ArticlePDF Available

A Simple Ocean Data Assimilation Analysis of the Global Upper Ocean 1950 95. Part II: Results

Authors:

Abstract and Figures

The authors explore the accuracy of a comprehensive 46-year retrospective analysis of upper-ocean temper- ature, salinity, and currents. The Simple Ocean Data Assimilation (SODA) analysis is global, spanning the latitude range 628S-628N. The SODA analysis has been constructed using optimal interpolation data assimilation combining numerical model forecasts with temperature and salinity profiles (MBT, XBT, CTD, and station), sea surface temperature, and altimeter sea level. To determine the accuracy of the analysis, the authors present a series of comparisons to independent observations at interannual and longer timescales and examine the structure of well-known climate features such as the annual cycle, El Nino, and the Pacific-North American (PNA) anomaly pattern. A comparison to tide-gauge time series records shows that 25%-35% of the variance is explained by the analysis. Part of the variance that is not explained is due to unresolved mesoscale phenomena. Another part is due to errors in the rate of water mass formation and errors in salinity estimates. Comparisons are presented to altimeter sea level, WOCE global hydrographic sections, and to moored and surface drifter velocity. The results of these comparisons are quite encouraging. The differences are largest in the eddy production regions of the western boundary currents and the Antarctic Circumpolar Current. The differences are generally smaller in the Tropics, although the major equatorial currents are too broad and weak. The strongest basin-scale signal at interannual periods is associated with El Nino. Examination of the zero- lag correlation of global heat content shows the eastern and western tropical Pacific to be out of phase (correlation 20.4 to 20.6). The eastern Indian Ocean is in phase with the western Pacific and thus is out of phase with the eastern Pacific. The North Pacific has a weak positive correlation with the eastern equatorial Pacific. Correlations between eastern Pacific heat content and Atlantic heat content at interannual periods are modest. At longer decadal periods the PNA wind pattern leads to broad patterns of correlation in heat content variability. Increases in heat content in the central North Pacific are associated with decreases in heat content in the subtropical Pacific and increases in the western tropical Pacific. Atlantic heat content is positively correlated with the central North Pacific.
Content may be subject to copyright.
F
EBRUARY
2000 311CARTON ET AL.
q2000 American Meteorological Society
A Simple Ocean Data Assimilation Analysis of the Global Upper Ocean 1950–95.
Part II: Results
J
AMES
A. C
ARTON
,G
ENNADY
C
HEPURIN
,
AND
X
IANHE
C
AO
Department of Meteorology, University of Maryland at College Park, College Park, Maryland
(Manuscript received 6 February 1998, in final form 10 February 1999)
ABSTRACT
The authors explore the accuracy of a comprehensive 46-year retrospective analysis of upper-ocean temper-
ature, salinity, and currents. The Simple Ocean Data Assimilation (SODA) analysis is global, spanning the
latitude range 628S–628N. The SODA analysis has been constructed using optimal interpolation data assimilation
combining numerical model forecasts with temperature and salinity profiles (MBT, XBT, CTD, and station), sea
surface temperature, and altimeter sea level. To determine the accuracy of the analysis, the authors present a
series of comparisons to independent observations at interannual and longer timescales and examine the structure
of well-known climate features such as the annual cycle, El Nin˜o, and the Pacific–North American (PNA)
anomaly pattern.
A comparison to tide-gauge time series records shows that 25%–35% of the variance is explained by the
analysis. Part of the variance that is not explained is due to unresolved mesoscale phenomena. Another part is
due to errors in the rate of water mass formation and errors in salinity estimates. Comparisons are presented to
altimeter sea level, WOCE global hydrographic sections, and to moored and surface drifter velocity. The results
of these comparisons are quite encouraging. The differences are largest in the eddy production regions of the
western boundary currents and the Antarctic Circumpolar Current. The differences are generally smaller in the
Tropics, although the major equatorial currents are too broad and weak.
