 
(meters)  between 
two  density  sur- 
faces that  bracket 
the  Labrador  Sea 
Water.  The  North 
Atlantic  Oscillation 
index  has  been 
overplotted  on  the 
thickness  axis  with 
high  index  shaded 
red  and  low  index 
blue.  Years  in 
which  surface  sa- 
linity anomalies 
occupied  the  La- 
brador Basin  are 
shaded  green. 
Above,  depth  pro- 
files of  temperature 
for  three  different 
years  contrasting 
the  Labrador  Sea 
Water  temperature 
in  its  cool  period 
before  World  War 
II  (1935),  the  peak 
of  warming  (1971), 
and  at  its  coldest 
point  (1993). 


OCEANUS   •   25 


Temperature  changes 
recorded  in  the 
1,500  to  2,500  meter 
layer  near  Bermuda. 
The  top  plots  show 
time  series  of  Ber- 
muda temperature 
anomaly  (red  curve) 
lagged  by  6  years, 
thickness  of  the  La- 
brador Sea  Water 
layer  (blue  curve)  in 
its  formation  area, 
and  temperature  of 
the  L.SW  core  (green 
curve).  The  lower 
plot  shows  a  lagged 
correlation  analysis 
for  Bermuda  tem- 
perature and  Labra- 
dor Sea  Water  (LSW) 
thickness,  which  is 
highest  for  lags  of  5 
to  7  years.  The  au- 
thors' interpretation 
is  that  subpolar 
thickness  anomalies 
result  in  variability 
of  the  volume  of 
LSW  entering  the 
subtropics.  A  large 
volume  of  LSW  shifts 
the  balance  of  influ- 
ence between  LSW 
and  Mediterranean 
Outflow  towards 
LSW.  When  the  LSW 
is  thick,  Bermuda 
sees  colder  (and 
fresher)  conditions 
about  6  years  later, 
while  a  thin  LSW 
source  results  in 
stronger  Mediterra- 
nean Outflow  influ- 
ence and  Bermuda 
sees  warmer  (and 
saltier)  conditions 
after  about  6  vears. 


red  for  high,  blue  for  low)  shows  trends  similar 
to  the  LSW  thickness:  declining  NAO  index  and 
thinning  LSW  from  the  1950s  to  1970,  a  pulse  of 
strong  westerlies  and  LSW  thickening  in  the  early 
1970s  followed  by  weak  westerlies  and  thin  con- 
ditions in  the  late  1970s,  then  extremely  strong 
westerlies  (high  NAO  index)  and  thick  condi- 
tions in  the  1990s.  Notice  also  in  the  figure  on 
the  previous  page  that  LSW  temperature  is  warm- 
ing during  low  NAO  periods  and  cooling  during 
years  of  high  NAO.  Thus  the  atmospheric  climate 
signal  becomes  imprinted  on  the  LSW  thickness 
and  temperature. 

The  thin  LSW  layers  of  1967-72  and  the  early 
1980s  correspond  to  buildups  of  extremely  low 
surface  salinity  conditions,  which  resulted  in  low 
surface  densities.  Because  this  surface  water  re- 
quired extraordinary  cooling  to  make  it  denser 
than  the  underlying  water,  it  completely  inhib- 
ited convection  in  the  Labrador  Basin.  These  two 
events,  referred  to  as  the  "Great  Salinity 
Anomaly"  and  the  "Lesser  Great  Salinity 
Anomaly,"  eventually  moved  around  the  subpo- 
lar gyre  (see  "If  Rain  Falls"  on  page  4).  The  first 
Great  Salinity  Anomaly  occupied  the  Labrador 
Basin  during  a  time  of  low  NAO  index  when 
weak  winds  would  have  reduced  the  convection 
depths  anyway.  However,  the  Lesser  Great  Salinity 
Anomaly  hindered  convection  in  the  early  1980s, 
which  resulted  in  a  thin  LSW  source  despite 
strong  westerlies  associated  with  a  relatively  high 
NAO  index.  Both  anomalies  ended  with  a  strong 
increase  in  convection  and  a  downward  mixing 
of  the  freshwater  cap,  which  dramatically  lowered 
the  LSW  core  salinity  and  temperature. 

The  subpolar  LSW,  carrying  the  imprinted 
climate  signals,  enters  the  subtropics  along  two 

Year  at  Bermuda 
1960  1970  1980 


1000 


$  1500 


2500 


Bermuda  Temperature  Anomaly 
lagged  by  6  years 


1950 


1960 


1970  1980 

Year  at  Labrador  Basin 


C  -0.6 
-t 

Oi 

°  -0.4 

c 
o 

1-0.2 


0.0 


4  6 

Time  Lag  in  Years 


principal  pathways:  The  deep  western  boundary 
current  transports  LSW  from  the  Labrador  Basin 
to  the  Caribbean  Islands,  and  the  Gulf  Stream 
and  North  Atlantic  Currents  carry  it  from  the 
western  boundary  out  into  the  interior  of  the 
ocean.  The  LSW  in  these  flows  is  strongly  stirred 
and  mixed  by  current  and  eddy  action  along 
these  pathways  and  in  the  ocean  interior.  In  the 
figure  on  page  24,  the  basin-scale,  deep-water 
properties  thus  represent  a  blending  of  the  LSW 
influence  with  other  influences,  principally  the 
Mediterranean  Overflow  Water.  The  resulting 
blended  water  mass,  known  as  the  Upper  North 
Atlantic  Deep  Water  (UNADW),  exhibits  a  tem- 
perature history  that  we  can  now  relate  to  varia- 
tions in  LSW  source  properties.  This  link  was 
previously  obscure  because  the  subtropical 
UNADW  temperature  signal  is  more  strongly 
influenced  by  the  LSW  thickness  history  than  by 
the  LSW  temperature  history.  Furthermore,  the 
time  the  ocean  requires  to  transport  and  mix  the 
LSW  into  the  subtropical  LINADW  introduces  a 
time  delay  to  the  link  between  these  signals. 
The  temperature  of  the  subtropical  mid 
depths  (1,000  to  2,500  meters)  has  generally 
warmed  since  the  1950s.  The  figure  below  left 
(red  curve)  shows  this  warming  trend  using  a 
long  time  series  measured  at  Bermuda  and  a 
recent  analysis  of  its  thermal  structure  by  Terry 
Joyce  and  Paul  Robbins  (see  "Bermuda's  Sta- 
tion S"  on  page  14).  When  a  time  lag  is  applied 
to  the  Bermuda  signal,  its  temperature  is  re- 
markably similar  to  the  LSW  thickness  (blue 
curve),  while  the  subpolar  LSW  temperature 
signal  (yellow  curve)  diverges  after  1975.  Corre- 
lation between  the  subtropical  temperature  and 
subpolar  thickness  signals  is  greatest  at  lags  of 

five  to  six  years  and 
1990  2000  implies  that  when 

subpolar  convection  is 
strong — and  the  LSW 
layer  is  thick — the 
subtropics  follow  five 
to  six  years  later  with 
cooler  temperatures  at 
mid  depths.  Con- 
versely, weak  convec- 
tion and  a  thin  LSW 
-0.2         layer  is  followed  by 
warmer  subtropical 
temperatures  approxi- 
mately six  years  later. 

To  place  this  rela- 
tionship into  a  geo- 
graphic context,  the 
figure  opposite  maps 
the  thickness  of  the 
LSW  density  layer  in  six 
different  time  frames, 


0.2 


0.1 


0.0 


-0.1 


1990 


10 


FALL/WINTER  1996 


each  spanning  about  7  years 
and  chosen  to  represent 
phases  of  LSW  source  varia- 
tion. As  noted  above,  the 
thickness  of  the  LSW  layer  in 
its  subpolar  formation  area 
changes  through  time  as  the 
convection  intensity  varies: 
The  Labrador  Basin  LSW 
source  is  thick  in  the  first  two 
time  frames,  extremely  thin 
in  1966-72  and  1980-86  (a 
thick  pulse  in  1973-79  sepa- 
rates these  two  periods),  and 
grows  to  extreme  thicknesses 
in  the  final  time  frame.  Away 
from  the  source  (near  the 
western  boundary  east  and 
south  of  Newfoundland  and 
east  of  New  England),  the 
layer  is  noticeably  thin  in  the 
1970s  and  1980s,  but  ro- 
bustly flooded  with  LSW  in 
the  1950s  and  1990s.  Over 
the  rest  of  the  subtropics 
(north  of  10°  latitude),  the 
LSW  layer  thickness  changes 
most  in  the  fourth  time 
frame  (1973-79)  when  the 
Mediterranean  Overflow 
Water  characteristics  (green 
colors)  are  extended  north- 
wards in  the  eastern  basin 
and  westwards  in  the  western 
basin.  Compare  the  areas 
around  the  Azores  and  south 
of  Bermuda  in  each  panel  to 
see  this  change. 

In  order  to  visualize  the 
impact  of  LSW  temperature 
and  thickness  anomalies  (changes  in  temperature 
and  thickness)  on  the  subtropics,  the  figure  on 
the  next  page  maps  the  temperature  and  thick- 
ness differences  in  one  time  frame  compared  to 
the  previous  time  frame  for  each  period  in  the 
figure  at  right.  The  patterns  of  anomalies  show 
large  areas  where  thickness  changes  correspond 
to  temperature  changes — where  the  layer  thins, 
temperatures  grow  warmer,  and  where  the  layer 
thickens,  temperatures  are  cooler— and  delineate 
where  LSW  exerts  a  strong  influence.  These  pat- 
terns also  show  consecutive  instances  where 
temperature  and  thickness  anomalies  of  one 
color  first  appear  in  the  Labrador  Basin,  rather 
quickly  move  southward  and  eastward  with  the 
western  boundary  current  and  North  Atlantic 
Current,  and  then,  one  time  frame  later,  anoma- 
lies of  the  same  color  appear  in  both  the  western 
and  eastern  subtropical  basins.  The  subtropical 


Thickness  of  the 
I,SW  density  layer 
for  six  consecutive 
time  periods. 


80:W       60:'W       40°W       20:'W 


0         250 


500  1,000  1,500  2,000 

Thickness  (meters)  O1500=  34.62-34.72 


2,500 


deep  water  anomalies  appear  to  lag  behind  the 
subpolar  LSW  signal  by  five  to  seven  years  as  the 
lagged  correlation  of  the  Bermuda  data  suggests. 

The  subtropical  temperature  anomalies  are 
large  compared  to  the  subpolar  temperature 
anomalies.  Because  a  signal  weakens  as  it  moves 
away  from  its  source,  these  subtropical  signals 
cannot  be  simply  the  advected  subpolar  tem- 
perature anomalies.  Rather,  the  time-delayed 
subtropical  response  to  LSW  source  variability 
represents  the  slow  adjustment  of  the  subtropi- 
cal deep  water  to  the  waxing  and  waning  of  LSW 
strength  so  clearly  visible  in  the  figure  above. 
The  thicker  the  LSW,  the  stronger  its  role  in 
mixing  with  Mediterranean  Overflow  water,  and 
this  is  manifested  as  an  eastward  and  southward 
erosion  of  the  influence  of  the  Mediterranean 
Overflow  Water  on  the  subtropical  deep  water. 
The  time  needed  for  the  LSW  to  circulate  and 


OCEANUS  •  27 


THICKNESS  DIFFERENCES 

i°W  60°W  40° W  20  "W       0 


TEMPERATURE  DIFFERENCES 

80°W  60°W  40°W  20"W       0° 


:l 

Nl 


1000 


300 


200 


100 


1-100 


-200 


-300 


1958-1965  compared  to  1950-1957 


1958-1965  compared  to  1950-1957 


1966-1972  compared  to  1958-1965  1966-1972  compared  to  1958-1965 


60  N 
50:'N 


0.20 


0.10 


0.08 


-0.08 


40"N 
30' N 
20°N 
10"N 
0'  CE^^^^^^^^^^K^^JE^^"^^^1'  1 1  n  i  n 

1973-1979  compared  to  1966-1972  1973-1979  compared  to  1966-1972 


1980-1986  compared  to  1973-1979  1980-1986  compared  to  1973-1979 


1987-1994  compared  to  1980-1996 


1987-1994  compared  to  1980-1996 


Thickness  difference  fields  (left  column)  and  temperature  difference  fields 
(right  column)  were  constructed  by  subtracting  thickness  or  temperature  in  two 
consecutive  time  frames  at  each  1 -degree  square  in  the  North  Atlantic.  The 
thickness  represents  the  LSW  density  layer  and  temperature  values  are  taken  at  a 
density  surface  in  the  middle  of  that  layer.  Green-blue  colors  indicate  layer 
thickening  and/or  cooling  in  one  time  frame  compared  to  the  previous  time 
frame;  yellow-red  colors  indicate  layer  thinning  and/or  warming. 


mix  into  the  subtropical  basins  results  in  a  de- 
layed appearance  of  the  response.  When  the 
LSW  is  thinner  than  normal,  the  Mediterranean 
Overflow  Water  exerts  more  influence,  and  this 
appears  as  a  westward  and  northward  extension 
of  the  thin,  warm,  and  salty  characteristics. 

Understanding  the  nature  of  the  subtropical 
temperature  variations  and  knowing  that  the 
subpolar  convection  has  been  extremely  strong 
from  1988  to  1995  enables  us  to  predict  that  the 
subtropical  mid  depths  will  continue  to  cool 
through  the  1990s.  Tracing  the  extremely  cold, 
fresh,  and  thick  signal  that  is  now  invading  the 
subtropics  (quite  pronounced  in  the  bottom 
panels  of  the  figure  on  the  next  page)  will  pro- 
vide us  with  valuable  information  concerning  the 
timing  and  geography  of  the  complex  mid-lati- 
tude circulation  system  whose  end  product,  the 
Upper  North  Atlantic  Deep  Water,  is  exported  to 
the  southern  ocean. 

Our  WHOI  colleague  Bob  Pickart  (see 
Oceanus,  Spring  1994)  has  tracked  the  penetra- 
tion of  the  extreme  LSW  along  the  deep  western 
boundary  current  and  Gulf  Stream  system  off 
New  England,  and  our  University  of  Miami  col- 
leagues Rana  Fine  and  Bob  Molinari  have  recently 
(summer  1996)  sighted  this  extreme  LSW  signal 
in  the  deep  western  boundary  current  off  Abaco 
in  the  Bahamas— one  of  the  most  exciting  and 
valuable  results  of  their  decade-long  monitoring 
program  at  that  location. 

We  are  planning  a  1998  field  experiment  to 
take  advantage  of  this  unique  climate  change 
signal  by  measuring  the  subtropical  western 
basin's  response  to  the  LSW  invasion  at  24°  N 
and  15°  N.  Because  of  the  time  delay  observed  in 
the  subtropical  response,  we  can  be  reasonably 
confident  that  we  will  be  in  the  right  places  to 
measure  this  extreme  LSW  event,  for  we  know 
through  the  continued  efforts  of  our  Canadian 
colleagues  John  Lazier  and  Allyn  Clarke  (Bedford 
Institute  of  Oceanography)  that  the  LSW  source 
continued  to  convert  through  the  winter  of  1995. 
They  report  a  cessation  of  deep  LSW  convection 
in  winter  1996.  If  that  cessation  is  longer-lived 
than  a  single  anomalous  winter  event,  then  we 
would  expect  it  to  appear  as  a  subtropical  climate 
change  signal  in  2000  to  2002. 

The  authors'  research  is  jointly  funded  by  the  National 
Science  Foundation-sponsored  World  Ocean  Circulation 
Experiment  Program  and  the  Climate  and  Global  Change 
Program  of  the  National  Oceanic  and  Atmospheric  Admin- 
istration. The  authors  thank  Terry  Joyce  for  collaboration 
in  producing  the  figure  on  page  26,  James  Hurrel  (Na- 
tional Center  for  Atmospheric  Research)  for  providing  the 
most  recent  update  of  the  NAO  index  data,  and  John 
Lazier  for  providing  the  Labrador  Basin  data  for  recent 
years  and  for  his  sustained  effort  for  more  than  30  years  in 
maintaining  critical  time-series  measurements  in  the 
hostile  environment  of  the  Labrador  Basin. 


28  *  FALLA/VINTER  1996 


Transient  Tracers  Track 
Ocean  Climate  Signals 


William  I.  Jenkins 

Senior  Scientist,  Marine  Chemistry  &  Geochemistry  Dept. 

William  M.  Smethie,  Jr. 

Senior  Research  Scientist,  Lamont-Doherty  Earth 
Observatory,  Columbia  University 

Transient  tracers  provide  us  with  a  unique 
opportunity  to  visualize  the  effects  of  the 
changing  climate  on  the  ocean.  They  trace 
the  pathways  climate  anomalies  follow  as  they 
enter  and  move  through  the  ocean  and  give  us 
valuable  information  about  rates  of  movement 
and  amounts  of  dilution.  This  knowledge  is 
important  for  developing  ocean-climate  models 


to  predict  long  term  climate  changes. 

Humankind's  activities  have  resulted  in  the 
release  of  a  number  of  globally  distributed  sub- 
stances into  the  environment.  These  substances 
enter  the  oceans,  and,  although  they  have  little,  if 
any,  impact  on  the  environment,  they  travel 
through  and  "trace"  the  biological,  chemical,  and 
physical  pathways  of  the  ocean.  The  distributions 
of  these  "tracers"  change  with  time.  For  example, 
isotopes  created  by  atmospheric  nuclear  weapons 
tests  in  the  1950s  and  1960s  were  introduced  in  a 
pulselike  fashion,  while  atmospheric  concentra- 
tions of  chlorofluorocarbons  (CFCs),  which 


A  bird's  eye  view  of 
the  distribution  of 
tritium  in  the 
North  Atlantic.  Pic- 
ture yourself  float- 
ing a  few  hundred 
miles  above  Nor- 
way, looking 
southwestward 
down  at  the  North 
Atlantic.  North 
America  is  in  the 
top  right  corner  of 
the  view, 
Greenland  to  the 
lower  right,  and 
parts  of  Europe, 
Great  Britain,  and 
Africa  are  visible 
on  the  lower  left. 
The  spikes  are 
ocean  islands.  The 
blue  "blanket"  is 
the  1  Tritium  Unit 
isosurface  (surface 
of  constant  tritium 
measured  in  1981). 
(One  Tritium  Unit 
equals  one  tritium 
atom  to  10 '"  hy- 
drogen atoms.) 
Underneath  this 
blanket  lies  water 
that  has  not  been 
appreciably  venti- 
lated (in  contact 
with  the  atmo- 
sphere) while  wa- 
ter above  this  level 
has  been  ventilated 
since  the  1960s. 


OCEANUS 


threaten  the  earth's  ozone  layer,  have  been  in- 
creasing with  time.  We  refer  to  these  substances 
in  the  ocean  as  "transient  tracers"  because  their 
distributions  are  evolving. 

Transient  tracers  are  valuable  tools  for  study- 
ing ocean  climate.  First,  because  they  are  new  to 
the  ocean  environment,  they  are  indicators  of 
"ocean  ventilation."  Ventilation  is  the  imposition 
of  atmospherically  derived  properties  on  water 
masses.  For  example,  waters  in  contact  with  the 
atmosphere  will  have  dissolved  oxygen  concen- 
trations increased  to  equilibrium  values  with  the 
atmosphere.  Providing  their  time  history  in  the 
atmosphere  is  known  and  the  manner  in  which 
they  are  transferred  to  the  ocean  is  understood, 
they  can  be  used  to  construct  and  test  models  of 
ocean  ventilation  and  circulation.  Observations 
of  their  distributions  in  the  ocean  and  time  series 
measurements  of  how  they  change  with  time  are 


2500 


1970 


1975 


1980 


A  time  series  of  tri- 
tium in  the  Sar- 
gasso Sea  near  Ber- 
muda The  plot  of 
tritium  vs.  depth 
and  time  shows  the 
sudden  arrival  of 
tritium  at  interme- 
diate depths  (1,000 
to  1,500  meters)  in 
the  late  1970s,  and 
at  deeper  depths 
(2,000  to  2,500 
meters)  in  the  late 
1980s.  These 
events  correspond 
to  the  onset  of 
cooling  at  these 
levels,  and  signal 
the  arrival  of  newly 
ventilated  waters  in 
response  to  climate 
changes  farther 
north. 


powerful  tools:  They  provide  direct  visualization 
of  climate  changes,  and  they  trace  the  pathways 
along  which  ocean  climate  perturbations  propa- 
gate into  the  oceans.  That  is,  changes  in  charac- 
teristics and  volumes  of  water  masses  due  to 
climate  variations  ultimately  influence  deeper, 
more  isolated  regions  of  the  oceans.  How  these 
changes  move  to  the  deep  ocean  from  regions  of 
contact  with  the  atmosphere  must  be  under- 
stood. This  process  is  an  important  mechanism 
whereby  the  oceans  couple  to  the  atmosphere  on 
longer  time  scales,  and  probably  plays  a  role  in 
determining  the  interannual  to  decadal  variations 
in  global  climate. 