The strongest basin-scale signal at interannual periods is associated with El Nin˜o. Examination of the zero-
lag correlation of global heat content shows the eastern and western tropical Pacific to be out of phase (correlation
20.4 to 20.6). The eastern Indian Ocean is in phase with the western Pacific and thus is out of phase with the
eastern Pacific. The North Pacific has a weak positive correlation with the eastern equatorial Pacific. Correlations
between eastern Pacific heat content and Atlantic heat content at interannual periods are modest. At longer
decadal periods the PNA wind pattern leads to broad patterns of correlation in heat content variability. Increases
in heat content in the central North Pacific are associated with decreases in heat content in the subtropical Pacific
and increases in the western tropical Pacific. Atlantic heat content is positively correlated with the central North
Pacific.
1. Introduction
Recently we have reported a new retrospective anal-
ysis of upper-ocean temperature, salinity, sea level, and
currents for the global ocean, 1950–1995 (Carton et al.
2000). The result of that research has been to provide
a synthesis of the historical oceanographic data record
in the form of a retrospective analysis. Here we describe
the results of an exploration of the accuracy of this
analysis.
The Simple Ocean Data Assimilation (SODA) algo-
rithm used to construct our retrospective analysis con-
sists of a numerical forecast model and an update pro-
cedure to provide corrections to the forecast. The update
Corresponding author address: Dr. James A. Carton, Department
of Meteorology, University of Maryland at College Park, 2417 Com-
puter and Space Science Building, College Park, MD 20742-2425.
E-mail: carton@metosrv2.umd.edu
procedure is based on a data assimilation algorithm
widely applied in atmospheric numerical weather pre-
diction called optimal interpolation (see Daley 1991).
In optimal interpolation the differences between ob-
served variables such as temperature, salinity, and sea
level and model forecasts of the same variables are used
to update the forecast. The interpolation coefficients,
otherwise known as gain matrices, are determined so as
to minimize the mean square error of the analysis. Our
implementation of this algorithm differs from the usual
implementation to the extent that we include spatial de-
pendence of the assumed error statistics and because of
our assumptions about bias in the model forecast.
SODA differs from the more sophisticated but much
more computationally intensive Kalman filter to the ex-
tent that we do not have prediction equations for the
temporal and spatial evolution of the error statistics.
Rather, the error statistics are determined a priori based
on a statistical analysis of errors from a preliminary
analysis. Our approach also bears similarity to 3D var-
312 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
T
ABLE
1. Ocean analysis experiments presented in the text. Each
experiment covers the period 1950–96. Further description of these
experiments is provided in Carton et al. (2000).
Experiment Description
Control analysis Basic analysis with detrended winds
1 Basic analysis except assuming significant fore-
cast error bias
2 Simulation with no subsurface updating
3 Basic analysis with climatological monthly winds
4 Basic analysis with complete winds
5 Basic analysis except with salinity updating with
observed salinity, but without T/Serror covari-
ance
6 Basic analysis except without T/Serror covari-
ance
7 Basic analysis except replacing model with cli-
mate temperature
8 Basic analysis except that TOPEX/Poseidon al-
timeter sea level is excluded from updating
procedure
T
ABLE
2. Statistical comparison of control analysis and tide guage sea level for selected stations. Record length is given, along with overall
correlation (Cor1) and correlation of 5-yr low-pass filtered records (Cor2). Christmas Island record comes in two parts, each of which has
been detrended. The comparison is with the combined record. Correlations marked with an asterisk exceed the 95% test of significance.
Name Location Years Cor1 Cor2
Atlantic
San Juan (Puerto Rico)
Tenerife (Canary Is.)
Bermuda, Is.
La Coruna (Spain)
188279N, 668059W
288299N, 168149W
328229N, 648429W
438229N, 88249W
26
38
40
27
0.50*
20.02
0.46*
0.28
0.34
20.21
0.72*
0.31
Pacific
Rikitea Is.