Observations  of  tracer  distributions  provide 
information  on  processes  that  are  very  difficult 
to  observe  any  other  way.  Mixing  and  dilution, 
for  example,  play  a  dominant  role  in  the  south- 
ward transport  of  material  along  the  deep  west- 
ern boundary  of  the  North  Atlantic.  It  has  long 
been  known  that  newly  ventilated  North  Atlantic 


Deep  Water  travels  southward  in  a  concentrated 
current,  hugging  the  western  edge  of  the  Atlantic 
basin.  Although  direct  current  measurements 
indicate  velocities  of  tens  of  centimeters  per 
second,  the  actual  average  propagation  rate  of 
tracers  down  the  western  boundary  is  only  one 
or  two  centimeters  per  second.  This  is  because 
there  is  a  tremendous  amount  of  entrainment 
and  mixing  associated  with  water  recirculating 
within  the  rest  of  the  basin.  The  mixing  slows  the 
progress  of  tracers  and  climate  anomalies.  Tran- 
sient tracers  are  perhaps  the  only  tools  for  mea- 
suring the  amount  of  interior  exchange  and 
downstream  propagation  rates. 

Tritium:  the  Cold  War  Legacy 

It  is  said  that  every  cloud  has  a  silver  lining, 
and  that  seems  to  be  true  even  if  it  is  a  mush- 
room cloud.  Although  the  atmospheric  testing  of 
nuclear  weapons  re- 
leased alarming 
amounts  of  radioactive 
debris  into  the  environ- 
ment, and  caused  un- 
told damage  to  both 
the  environment  and 
human  health,  it  also 
provided  oceanogra- 
phers  with  some 
unique  tools  to  study 
ocean  circulation  and 
ventilation.  We  have 
had  the  opportunity  to 
observe  the  entry  of 
these  substances  into 
the  oceans,  and  to  see 
how  they  are  moved 
around  by  physical, 
chemical,  and  biological  processes. 

One  of  these  bomb-produced  tracers  is  tri- 
tium, a  radioactive  isotope  of  hydrogen.  There  is 
very  little  natural  tritium  in  the  world  (the  entire 
global  inventory  would  only  weigh  a  few  kilo- 
grams!), but  several  hundred  kilograms  were 
created  in  the  hydrogen  bomb  explosions.  This 
tritium  was  injected  into  the  stratosphere,  almost 
immediately  oxidized  to  water,  and  fairly  rapidly 
rained  out  onto  the  surface  of  the  earth.  Since  it 
is  chemically  bound  up  as  part  of  the  water  mol- 
ecule (it  is,  after  all,  hydrogen),  tritium  is  an 
ideal  tracer  of  the  hydrologic  system.  Further, 
since  the  bulk  of  the  atmospheric  weapons  test- 
ing was  done  in  the  northern  hemisphere,  most 
of  the  tritium  was  deposited  in  the  high  latitude 
northern  hemisphere.  Thus  it  is  an  ideal  tracer  of 
ventilation  and  water  mass  formation  in  the 
North  Atlantic. 

We  can  see  this  in  the  figure  on  the  previous 
page,  which  provides  a  bird's  eye  view  of  tritium 


1985 


/INTER  1996 


distribution  in  the  North  Atlantic  in  1981.  Picture 
yourself  hovering  somewhere  over  Norway,  at  an 
altitude  of  a  few  hundred  miles.  You  are  looking 
southwestward  and  downward  on  an  isosurface 
of  tritium  in  the  North  Atlantic.  An  isosurface  is 
the  two-dimensional  analog  of  a  contour  line  on 
a  map,  shown  here  in  a  three-dimensional  ocean. 
The  blue  "blanket"  you  see  in  the  picture  is  the 
1  Tritium  Unit  isosurface,  where  we  find  a  ratio 
of  1  tritium  atom  to  100,000,000,000,000,000 
hydrogen  atoms.  This  isosurface  corresponds  to 
about  5  or  10  percent  of  the  maximum  surface 
water  concentrations  of  tritium  during  the  mid 
1960s,  when  it  was  at  its  peak.  Ail  the  water  be- 
neath this  blanket  has  remained  relatively  iso- 
lated from  the  tritium  invasion,  and,  conversely, 
all  the  water  above  this  blanket  has  been  at  the 
sea  surface,  exposed  to  the  atmosphere,  and  thus 
ventilated  or  otherwise  involved  in  interaction 
with  the  surface  ocean  in  the  15  to  20  years  be- 
tween the  bomb  tests  and  this  survey. 

The  blanket  lies  at  about  500  to  1,000  meters 
depth  in  the  subtropics,  but  deepens  to  1,500  to 
2,000  meters  just  south  of  the  Gulf  Stream  off  the 
New  England  coast.  This  is  the  effect  of  the  Gulf 
Stream  recirculation,  a  tight  gyre  that  effectively 
ventilates  the  upper  part  of  the  ocean  in  this 
region.  There  is  also  a  fold  extending  southward 
from  this  region  at  about  1,200  meters  depth, 
marking  the  intrusion  of  intermediate  depth 
waters  toward  the  tropics.  Most  notably,  however, 
is  the  dramatic  dive  that  the  "blanket"  takes  to 
the  north,  disappearing  into  the  ocean  floor.  The 
track  along  which  this  happens  parallels  the  Gulf 
Stream  Extension/North  Atlantic  Drift.  All  the 
waters  north  of  this  line  have  been  ventilated  to 
the  ocean  floor  on  10  to  20  year  time  scales.  This 
is  a  powerful  statement  regarding  the  time  scales 
of  ocean  ventilation,  and  has  profound  implica- 
tions concerning  how  rapidly  climatic  variations 
can  propagate  through  the  oceans. 

We  can  see  this  in  yet  another  way.  The  figure 
at  left  is  a  plot  of  tritium  vs.  depth  for  a  time 
series  near  Bermuda  from  the  late  1960s  to  the 
late  1980s.  The  tritium  data  has  been  adjusted  for 
decay  to  the  same  date  (January  1 ,  1 981 )  to  allow 
us  to  more  clearly  see  the  time  trends.  The  most 
obvious  trend  is  the  downward  propagation  of  a 
tritium  maximum  (deeper  red)  from  the  surface 
in  the  late  1960s  to  about  400  meters  depth  in 
the  late  1980s,  a  downward  penetration  rate  of 
18  meters  per  year.  However,  from  the  perspective 
of  climate  change,  the  most  important  signal  is 
the  sudden  increase  in  tritium  (green)  at  about 
1,500  meters  in  the  late  1970s  and  at  2,500 
meters  in  the  mid  1980s.  Both  of  these  features 
correspond  to  the  onset  of  significant  cooling 
events  seen  in  the  deep  water  at  this  station.  The 
correspondence  between  the  sudden  tritium 


67       66     65      64 


62 


10      9 


\ 


March  1991 


0     50     100    150    200    250  0     50     100    150    200    250 

3.0     2.5      20     16      14     12      1.0     0.9     0,8     0.7     0.6     0.5     0.4     0.3     0.2     0.1      0.0 

J  p/como/es/ 


4500 


5000 


SO 


Distance  (kilometers) 


increases  and  cooling  offsets  is  highly  suggestive 
of  significant  changes  in  deep  water  ventilation  at 
those  times.  Indeed,  if  we  could  take  another 
"picture"  of  the  tritium  blanket  shown  in  the 
figure  on  page  29,  we  would  see  that  it  has  been 
pushed  further  downward  and  southward  by  a 
climatic  event. 

Another  Kind  of  Cold  War  Legacy 

The  development  and  manufacture  of  chlorof- 
luorocarbons  (CFCs)  for  use  in  refrigerators  and 
air  conditioners  (and  later  as  spray  can  propel- 
lants)  seemed  like  a  good  thing  at  the  time:  CFCs 
were  easy  to  manufacture,  nontoxic,  chemically 
inert,  and  stable.  Production,  use,  and  ultimate 
release  of  CFCs  into  the  atmosphere  increased 
annually  in  an  almost  exponential  fashion  from 
their  introduction  in  the  1930s.  The  unfortunate 
influence  of  these  compounds  on  the  ozone 
layer,  however,  has  lead  to  international  reduc- 
tion in  their  manufacture  and  use.  In  1990,  the 
US  and  55  other  nations  agreed  to  end  CFC  pro- 
duction by  2000.  Meanwhile,  however,  oceanog- 


Four  chlorofluoro- 
carbon  (CFC)  sec- 
tions taken  at  vari- 
ous times  along 
55°  W  south  of  the 
Grand  Banks.  Note 
the  absence  of  any 
significant  CFC  sig- 
nal at  the  depth  of 
the  Labrador  Sea 
Water  (about  1,500 
meters  depth)  in 
1983,  but  the  sud- 
den flooding  of 
these  depths  with 
CFCs  in  the  later 
sections,  as  newly 
formed  Labrador 
Sea  Water  flows 
around  the  Grand 
Banks  and  into  the 
Sargasso  Sea.  These 
changes  corre- 
spond to  the  tri- 
tium increases  seen 
in  the  Bermuda 
time  series  (see  fig- 
ure opposite). 


raphers  have  found  another  "silver  lining"  in  this 
ecological  cloud,  which  has  permitted  us  to  study 
ocean  ventilation  and  circulation:  Waters  that  have 
been  in  contact  with  the  atmosphere  in  the  past 
few  decades  have  taken  up  some  of  these  com- 
pounds, and  hence  have  been  labeled  in  a  distinc- 
tive way.  Thus  the  distribution  of  CFCs  in  ocean 
water  provides  us  with  important  clues  regarding 
the  pathways  of  newly  formed  water  masses. 

In  the  1970s  and  early  1980s,  there  was  not 
much  winter  time  convection  and  formation  of 
Labrador  Sea  Water.  In  fact,  tracer  sections  (lines 
of  stations)  taken  across  the  Deep  Western 
Boundary  Current  to  the  south  showed  a  decided 
lack  of  newly  venti- 


A 


2000  4000  6000  8000 

Down  Stream  Distance  (kilometers) 


10000 


The  downstream 
evolution  of  tri- 
tium (upper  panel, 
in  tritium  units) 
and  tritium-he- 
lium age  (lower 
panel,  in  years)  vs. 
distance  in  the 
core  of  the  deep 
western  boundary 
current.  Note  the 
approximately  ten- 
fold reduction  in 
tritium  content  in 
the  Deep  Western 
Boundary  Current 
core  due  to  dilu- 
tion with  older, 
surrounding  deep 
water,  and  the  lin- 
ear increase  in  age 
downstream.  The 
age  increase  is  con- 
sistent with  a 
mean  speed  of 
about  1.5  centime- 
ters per  second. 


lated  Labrador  Sea 
Water.  This  hiatus  is 
somehow  related  to 
the  complex  interplay 
between  changing 
climatic  conditions  in 
the  area  and  the  fresh- 
water outflow  and 
budgets  of  the  Arctic. 
The  late  1980s  and 
early  1990s  have  her- 
alded a  dramatic 
change  in  climatic 
conditions  in  the  La- 
brador Sea.  These 
changes  have  resulted 
in  the  production  of  a 
large  amount  of  Labra- 
dor Sea  Water,  which  is 

now  invading  the  ocean  interior.  You  can  see  this 
beginning  to  happen  in  the  figure  on  the  previ- 
ous page,  a  time  series  of  CFC  sections  made 
along  55°  W  south  of  the  Grand  Banks.  Labrador 
Sea  Water  occurs  in  these  sections  at  a  depth  of 
about  1,500-2,000  meters  (the  middle  heavy 
dashed  line).  Notice  that  there  was  very  little 
CFC- 11  in  this  water  in  the  1983,  although  there 
is  a  CFC  tongue  at  a  shallower  level  characteristic 
of  waters  that  are  formed  in  the  southeastern 
corner  of  the  Labrador  Sea  (the  shallowest  heavy 
dashed  line).  Below  the  Labrador  Sea  Water  core, 
there  is  a  weak  but  detectable  CFC  core  in  waters 
characteristic  of  Denmark  Straits  Overflow  water 
(marked  here  by  the  deepest  dashed  line).  Com- 
bined, these  three  water  masses  form  the  Deep 
Western  Boundary  Current  system  of  the  North 
Atlantic,  and  are  responsible  for  the  southward 
transport  of  newly  ventilated  waters. 

However,  in  the  1990s,  there  is  a  sudden  in- 
crease in  the  amounts  of  CFCs  in  the  Labrador 
Sea  Water  core,  as  well  as  a  steady  increase  in  the 
deeper  core  associated  with  waters  from  the  Den- 
mark Straits  overflow.  This  increase  is  continuing 
through  the  1990s  and  is  direct  evidence  of  the 


newly  ventilated  waters'  arrival.  This  again  is  a 
signature  of  the  penetration  of  climatic  anoma- 
lies into  the  ocean  interior. 

While  the  55°  W  sections  capture  the  invasion 
of  the  newly  formed  waters  into  the  northern 
Sargasso  Sea,  the  pathway  southward  is  not  a 
simple  one.  The  Deep  Western  Boundary  Current 
is  not  a  continuous  ribbon  of  flow  extending  all 
the  way  from  the  Grand  Banks  to  the  equator, 
but  rather  a  composite  consisting  of  series  of 
interconnected  gyres  lined  up  along  its  path.  A 
fluid  parcel  that  passes  by  the  Grand  Banks  may 
spend  most  of  its  time  looping  through  these 
gyres,  and  only  part  of  its  time  in  the  Deep  West- 
ern Boundary  Current.  This  is  why  the  mean 
propagation  speed  of  tracers  down  the  western 
boundary  is  only  1  to  2  centimeters  per  second, 
while  velocities  in  the  actual  core  of  the  Deep 
Western  Boundary  Current  are  10  to  20  centime- 
ters per  second. 

The  figure  at  left  is  a  plot  of  tritium  and  tri- 
tium-helium age  in  the  core  of  the  Deep  Water 
Boundary  Current  vs.  distance  downstream  from 
its  origin.  We  see  that  the  core  becomes  progres- 
sively older  (about  20  years  in  10,000  kilometers) 
corresponding  to  a  speed  of  about  1.5  centime- 
ters per  second.  Notably,  the  tritium  concentra- 
tion in  the  core  decreases  more  than  tenfold 
downstream,  partly  due  to  decay,  but  largely  due 
to  dilution  and  mixing  with  older,  tritium-free 
waters.  This  process  of  mixing  is  an  important 
mechanism  for  ventilating  the  abyssal  ocean.  It's 
through  this  process  that  climate  anomalies  make 
their  way  into  the  deep  ocean. 

Thus  the  transient  tracers  are  telling  us  some- 
thing very  important  about  the  propagation  of 
climatic  changes  into  the  deep  ocean.  They  high- 
light the  pathways  and  give  us  the  rates  of  move- 
ment and  dilution  in  the  ocean.  This  information 
is  valuable  because  the  ocean  provides  the  long 
term  memory  and  feedback  in  the  coupled 
ocean-atmosphere-climate  system,  and  is  the  key 
to  beginning  to  make  long  term  predictions  in 
our  ever  changing  climate. 

The  research  discussed  in  this  article  was  supported  by 
the  National  Science  Foundation  and  the  Office  of  Naval 
Research. 

Bill  lenkins  nailed  life  as  a  nuclear  physicist  but  drifted  into 
environmental  sciences  out  of  a  secret  yearning  to  become  a 
forest  ranger.  Not  having  a  good  sense  of  direction,  and  tearing 
bliicl;  flies,  however,  he  ended  up  as  an  oceanographer  on  Cape 
Cod.  He  joined  the  WHO/  Chemistry  Department  (now  the 
Department  of  Marine  Chemistry  and  Geochemistry)  in  1974. 

Bill  Smethie's  interest  in  oceanography  began  during  childhood 
summers  spent  at  his  grandfather's  log  cabin  an  the  Virginia 
side  of  the  Potomac  River.  He  embarked  on  his  first  oceano- 
graphic  cruise  at  age  7  when  he  attached  a  makeshift  sail  to  his 
inner  tube  ami  set  sail  for  the  other  side  of  the  river.  His  doting 
tuints  prevented  him  from  making  it  to  the  other  side,  but  ever 
since  he  has  had  a  never-ending  curiosity  for  what  lies  beyond 
the  horizon.  He  joined  the  Geochemistry  Division  of  Lamont- 
Doherty  Geological  Obseivatoiy  in  1979. 


L/WINTER  1996 


New  Data  on  Deep  Sea  Turbulence 
Shed  Light  on  Vertical  Mixing 

Rough  Seafloor  Topography  Has  Far-Reaching  Effect 


John  M.  Toole 

Senior  Scientist,  Physical  Oceanography  Department 

The  global  thermohaline  circulation  is  basi- 
cally a  wholesale  vertical  overturning  of 
the  sea,  driven  by  heating  and  cooling, 
precipitation  and  evaporation.  (Changes  in 
temperature=f/!crmo,  changes  in  salinity=/w/mi'.) 
Bottom  waters  move  equatorward  from  their 
high-latitude  regions  of  formation  (the  cold  limb 
of  the  circulation),  upwell,  and  return  poleward 
at  intermediate  depth  and/or  the  surface  (the 
warm  limb).  As  the  bottom  waters  are  colder 
than  the  overlying  waters,  this  circulation  is  re- 
sponsible for  a  large  fraction  of  the  ocean's 
poleward  heat  transport.  In  addition,  these  flows 
often  redistribute  fresh  water,  as  the  northward 
and  southward  moving  waters  generally  have 
different  salinities. 

These  oceanic  heat  and  water  transports  play  a 
significant  role  in  Earth's  climate.  The  earth  gains 
heat  from  the  sun  at  low  latitude,  and  radiates 
heat  back  to  space  about  the  poles.  To  maintain  a 
quasi-steady  state,  the  ocean-atmosphere  system 
must  carry  heat  from  low  to  high  latitude.  At 
mid-latitudes,  where  the  poleward  heat  flux  is 
maximum,  the  oceanic  and  atmospheric  contri- 
butions are  about  equal.  One  component  of  the 
atmospheric  heat  transport  involves  evaporation, 
water  vapor  transport,  and  its  subsequent  con- 
densation. Net  north/south  water  vapor  transport 
in  the  atmosphere  is  balanced  by  liquid  water 
transport  by  rivers  and  ocean  currents. 

For  almost  200  years,  since  the  writing  of 
Count  Rumford  in  1797,  there  has  been  a  basic 
understanding  of  the  cold  limb  of  the  thermoha- 
line circulation.  The  combination  of  atmospheric 
cooling,  evaporation,  and,  in  some  cases,  salt 
rejection  during  the  formation  of  sea  ice  causes 
surface  waters  at  high  latitudes  to  become  suffi- 
ciently dense  that  they  sink  to  the  ocean  bottom. 
These  newly  formed  deep  waters  subsequently 
spread  horizontally  within  the  constraints  of  the 
seafloor's  bathymetry  to  renew  the  deep  waters 
found  in  the  interiors  of  the  world's  oceans. 
There  are  two  principal  formation  sites  for  dense 
bottom  water:  the  Greenland  and  Norwegian 
Seas  of  the  northern  North  Atlantic  Ocean,  and 


around  the  Antarctic  continent,  particularly 
within  the  Weddell  Sea.  Together,  these  source 
regions  export  some  20  to  30  million  cubic 
meters  per  second  of  bottom  water  to  the  other 
ocean  basins.  (For  comparison,  the  chiefly  wind- 
driven  Gulf  Stream,  Kuroshio,  and  Agulhas  Cur- 
rents carry  in  excess  of  100  million  cubic  meters 
per  second  within  horizontal  circulations.) 

The  processes  involved  with  the  return  limb  of 


The  principal  tool  for  work  described  in  this  article  is  the  high  resolution 
profiler.  It  records  temperature,  salinity,  pressure,  and  horizontal  velocity  10 
times  per  second  on  descent  to  the  ocean  floor,  then  returns  to  the  surface.  For  a 
detailed  discussion  of  the  instrument's  development,  see  the  Spring/Summer 
1995  issue  of  Oceanm. 


PACIFIC 


Circulation  sche- 
matic of  the 
world's  major  wa- 
ter masses  (also  see 
inside  front  cover). 
Of  concern  here 
are  the  mixing  pro- 
cesses that  modify 
the  bottom  and 
deep  waters  within 
the  cold-to-warm 
limbs  of  the  over- 
turning circulation. 