Noumea II
Pago Pago Is.
Christmas Is. (two parts)
Pahnpei-b Is.
Majuro-b Is.
Kwajalein Is.
Yap Is.
Guam Is.
23889S, 1348579W
228189S, 1668269E
148179S, 1708419W
18599N, 1578299W
68599N, 1588149E
7869N, 1718229E
88449N, 1678449E
98319N, 138889E
138269N, 1448399E
20
22
40
14
17
17
45
21
41
0.15
0.52*
0.34*
0.86*
0.86*
0.79*
0.73*
0.66*
0.70*
0.43
0.77*
0.56*
0.88*
0.71*
0.81*
0.74*
0.91*
0.45*
Johnston Is.
Wake Is.
Hilo, Hawaii
Honolulu, Hawaii
French Frigate Shoals
Midway Is.
Sitka, Alaska
168459N, 1698319W
198179N, 1668379E
198449N, 155849W
218189N, 1578529W
238529N, 1668179W
288139N, 1778229W
57839N, 1358209W
41
38
45
45
17
35
35
0.66*
0.54*
0.44*
0.67*
0.44
0.07
0.07
0.40*
0.67*
0.35*
0.70*
0.77*
0.44*
0.44*
iational assimilation methods when the mean square er-
ror is minimized such as that described in Ji et al. (1995).
Comparison of the results shows correlations of heat
content exceeding 80% in the tropical Pacific (Chepurin
and Carton 1999). Bennett (1990) and Wunsch (1996)
provide comprehensive discussions of the alternatives
in formulating updating algorithms. Malonote-Rizzoli
(1996) reviews many current implementations.
The general circulation ocean model on which our
analysis is based uses the full Geophysical Fluid Dy-
namics Laboratory Modular Ocean Model 2.b primitive
equation code, with conventional choices for mixing,
etc. The domain of this analysis is global, extending
from 628Sto628N. The model horizontal resolution is
2.5830.58in the Tropics, expanding to a uniform 2.58
31.58resolution at midlatitude. No attempt is made to
resolve midlatitude eddy processes. At the polar bound-
aries the temperature and salinity fields are relaxed to
climatology. We make no attempt to model cryospheric
or deep-water formation processes explicitly. A weak
5-yr relaxation of the global temperature and salinity
fields is included in order to reduce forecast bias indeep-
water masses. Bottom topography is included. The mod-
el has 20 levels in the vertical, with 15-m resolution in
the upper 150 m. In this model sea level is obtained
through diagnostic calculation from the mass and mo-
mentum fields. Winds are provided by an analysis of
historical shipboard measurements by da Silva et al.
(1994) prior to 1993 and by the National Centers for
Environmental Prediction after that date.
The main datasets to constrain the model forecast are
the hydrographic data contained in the World Ocean
Atlas 1994 (WOA-94; Levitus et al. 1994), additional
hydrography obtained from a variety of sources, sea
surface temperature (Reynolds and Smith 1994), and
altimetry from the Geosat, ERS-1, and TOPEX/Posei-
don satellites. The hydrographic dataset exceeds 50 000
profiles per year for much of our 46-yr period of interest.
This dataset has been collected using several different
instruments. In the interval 1950–69, most of the tem-
perature measurements were made with the mechanical
bathythermograph. This subset reached a maximum of
70 000 measurements per year in 1968, but is limited
to sampling temperature in the upper 2–300 m. After
1969 the mechanical bathythermograph was largely re-
F
EBRUARY
2000 313CARTON ET AL.
F
IG
. 1. Observed and control analysis sea level time series at Kwa-
jalein Island (98N, 1688E). The seasonal cycle and a linear trend has
been removed from both records.
F
IG
. 2. Five-year low-pass filtered observed and control analysis
sea level at three locations: San Juan, Puerto Rico (188N, 668W),
Honolulu (218N, 1588W), and Kwajalein Island (98N, 1688E). A linear
trend has been removed from all records. Units are centimeters.
placed by the expendable bathythermograph. Two other
small subsets, conductivity–temperature–depth and bot-
tle measurements, are important because they provide
salinity as well as temperature and because they extend
more deeply into the water column.