Subantarctic  Mode  Water 
Antarctic  Intermediate  Water 
Red  Sea  Overflow  Water 
Antarctic  Bottom  Water 
North  Pacific  Deep  Water 
Antarctic  Circumpolar  Current 
Grcumpolar  Deep  Water 
North  Atlantic  Deep  Water 
Upper  Intermediate  Water 
Indian  Ocean  Deep  Water 


the  thermohaline  circulation — the  transforma- 
tion of  these  bottom  waters  to  lower  density,  and 
their  upwelling  and  eventual  return  to  the  high- 
latitude  cooling  zones — are  less  well  understood. 
An  upwelling  of  deep  and  bottom  waters  is  be- 
lieved to  be  fed  by  the  continual  supply  of  new 
bottom  water:  Dense  new  waters  intrude  below 
older  waters  and  force  them  upwards.  The  bot- 
tom water  source  strength  of  20  to  30  million 
cubic  meters  per  second  translates  into  a  globally 
averaged  upwelling  rate  at  mid-ocean  depth  of 
about  3  meters  per  year.  This  upwelling  has  both 
dynamical  and  thermodynamical  implications. 

To  maintain  a  steady-state  temperature  distri- 
bution in  the  face  of  this  upwelling  of  cold  water, 
a  compensating  warming  is  required.  This  warm- 
ing may  be  accomplished  by  internal  mixing  of 
the  deep  ocean.  Models  exploring  the  thermody- 
namic  balance  between  the  downward  diffusion 
of  heat  associated  with  mixing  by  turbulent  ed- 
dies and  the  upwelling  of  cold  water  were  pub- 
lished by  Klaus  Wyrtki  (University  of  Hawaii) 
and  Walter  Munk  (Scripps  Institution  of  Ocean- 
ography) in  the  mid  1960s.  At  about  the  same 
time  Wyrtki's  and  Munk's  papers  appeared, 
Henry  Stommel,  considering  the  dynamical  ef- 
fects of  deep  upwelling,  proposed  the  existence  of 
abyssal  gyre  circulations  involving  poleward  deep 
flow  in  the  ocean  interiors  fed  by  a  series  of  west- 
ern boundary  currents.  These  boundary  flows 
ultimately  connect  to  the  high-latitude  bottom 
water  formation  sites.  Twenty  years  later  Frank 
Bryan  (National  Center  for  Atmospheric  Re- 
search) published  a  study  of  an  idealized,  three- 
dimensional  ocean  model  showing  a  direct  rela- 
tionship between  the  intensity  of  the  vertical 
mixing  and  the  strength  of  the  thermohaline 
overturning  circulation.  These  theoretical  ideas 
linking  diffusion,  upwelling,  and  the  deep  cur- 


rent systems  have  guided  research  on  abyssal 
circulation  for  the  past  three  decades. 

But  how  much  vertical  diffusion  is  there  in 
the  oceans,  and  what  processes  sustain  it? 
Munk's  application  of  his  model  to  data  from 
the  North  Pacific  Ocean  required  a  downward 
diffusive  heat  flux  about  1,000  times  larger  than 
that  caused  by  molecular  diffusion  (the  process 
whereby  differences  in  temperature  or  concentra- 
tion of  a  dissolved  substance  are  removed  by  the 
random  motion  of  molecules).  More  recent 
studies  concerning  vertical  diffusive  heat  fluxes 
in  semi-enclosed  basins  also  required  downward 
diffusive  heat  fluxes  thousands  of  times  greater 
than  those  due  directly  to  molecular  diffusion. 
All  of  the  researchers  involved  invoke  turbulent 
mixing  as  the  mechanism  supporting  these  large 
diffusive  heat  fluxes. 

Ocean  turbulence  is  the  focus  of  a  subgroup  of 
physical  oceanographers  specializing  in  micro- 
structure,  that  is,  temperature  and  velocity  struc- 
tures occurring  at  spatial  scales  directly  influ- 
enced by  seawater's  molecular  viscosity  and 
thermal  diffusivity — typically  around  one  centi- 
meter. These  scientists  have  extensively  sampled 
the  upper  ocean  in  recent  years.  Apart  from  the 
surface  layer  (which  is  actively  mixed  by  wind 
and  waves),  the  shallow  ocean  margins,  and 
highly  sheared  flows  like  the  equatorial  undercur- 
rent, the  microstructure  data  suggest  turbulent 
diffusive  fluxes  some  ten  times  smaller  than  the 
studies  mentioned  above.  This  seeming  discrep- 
ancy caused  some  to  question  the  models  used  to 
deduce  the  intensity  of  vertical  diffusion  from 
microstructure  data,  and  whether  sufficient  data 
had  been  gathered  to  adequately  describe  ocean 
microstructure.  Relatively  weak  mixing  in  the 
upper  ocean  away  from  boundaries  was,  however, 
recently  confirmed  by  a  nontoxic  chemical  tracer 
release  experiment  in  the  Northeast  Atlantic  led 
by  Jim  Led  well. 

The  apparent  contradiction  between  micro- 
structure-based  and  indirectly  determined  esti- 
mates of  vertical  diffusion  might  actually  reflect  a 
real  difference  with  depth  in  the  ocean.  Part  of 
the  problem  is  that  the  indirect  estimates  of  verti- 
cal mixing  have  been  derived  tor  the  deep  ocean, 
while  the  bulk  of  the  microstructure  observations 
are  from  the  top  1  kilometer  of  the  ocean.  Ray 
Schmitt,  Kurt  Polzin,  and  I  have  recently  ad- 
dressed this  issue  with  a  series  of  cruises  on 
which  we  acquired  full-ocean-depth  profiles  of 
temperature  and  velocity  microstructure.  We  find 
evidence  of  enhanced  turbulent  mixing  in  the 
deep  ocean  near  the  bottom,  particularly  in  re- 
gions where  the  bottom  is  rough.  The  zone  of 
enhanced  mixing  extends  upward  to  several  hun- 
dred meters  above  the  bottom,  a  span  much 
greater  than  that  of  the  traditional  bottom 


34  «'  FALL/WINTER  1996 


boundary  layer,  a  roughly  10-meter- 
thick,  vertically  homogenized  layer 
that  is  maintained  by  bottom-gen- 
erated turbulence.  Our  data  also 
show  strong  internal  waves  at  these 
sites,  and  we  believe  the  enhanced 
mixing  is  sustained  by  the  breaking 
of  these  internal  waves,  which  are 
both  generated  at  and  reflected 
from  the  rough  bottom. 

These  observations  also  docu- 
ment striking  horizontal  patterns 
in  the  turbulent  mixing  at  depth. 
Our  current  study  (a  joint  micro- 
structure-tracer  experiment  in 
collaboration  with  Jim  Ledwell) 
is  now  underway  in  the  Brazil 
Basin,  the  region  where  Nelson 
Hogg  and  colleagues  inferred 
significant  vertical  diffusion  from 
a  heat  budget  for  the  bottom  wa- 
ters. In  the  interior  of  the  basin 
where  the  bottom  is  smooth,  the 
microstructure  data  imply  turbu- 
lent fluxes  less  than  a  tenth  of 
Hogg  and  colleagues'  basin-aver- 
aged value.  In  contrast,  above  the 
rough  flanks  of  the  Mid-Atlantic 
Ridge  in  the  eastern  third  of  the 
basin,  we  deduce  turbulent  fluxes 
greater  than  their  figure. 

We  find  that  the  horizontally 
averaged  turbulent  heat  flux  for 
our  study  region,  based  on  the 
microstructure  data  now  in  hand, 
is  in  near  accord  with  that  derived 
from  the  bottom  water  heat  bud- 
get. Our  results  suggest  that  vertical 
diffusion  in  the  deep  ocean  is 
dominated  by  turbulent  mixing 
near  rough  bathymetric  structures, 
a  refinement  of  Munk's  hypothesis  that  it  occurs 
generally  near  the  bottom.  Greater  average  tur- 
bulent fluxes  may  be  achieved  at  depth  than  in 
the  upper  ocean  because  a  larger  fraction  of  the 
deep  ocean  is  in  close  proximity  to  the  bottom. 
Spatially  variable  mixing  in  turn  implies  exist- 
ence of  horizontal  circulations  to  distribute 
modified  waters  from  these  mixing  zones 
throughout  deep  basins.  Moreover,  given  the 
dynamical  links  between  mixing,  upwelling,  and 
circulation,  our  findings  hint  that  the  deep  gyres 
predicted  by  Stommel  might  be  highly  distorted 
in  the  real  ocean. 

The  scientific  community  is  just  beginning  to 
document  the  intensity  and  patterns  of  mixing  in 
the  ocean  abyss.  It  is  not  surprising  that  mixing  in 
ocean  climate  models  has  so  far  been  generally 
taken  as  spatially  uniform.  Much  work  remains 


,ent  mixing  in  surface  layer  driven  by  win, 
internal  wavebreaking  supports  fhe  deeper  turbulent  mixing 


energy  propagates  away  from  peaks  of 
bathymetry  as  narrow  beams  of 
internal  waves 

•  possibly  larger  amplitudes  and 
more  irregular  motion  where 
'—ams  intersect 


little  mixing  in 
ocean  interior 


tense  mixing  over 
lugh  bathymetry 


to  be  done,  both  observational  and  theoretical,  to 
fully  understand  the  role  of  turbulent  mixing  in 
the  ocean's  thermohaline  circulation. 

The  research  discussed  in  this  article  was  supported  by  the 
National  Science  Foundation  Initial  development  of  the 
High  Resolution  Profiler  was  supported  by  the  Department 
of  Defense  and  the  Office  of  Naval  Research. 

Attraction  to  the  sea  and  ocean  science  began  for  lohn  Toole 
with  a  keen  interest  in  sailing.  He  maintains  an  eclectic 
research  program  at  WHO/  that  includes  study  of  basin-scale 
circulations  and  the  processes  of  ocean  mixing.  Developing 
understanding  of  the  cold-to-warm  limb  of  the  thermohaline 
circulation  represents  a  synthesis  of  research  supported  by 
grants  from  the  National  Science  Foundation  and  the  Office 
of  Naval  Research.  With  WHO/  colleagues  and  his  wife  (and 
chief  foredeck  crew),  he  also  continues  to  campaign  sailing 
race  courses  through  the  summer,  as  research  cruises  and 
meetings  permit. 


A  schematic  draw- 
ing of  turbulent 
processes  at  work 
in  the  ocean. 


Computer  Modelers 

Simulate  Real  and  Potential  Climate, 

Work  Toward  Prediction 

Combining  Equations  and  Data  Pushes  Computers'  Limits 


Rui  Xin  Huang 

Associate  Scientist,  Physical  Oceanography  Department 

Jiayan  Yang 

Assistant  Scientist,  Physical  Oceanography  Department 

Although  weather  forecasting  is  accepted  by 
the  public  as  part  of  daily  life,  oceanic 
forecasting  is  not  yet  so  advanced.  There 
are,  however,  successful  examples  of  oceanic 
forecasting — one  is  the  newly  developed  skill  to 
predict  El  Nino/Southern  Oscillation  (ENSO) 
events,  largely  due  to  improvements  in  ocean 
modeling  (see  following  article). 

In  1982  and  1983,  Eastern  Australia  and  Indo- 
nesia experienced  the  century's  worst  drought, 
which  led  to  devastation  in  agricultural  regions 
and  rain  forests,  and  even  to  loss  of  hundreds  of 


Authors  Rui  Xin 
Huang,  right,  and 

Jiayan  Yang  col- 
laborate on  ocean 
process  study  and 
climate  prediction 
models. 


lives.  These  were  just  some  regional  impacts  of 
what  is  now  known  as  the  most  severe  ENSO 
event  on  record,  an  anomalous  warming  of  sea 
surface  temperature  in  the  eastern  Tropical  Pacific 
Ocean  that  occurs  once  every  three  to  seven  years. 
The  change  in  oceanic  conditions  associated  with 
ENSO  is  usually  accompanied  by  atmospheric 
shifts,  and  together  these  phenomena  lead  to 
droughts  in  some  parts  of  the  world  and  flooding 
in  others.  It  is  estimated  that  total  worldwide 
damage  caused  by  the  1982-83  ENSO  was  more 
than  $10  billion. 

Another  climate  variation,  the  North  Atlantic 


Oscillation  (NAO),  a  shift  of  atmospheric  pres- 
sure fields  between  Iceland  (65°  N)  and  the 
Azores  (40°  N)  on  decadal  time  scales,  is  known 
to  change  weather  conditions  in  Europe  and 
North  America  (see  Box  on  page  13).  ENSO  and 
NAO  are  just  two  examples  of  how  natural  cli- 
matic fluctuations  can  dramatically  affect  the 
world  economy  and  our  daily  lives.  Earth's  cli- 
mate changes  ceaselessly,  and  it  will  surely  con- 
tinue to  evolve,  possibly  in  a  more  complicated 
manner  due  to  increasing  atmospheric  concentra- 
tions of  greenhouse  gases.  Thus  there  are  pressing 
reasons  to  improve  our  understanding  of  severe 
climate  variations,  such  as  ENSO  and  NAO 
events,  and  even  to  predict  them  before  they  oc- 
cur so  that  the  public  can  be  informed  and  policy 
makers  can  prepare  for  possible  natural  disasters. 
Because  the  future  is  unobservable,  we  must  rely 
on  numerical  models  for  such  forecasting. 

Geologic  studies  of  Earth's  history  show  that 
the  world  ocean  has  changed  profoundly  over 
time.  Modern  observations  indicate  that  there 
have  been  noticeable  changes  in  world  ocean 
circulation  even  during  recent  decades.  As  our 
knowledge  advances,  so  does  our  understanding 
of  the  ocean's  importance  in  the  climate  system. 
Driven  by  wind  stress  as  well  as  heat  and  freshwa- 
ter fluxes,  oceanic  currents  redistribute  heat 
across  the  globe  and  regulate  our  climate.  The 
ocean's  enormous  capacity  to  store  heat  also 
buffers  climate  changes.  Since  the  ocean  and  the 
atmosphere  exchange  momentum,  heat,  and 
fresh  water  across  their  interfaces,  variation  in 
one  fluid  system  can  lead  to  changes  in  the  other, 
often  in  a  chain  reaction  that  amplifies  initially 
small  deviations.  Many  climate  phenomena,  such 
as  ENSO,  result  from  such  interactions,  which 
can  occur  over  a  wide  spectrum  of  time  scales.  So, 
even  though  the  weather  can  be  forecast  for  a 
week  without  considering  oceanic  circulation, 
climate  on  time  scales  longer  than  a  month  must 
include  the  ocean. 

Temporal  evolution  and  spatial  variation  of 
the  ocean  are  constrained  by  physical  laws  and 
such  external  forcing  fields  as  wind  stresses  and 
surface  buoyancy  fluxes.  The  essence  of  climate 


FALL/WINTER  1996 


Satellite 


Land-based 

Forecasting 

Facility 


modeling  is  to  integrate  the  dynamic  equations 
for  these  climate  components  forward  in  time, 
starting  with  conditions  that  are  often  based  on 
actual  observations.  The  basic  idea  of  computer 
simulation  is  to  organize  physical  equations  into 
a  net  of  grids  arranged  to  cover  a  spatial  domain, 
such  as  the  tropical  Pacific  Ocean  for  ENSO  pre- 
dictions, or  the  global  oceans  for  carbon  cycle 
assessments,  and  then  predict  the  climatic  state  at 
each  grid  in  the  future  based  on  its  initial  condi- 
tion and  its  subsequent  interactions  with  sur- 
rounding grids  as  the  conditions  in  each  change. 
Though  the  evolution  of  a  climate  event  is 
unrepeatable  and  beyond  our  control,  computer 
simulations  can  be  repeated  many  times  by  vary- 
ing the  mathematic  representation  of  conditions 
and  forces  at  work  in  each  grid.  Thus,  numerical 
models  are  very  powerful  tools  for  testing  scien- 
tific hypotheses  and  for  examining  important 
climate  processes. 

One  good  example  is  a  study  by  Frank  Bryan, 
who  conducted  numerical  experiments  in  the 
early  1980s  while  a  graduate  student  at  the  Geo- 
physical Fluid  Dynamics  Laboratory  in  Princeton, 
N(.  By  running  an  idealized  model  for  the  Atlan- 
tic Ocean,  he  showed  that  deep  water  formation 
in  the  North  Atlantic  could  be  shut  off  by  a 
strong  salinity  perturbation  in  the  subpolar  ba- 
sin. If  such  changes  were  to  take  place,  the  North 
Atlantic's  poleward  heat  flux  would  be  substan- 
tially reduced,  and  the  European  climate  would 
be  remarkably  less  mild.  His  modeling  results  are 


consistent  with  a  paleoclimatic  record  that  shows 
deep  water  formation  was  interrupted  about 
12,000  years  ago  in  an  event  known  as  the 
Younger  Dryas  when  melted  glacial  water  flooded 
the  subpolar  North  Atlantic  Ocean,  resulting  in 
Northern  Hemisphere  cooling  and  even  reduc- 
tion in  the  deglaciation  process. 

The  accuracy  of  a  numerical  model  depends 
on  how  well  and  how  realistically  it  approxi- 
mates boundary  conditions,  external  forcing,  and 
key  processes  that  govern  the  real  climate  system. 
The  ideal  model  would  use  a  very  fine  spatial  grid 
and  a  small  integration  time  step  to  resolve  spa- 
tial and  temporal  structures  of  all  important 
physical  processes  in  the  system.  For  instance, 
mesoscale  eddies,  whirling  parcels  of  fluid  typi- 
cally 40  to  200  kilometers  wide  in  the  subtropics, 
play  significant  roles  in  redistributing  momen- 
tum, heat,  salts,  and  other  dissolved  matter  in  the 
ocean.  Excluding  such  processes  will  definitely 
lead  to  inaccurate  model  representations  of  the 
oceans.  However,  most  of  the  current  climate 
models  use  resolutions  (spacing  between  two 
adjacent  grids)  on  the  order  of  100  kilometers  or 
coarser,  too  coarse  to  explicitly  resolve  the  spatial 
structures  of  mesoscale  eddies. 

Increasing  the  horizontal  and  vertical  resolu- 
tion of  the  models  10  times  requires  increasing 
the  number  of  grid  points  1,000  times.  In  addi- 
tion, the  allowable  time  step  for  maintaining  the 
numerical  calculation's  stability  will  be  at  least  10 
times  smaller  (the  time  step  should  be  smaller 


The  authors  envi- 
sion a  forecasting 
center  that  would 
receive  near-real- 
time data  from  a 
variety  of  sources 
by  satellite  trans- 
mission for  con- 
stant updating  of 
ocean  climate 
models.  The  data 
sources  might  in- 
clude those  shown 
here.  The  surface 
mooring  transmits 
meteorological 
data  as  well  as  in- 
formation from  the 
string  of  instru- 
ments below  it. 
Slocum,  described 
on  page  6,  surfaces 
once  a  week  or  so 
at  the  top  of  its  tra- 
jectory to  transmit 
temperature  and 
salinity  records 
from  ocean  depths, 
and  SeaSoar  data 
on  a  variety  of  up- 
per ocean  charac- 
teristics is  beamed 
from  the  ship.  Data 
collected  by  satel- 
lites would  include 
sea  surface  eleva- 
tion, wind  stress, 
temperature,  and 
perhaps  salinity. 


OCEANUS 


than  the  grid  size  divided  by  the  maximum  veloc- 
ity). Thus,  we  need  a  computer  that  is  10,000 
times  faster  and  has  1,000  times  more  memory. 
For  such  high  resolution,  even  the  fastest  com- 
puters, such  as  the  128-processor  CMS  computer, 
which  reaches  a  speed  of  3  gigaflops  (3  billion 
float  point  operations  per  second),  cannot  yet 
accommodate  a  long-term  global  simulation. 
Thus  one  key  aspect  of  climate  modeling  involves 
how  to  accurately  represent  important  processes 
that  cannot  be  explicitly  resolved  due  to  model 
resolution  limits.  This  is  called  the  subgrid-scale 
parametrization. 

Another  important  process  that  requires  care  is 
the  direction  along  which  mixing  occurs.  In 
many  ocean  general  circulation  models 
(OGCMs),  especially  those  formulated  in  fixed 
spatial  grids,  mixing  in  a  model  grid  is  repre- 
sented by  some  form  of  averaging  its  properties 
with  those  in  surrounding  grids.  In  the  real 
world,  mixing  is  most  likely  to  occur  along  con- 
stant density  surfaces  so  that  the  mixing  processes 
do  not  work  against  the  buoyancy.  Recent  work 
by  lames  McWilliams  and  Peter  Gent  and  their 
colleagues  at  the  National  Center  for  Atmo- 
spheric Research  has  significantly  improved  the 
performance  of  ocean  models  that  use  spatially 
fixed  grids. 

Rapidly  developing  computer  technology  allows 
climate  modelers  to  work  at  ever  finer  resolution 
as  they  aim  explicitly  to  model  the  structures  and 
temporal  evolution  of  eddies.  For  instance,  Albert 
Semtner  and  his  colleagues  at  the  Naval  Postgradu- 
ate School  in  Monterey,  California,  have  used  an 
OGCM  with  a  horizontal  resolution  of  about  10 
kilometers  to  study  circulation  in  the  Arctic  Ocean 
and  the  Greenland  and  Norwegian  Seas.  Their 
eddy-resolving  model  captures  many  observed 
frontal  structures,  such  as  sharply  defined  features 
often  associated  with  strong  jets,  that  coarse  reso- 
lution models  have  not  been  able  to  capture. 