The altimeter sea level used here is based on the
NASA Pathfinder Project version 2.1 to which we have
added all the standard corrections for geophysical ef-
fects and then averaged alongtrack into 18bins. No other
interpolation was carried out. The altimeter dataset be-
gins November 1986 with the Geosat Exact Repeat Mis-
sion. No altimeter data is available from the end of
Geosat in the fall of 1989 until the beginning of ERS-1
in spring 1992.
In addition to the basic analysis, which we refer to
as the control analysis, a series of analysis experiments
(see Table 1) has been carried out to determine some
of the properties and sensitivities of the analysis. The
control analysis and analysis experiments begin January
1950 and continue through December 1995. The anal-
ysis fields for each experiment consist of 552 monthly
averages of temperature, salinity, and the horizontal
components of velocity at 20 levels. Comparison of the
experiments reveals important sensitivities of the anal-
ysis to changes in parameters, boundary conditions, etc.
Some review of these results is provided in this paper.
A more extensive discussion is provided in Carton et
al. (2000).
Here our examination of analysis error focuses on
comparison to independent observations on interannual
and longer timescales. We limit our comparison to ex-
amination of correlations and root-mean-square (rms)
differences and a more detailed examination of one ex-
ample for each dataset. Section 2 examines the temporal
variability of the analysis by comparison to tide gauge
sea level time series from the Permanent Service for
Mean Sea Level. In Section 3 our examination focuses
on the spatial structure of analysis error. Datasets with
good spatial coverage include altimeter sea level, global
hydrographic sections, and surface drifter velocity.
2. Time series comparisons
In this section we present a comparison to island tide
gauge records. Comparison to temperature and salinity
at Bermuda is provided in Carton et al. (2000). Together
these datasets allow a detailed examination of the ac-
curacy of the analysis at interannual to decadal periods.
The behavior of the analysis at short seasonal periods
is also discussed in Carton et al. (2000).
Tide gauge comparison
Approximately 1800 tide gauge records are available
from the Permanent Service for Mean Sea Level or the
314 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 3. Sea level error estimated using three years of TOPEX/Poseidon altimetry. (a) Root-
mean-square difference between altimeter and expt 8 analysis sea level in which altimeter sea
level observations have been excluded from the analysis. (b) Rms difference between altimeter
and control analysis sea level in which temperature and salinity are constrained by altimeter
observations. Contour intervals are 2 and 1 cm.
Tropical Ocean Global Atmosphere sea level archives.
However, many gauges are in locations that are un-
suitable for observing the large-scale circulation, while
others have records that are too short for our purposes.
After examining the datasets we have identified 20 sta-
tions, mainly from islands in the North Atlantic and
Pacific (in the latitude band 238S–578N), with records
that each exceed 17 years in length and do not seem
excessively gappy or otherwise contaminated.
At each of the station locations we have carried out
a comparison between annually averaged observations
and control analysis sea-level time series. The compar-
ison is summarized in Table 2. Correlations between
observed and control analysis sea level is presented as
Cor1. The lowest correlations are for continental sta-
tions such as La Coruna, Spain, and for some of the
subtropical and midlatitude islands such as the Canary
Islands. The agreement at islands in the tropical Pacific
such as Kwajalein (98N, 1678E: Fig. 1) is generally ex-
cellent. A succession of high and low sea level events
in the record at Kwajalein reflects the importance of El
Nin˜o throughout the tropical Pacific. The earliest strong
event in this record is the El Nin˜o of 1957–58, which
is indicated by a rise in sea level followed by a 9-cm
drop. The strongest event overall is the 1982–83 El Nin˜o
when sea level dropped by 20 cm. The differences be-
tween observed and analyzed sea level appear to be of
longer than interannual timescale. We examine the de-
cadal behavior of these records below. Interestingly, the
anomaly correlations in Table 2 are quite comparable
to correlations reported previously by Miller and Cane
(1996) for much shorter 2-yr intervals that also included
the seasonal cycle.