There  are  two  major  categories  in  ocean  cli- 
mate modeling,  process  studies  and  climate  pre- 
dictions. Process  studies  aim  to  understand  im- 
portant processes  that  operate  the  real  climate 
system,  to  identify  mechanisms  that  give  rise  to 
climate  variations,  or  to  explain  particular  pat- 
terns observed  in  the  real  world.  Such  studies 
often  involve  a  hierarchy  of  models,  from  simple 
ones  to  full,  three-dimensional  models.  Climate 
predictions,  like  numerical  weather  forecasts, 
attempt  to  determine  future  climate  states  based 
on  available  knowledge  about  how  the  climate 
system  evolves  in  time.  Climate  predictions  al- 
ways benefit  from  progress  in  process  studies.  For 
instance,  tremendous  advances  in  process  studies 
of  tropical  air-sea  interactions  in  the  1980s  led  to 
great  success  in  predicting  ENSO  events  a  year 
ahead  of  time.  Simple  models  play  very  active 


roles  in  process  studies.  For  example,  the  late 
Henry  Stommel  used  a  very  simple  two-box 
model  to  elucidate  how  the  distinct  difference 
between  air-sea  fluxes  of  heat  and  fresh  water  lead 
to  multiple  equilibrium  states  in  the  oceanic 
thermohaline  circulation.  This  seminal  work  laid 
the  foundation  for  our  understanding  of  the  sta- 
bility and  variability  of  the  Atlantic  overturning 
circulation.  More  comprehensive  models,  which 
can  resolve  mesoscale  eddies  and  include  ocean- 
atmosphere  interactions,  have  been  used  to  verify 
Stommel's  work  and  to  gain  further  understand- 
ing of  the  oceanic  thermohaline  circulation. 

Climate  prediction  models  must  prove  they 
can  describe  observed  climate  evolutions  in  the 
past  before  they  can  be  trusted  for  future  predic- 
tions. Therefore,  modelers  need  observations  to 
calibrate  and  verify  their  models.  Unlike  atmo- 
spheric data  sets,  oceanic  observations  have  rela- 
tively short  records  and  sparse  coverage  in  space. 
New  observing  technologies,  like  satellite  remote 
sensing,  acoustic  tomography,  and  the  long-lived 
floats  that  Ray  Schmitt  describes  on  page  6,  will 
certainly  expand  the  observing  capacity  and  sup- 
port climate  modeling.  A  global  observational 
network  is  likely  to  be  a  combination  of  satellite- 
borne  instruments  (which  can  measure  sea  sur- 
face elevation,  wind  stress,  temperature,  and  per- 
haps salinity);  automatic  instruments,  such  as 
buoys  and  data-transmitting  floats;  and  tradi- 
tional shipboard  instruments.  Data  collected  by 
these  instruments  will  be  sent  to  a  land-based 
forecasting  center,  where  the  most  powerful 
supercomputers  will  merge  current  oceanographic 
data  and  forecast  oceanic  conditions  for  the  near 
future.  With  the  rapid  advance  of  computer  tech- 
nology and  our  understanding  of  ocean  physics, 
oceanic  forecasting  will  eventually  become  a 
reality— perhaps  early  in  the  21st  century,  marine 
and  climate  forecasting  will  become  routine. 
The  research  discussed  in  this  article  was  supported  by  the 
National  Science  Foundation  and  the  National  Oceanic 
and  Atmospheric  Amdinistration's  Global  Climate 
Change  Program. 

Rui  Xin  Huang's  primary  research  interest  is  large-scale  oceanic 
circulation,  including  wind-driven  g}>res  and  thermohaline 
emulation.  When  he  was  in  high  school,  his  dream  was  to 
become  an  inventor  like  Edison.  Through  a  long  and  winding 
road,  he  came  to  Woods  Hole,  and  found  oceanography  an 
exciting  field.  He  also  likes  swimming,  gardening,  and,  above 
all,  pilules  and  games. 

During  his  graduate  student  and  postdoctoral  years,  liayan  Yang 
was  interested  mainly  in  tropical  air-sea  interaction.  He  decided 
to  take  "a  short  break"  away  from  the  tropics  to  do  a  small, 
high-latitude  oceanography  project  when  he  was  a  postdoctoral 
fellow  at  the  University  of  California,  Los  Angeles.  He  has 
stayed  in  high-latitude  oceanography  eivr  since.  He  moved  from 
Los  Angeles  to  Neu*  England  to  get  a  bit  closer  to  (though  still 
far  way  from)  sea-ice  margins.  In  his  leisure  time,  he  likes 
hiking,  swimming,  biking,  and  karaoke. 


38  •  FALL/WINTER  1996 


The  El  Nino/Southern 
Oscillation  Phenomenon 

Seeking  Its"Trigger"  and  Working  Toward  Prediction 


Lewis  M.  Rothstein 

Associate  Professor,  Graduate  School  of  Oceanography, 
University  of  Rhode  Island 

Dake  Chen 

Research  Scientist,  Lamont-Doherty  Earth 
Observatory,  Columbia  University 

The  El  Nino/Southern  Oscillation  (ENSO) 
phenomenon,  an  eastward  shift  of  warm 
water  in  the  tropical  Pacific  and  associated 
effects  on  the  atmosphere,  is  at  the  heart  of  glo- 
bal interannual  climate  variability.  The  just  com- 
pleted, decade-long  Tropical  Ocean/Global  At- 
mosphere (TOGA)  program  was  dedicated  to 
understanding  and  working  toward  predicting 
ENSO  by  bringing  together  oceanographers  and 
atmospheric  scientists  in  a  coordinated  observa- 
tional and  numerical  modeling  research  pro- 
gram. TOGA  has  not  answered  all  the  questions: 
We  have  not  uncovered  the  physical  mechanisms 
of  the  elusive  ENSO  "trigger"  nor  have  our  best 
coupled  air/sea  numerical  models  been  as  suc- 
cessful in  predicting  the  rather  irregular  ENSO 
signal  of  the  1990s  as  uhey  were  in  predicting  the 
regular  events  of  the  1 980s  and  hindcasting  the 
events  of  the  late  1960s  through  the  1970s. 

Prediction  is  the  ultimate  goal  of  ENSO  re- 
search. It  is  also  the  ultimate  test  for  an  ENSO 
model  and  the  theory  underlying  the  model. 
During  the  last  decade,  a  number  of  forecast 
models  have  shown  predictive  skills  in  both  ret- 
rospective and  real  time  forecasting,  and  they  are 
now  being  used  for  routine  ENSO  prediction. 
Nevertheless,  the  skill  of  even  the  best  available 
models  is  far  from  perfect,  and  there  is  still  con- 
siderable room  for  improvement  in  modeling, 
observation,  and  forecasting  techniques. 

Factors  that  limit  the  current  skill  of  ENSO 
forecasts  include: 

•  an  inherent  limit  to  predictability  because  of 
the  chaotic  and  random  nature  of  the  natural 
system, 

•  model  flaws  such  as  oversimplified  physics, 

•  gaps  in  the  observing  system,  and 

•  flaws  in  the  way  the  data  is  used  (data  assimila- 
tion and  initialization). 


It  seems  likely  that  the  inherent  predictability 
limit  for  ENSO  is  years  rather  than  weeks  or 
months,  though  more  theoretical  study  is  needed 
in  this  area.  The  observing  system  is  improving, 
but  still  far  from  satisfactory.  Thus  a  challenge 
facing  the  modelers  is  to  improve  model  forecasts 
by  making  the  most  reasonable  and  efficient  use 
of  available  data. 

In  the  past  few  years  much  effort  has  been 
devoted  to  assimilating  various  observational 
data  into  die  initial  state  of  forecast  models.  The 
most  common  approach  is  to  improve  die  ini- 
tial ocean  conditions  by  assimilating  observa- 
tions of  sea  surface  temperature,  thermocline 
(region  of  rapid  temperature  decline)  deptJi,  or 
sea  level  into  an  ocean  model  prior  to  coupling 
it  with  an  atmosphere  model.  One  problem 


Time  series  of  ob- 
served (red)  and 
forecast  (blue)  El 
Nino  sea  surface 
temperature 
anomalies.  Fore- 
casts with  0,  6,  12, 
and  18  month 
leads  are  shown  in 
different  panels, 
and  the  observed 
anomalies  are  re- 
peated from  panel 
to  panel. 


13 
O 

c 

I 

2 
I 


•E 


I1"!"1!"1!"1!"1!1"!1"!1"!1"!"1!1"!"1!" 

0  Month  Lead 


I"'!'"!'"!"'!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"!'"! 


72       74       76       78       '80       '82       '84       '86       '88       '90       '92       '94       '96 

Year 


OCEANUS 


with  this  approach  is  that  no  attention  is  paid 
to  the  ocean-atmosphere  interaction  during 
initialization,  so  the  coupled  system  may  not  be 
well  balanced  initially  and  may  experience  a 
shock  when  the  forecast  starts.  A  new  initializa- 
tion/assimilation procedure  significantly  im- 
proves the  predictive  skill  of  one  of  our  most 
promising  coupled  models,  which  was  con- 
structed by  Mark  Cane  and  Steve  Zebiak 
(Lamont-Doherty  Earth  Observatory). 

In  the  new  methodology  the  model  is  initial- 
ized in  a  coupled  manner,  using  a  simple  data 
assimilation  scheme  in  which  the  coupled  model 
wind  stress  anomalies  are  "nudged"  toward  ob- 


120  E 


This  diagram  illus- 
trates weather 
pathways  in  the 
North  Pacific  sub- 
tropical/tropical 
upper  ocean  and 
the  main  horizon- 
tal gyres  and  me- 
ridional-vertical 
cells  of  the  region's 
ocean  circulation. 


overpredicted,  and  the  short  warm  episodes  in 
1993  and  late  1994  are  missed. 

Although  the  predictive  skill  of  this  model  is 
most  likely  limited  by  its  highly  reduced  physics, 
the  skill  of  more  sophisticated  coupled  ocean- 
atmosphere  general  circulation  models  presently 
does  not  exceed  that  of  the  model  described 
above,  at  least  in  terms  of  the  tropical  Pacific  sea 
surface  temperature.  In  order  to  predict  the  glo- 
bal impact  of  ENSO,  a  two-tiered  approach 
appears  to  be  reasonable:  A  physics-simplified, 
coupled  model  is  first  used  to  predict  tropical  sea 
surface  temperature  fields,  and  these  fields  are 
then  used  as  boundary  conditions  for  a  more- 
complete-physics,  global  atmospheric 
XModel..KurOShi0....  general  circulation  model  to  predict 

the  global  distribution  of  atmospheric 
disturbances.  Scientists  are  rigorously 
pursuing  this  kind  of  research. 

A  second  area  of  progress  concerns 
improved  understanding  of  the  cou- 
pling between  different  depths  and 
different  regions  of  the  ocean.  A 
popular  ENSO  paradigm  that  emerged 
in  the  late  1980s  was  based  on  the 
a  observed,  rather  regular  rythms  of 
f  ENSO  conditions  during  a  span  of  25 
=  years  before  1990.  The  "delayed  oscil- 
~  lator"  mechanism  emphasizes  east- 
ward-propagating equatorial  wave 
processes,  *  westward-propagating  off- 
equatorial  signals,  and  their  asymmetric 
reflections  at  eastern  and  western  bound- 
aries respectively.  Despite  the  irregularities 
in  the  1990s  ENSO,  this  wave  propagation/ 
reflection  paradigm  is  still  compelling;  it  can 
accomodate  irregularities  in  the  ENSO  signal  by 
combining  the  tropical  signal  with  longer-term 
variability  in  the  subtropics. 

A  number  of  studies  have  sought  to  under- 
stand how  tropical  variability  is  linked  to  the  mid 
latitudes.  Ocean  circulation  may  provide  the 
links  via  several  different  pathways  that  are  sum- 
marized schematically  above.  These  are  not 
simple,  direct  north-south  flows;  the  existence  of 
vigorous  zonal  current  systems  complicate  the 
picture.  In  the  upper  layers  of  the  ocean,  up- 
welled  waters  along  the  equator  flow  into  the 
subtropics,  mainly  through  the  mid-latitude 
western  boundary  current  (the  Kuroshio).  There 
is  an  additional  interior  ocean  pathway,  through 
the  eastward  Subtropical  Countercurrent,  that 
more  directly  feeds  subtropical  sites  where  sur- 

*Flow  along  the  equator  tends  to  be  trapped  there.  The 
Coriolis  force,  due  to  the  earth's  rotation,  turns  water  that 
flows  south  back  to  the  north  and  water  that  flows  north 
back  to  the  south.  Because  of  this  trapping,  physical  ocean- 
ographers  recognize  the  equator  as  a  waveguide,  where  co- 
herent signals  or  waves  can  be  seen  to  propagate  east-west 
for  long  distances. 


Model  Eastward 

Subtropical 
Countercurrent 


Re-circulating 

Tropical 

Gyre 


300m 


servations.  The  new  procedure  improves  the 
model's  predictive  ability  as  measured  by  a  vari- 
ety of  statistical  scores.  It  also  eliminates  the  so- 
called  "spring  prediction  barrier,"  a  marked  drop 
of  skill  in  forecasts  that  try  to  predict  across  the 
boreal  spring,  lound  in  many  previous  ENSO 
forecast  systems.  The  success  of  the  new  initial- 
ization procedure  is  attributed  to  its  explicit 
consideration  of  ocean-atmosphere  coupling, 
and  the  associated  reduction  of  initialization 
shock  and  random  noise. 

As  an  example,  the  forecasts  made  by  the  im- 
proved model  are  compared  to  observations  in 
the  figure  on  page  39  in  terms  of  the  sea  surface 
temperature  anomaly  averaged  over  an  area  in 
the  eastern/central  equatorial  Pacific  (5°  S  to  5°  N 
and  90°  W  to  150°  W).  The  model  is  capable  of 
forecasting  ENSO  more  than  one  year  in  advance. 
The  large  warming  and  cooling  events  in  the 
1980s  are  particularly  well  predicted.  However, 
the  model  does  a  poorer  job  for  the  1970s  and 
1990s:  The  1976-77  event  is  largely 


40  «  FALL/WINTER  1996 


face  water  moves  deeper  into  the  ocean.  These 
interior  pathways  are  associated  with  a  recirculat- 
ing  tropical  gyre  in  and  just  helow  the  mixed 
layer  in  the  northeastern  tropics.  Below  the  mixed 
layer,  thermocline  water  from  the  suhtropics  to 
the  tropics  zigzags  almost  zonally  across  the 
basin,  succeeding  in  flowing  toward  the  equator 
only  along  zonally  narrow,  southward  flowing 
conduits.  The  low-latitude  western  boundary 
currents  serve  as  the  main  southward  circuit  for 
the  subducted  (water  moving  from  the  surface  to 
depth),  subtropical  thermocline  water. 

A  model  constructed  by  the  authors  also  indi- 
cates important  direct  flow  of  thermocline  water 
through  the  ocean  interior,  confined  to  the  far 
western  Pacific  (away  from  the  low-latitude  west- 
ern boundary  currents)  along  10°  N.  These  south- 
ward flowing  waters  are  then  swept  eastward  by 
the  North  Equatorial  Countercurrent,  finally 
penetrating  to  the  equator  in  the  central  and 
eastern  Pacific.  The  water  pathways  in  the  sub- 
tropical thermocline  essentially  reflect  the  surface 
gyre  circulation. 

Along  with  our  colleagues  Ronghua  Zhang 
(University  of  Rhode  Island)  and  Antonio  I. 
Busalacchi  (NASA  Goddard  Space  Flight  Center), 
we  have  examined  the  interannual  variability  of 
these  subtropical/tropical  pathways  and  found 
important  propagating  subsurface  ENSO  signa- 
tures in  the  subtropical  Pacific.  There  appears  to 
be  continual  movement  of  subsurface,  basin-scale 
anomalies  that  can  then  affect  sea  surface  tem- 
perature (SST)  anomalies,  especially  in  sensitive 
regions  where  the  thermocline  is  shallow.  These 
SST  anomalies  can  then  trigger  coupled  air/sea 
interactions.  A  clear  pattern  of  moving  anomalies 
is  less  obvious  at  the  sea  surface.  The  systematic 
subsurface  propagation  is  reminiscent  of  the 
delayed  oscillator:  eastward  along  the  equator, 
westward  off  the  equator  with  apparent  further 
propagation  along  the  eastern  and  western 
boundaries.  Off  the  equator,  subsurface  propaga- 
tion of  anomaly  patterns  initiates  an  SST 
anomaly  in  the  North  Equatorial  Countercurrent 
regions  of  the  western  Pacific,  which  then  intensi- 
fies and  moves  into  the  equatorial  waveguide, 
consistent  with  the  mean  water  pathways  found 
above.  We  speculate  that  this  could  be  a  mecha- 
nism for  initiating  coupled,  air-sea  interactions 
that  can  begin  to  evolve  as  an  ENSO  event.  The 
cycling  time  of  the  subsurface  anomaly  patterns 
may  determine  the  ENSO's  frequency.  We  look 
forward  to  continuing  our  investigations  to  so- 
lidify these  assertions. 

One  challenge  for  the  newly  established  Cli- 
mate Variability  (CLIVAR)  program  will  be  to 
uncover  the  ENSO  triggering  mechanism  and 
enable  intelligent  design  of  a  long-term  ocean 
and  atmosphere  monitoring  system.  CLIVAR  is 


MBI.  WHOI    LIBRARY 


the  oceanographic  and  atmospheric  scientific 
community's  new  program  of  climate  prediction. 
Its  focus  is  on  understanding  the  coupled  air/sea 
system's  variability  on  seasonal-to-interannual- 
to-interdecadal  time  scales  for  the  purpose  of 
determining  predictability,  and  then  prediction. 
Those  observations  would  then  feed  into  coupled 
air/sea  numerical  models  for  the  purpose  of  long 
lead  time  forecasting,  much  like  the  present-day 
weather  forcasting  systems.  However,  the 
interannual  ENSO  signal  does  not  exhibit  a 
simple  ryhthm;  there  are  clearly  influences  on 
longer  (decadal)  time  scales  that  need  to  be  con- 
sidered. There  are  clues  as  to  what  those  signals 
might  be  (for  example,  the  North  Atlantic  Oscil- 
lation— see  page  13),  but  we  are  still  in  the  early 
stages  of  identifying  these  signals. 

The  natural  system  is  not  easily  divided  ac- 
cording to  time  scales;  it  is  a  fully  nonlinear  sys- 
tem. If  we  are  to  understand  and  eventually  pre- 
dict global  interannual  variability,  we  must  not 
limit  ourselves  to  monitoring  the  air/sea  system 
over  a  few  interannual  cycles.  Permanent  moni- 
toring systems  are  needed.  It  is  the  primary 
charge  of  CLIVAR  to  help  design  such  a  monitor- 
ing system  while,  at  the  same  time,  supporting 
the  evolution  of  the  numerical  prediction  systems 
that  will  issue  the  forecasts. 

The  authors'  ENSO  research  is  supported  by  the  National 
Oceanic  and  Atmospheric  Administration,  the  TOGA 
Program  on  Seasonal  to  Interannual  Prediction,  and  the 
National  Aeronautics  and  Space  Administration 

Lew  Rothstein  started  his  career  as  a  physical  oceanographer  on 
the  beautiful  campus  of  the  University  of  Hawaii.  He  is  now  a 
professor  at  the  University  of  Rhode  Island  and  an  editor  of  the 
Journal  of  Geophysical  Research.  Dake  Chen  is  also  a  physi- 
cal oceanographer  by  training.  He  worked  with  Lew  on  various 
tropical  ocean  models  at  the  University  of  Rhode  Island  before 
he  joined  the  senior  staff  of  the  Lamont-Doherty  Earth  Observa- 
tory  last  summer.  Both  of  them  are  fond  of  building  computer 
models  of  the  ocean  and  atmosphere,  not  only  for  scientific 
research  but  also  for  fun. 


Servicing  an  Auton- 
omous Tempera- 
ture Line  Acquisi- 
tion System 
(ATLAS)  mooring 
of  the  Tropical 
Ocean  Global  At- 
mosphere (TOGA) 
program's  Tropical 
Atmosphere-Ocean 
(TAO)  Array  in  the 
Pacific  Ocean. 
ATLAS  moorings 
measure  surface 
winds,  air  tempera- 
ture, relative  hu- 
midity, sea  surface 
temperature,  and 
subsurface  temper- 
ature to  depths  of 
500  meters. 


OCEANUS 


1930 


Woods  Hole  Oceanographic  Institution 

Woods  Hole,  MA  02543 
508-457-2000