In order to determine the quality of the comparison
at frequencies longer than interannual we next filter all
time series with a 5-yr low-pass filter. A linear trend is
also removed in order to eliminate unmodeled geo-
physical effects such as geologic uplift and global sea
level rise due to continental ice melt as well as model
bias. The resulting gauge comparisons are labeled Cor2
F
EBRUARY
2000 315CARTON ET AL.
T
ABLE
3. Rms differences between observed and control analysis
temperature and salinity along WOCE hydrographic sections.
Name Location Rms (T)
(8C) Rms (S)
(psu)
Atlantic
A1E
A11
A16
A9
AR4E
AR15
52.28N, 238W–148W
44.68S, 508W–128W
448–598N, 208W
198S, 368W–138W
58S–58N, 34.58W
68S–28N, 34.58W
0.43
1.29
0.32*
0.36*
0.93
0.74
0.047
0.195
0.031
0.156
0.138
0.113
Indian
I5 33.58S, 388–728E 0.45 0.058
Pacific
P1-2
P1-3
P4-2
P4-3
P4-4
P6W
478N, 1478W–1268W
478N, 1478W–1268W
9.38N, 1608E–1608W
9.38N, 1658W–1108W
9.38N, 1608E–1608W
308S, 1548E–1788E
0.29*
0.32*
0.62
0.54
0.51
0.66
0.130
0.091
0.074
0.062
0.053
0.052
P6C
P6E
P16S
P16C
P17S
P17C-2
P17C-3
32.38S, 1658W–1208W
32.38S, 1108W–888W
338S–178S, 150.38W
17.38S–9.38N, 150.68W
198C–68S, 1348W
68S–0.48N, 134.68W
0.48N–34.48N, 134.68W
0.55
0.56
0.50
0.72
0.51
0.44
0.82
0.060
0.073
0.067
0.116
0.124
0.067
0.105
PR3
PR13J
PR13N
PR16
PR20
348N–42.28N, 1448E
258N–488N, 1658E
43.28S, 1488E–1668E
18N–68N, 1108W
21.5N
1.36
0.78
1.41*
1.01
0.87
0.120
0.160
0.065
0.150
0.071
* Temperature (but not salinity) from these cruises has been as-
similated in the control analysis.
in Table 2. In two-thirds of the stations in the tropical
Pacific the correlation of observed and analysis sea level
is improved by low-pass filtering. In mid and high lat-
itude the correlation also improves, but generally re-
mains below 0.5. The arithmetic average correlation of
the two sets of records are 0.48 and 0.56. If these values
are representative they would suggest that the control
analysis explains 23% of the interannual sea level var-
iance, increasing to 31% on timescales between 5 and
25 years. A significant fraction of the unexplained var-
iance is due to mesoscale processes that cannot be re-
solved by our analysis and is of less interest to us (the
contribution of mesoscale variability is evident in the
comparison of the nearby Honolulu and Hilo time se-
ries).
The three pairs of low-pass filtered time series shown
in Fig. 2 include one (San Juan) that has a correlation
of less than 0.5 and two (Honolulu and Kwajalein) that
exceed 0.7. At all three stations a visual comparison
reveals substantial similarity. The main differences at
San Juan seem to be in the specific timing of the anom-
alies rather than their amplitude or duration. At Hon-
olulu the major differences between observed and anal-
ysis sea level occur prior to 1960, when the subsurface
data coverage was less complete.
3. Additional comparisons
The comparisons reported in this section involve da-
tasets with more limited temporal coverage but with
expanded spatial coverage (altimeter sea level, WOCE
hydrography, and drifter and moored currents).
a. Altimeter sea level
Although the time series comparisons discussed
above provide information about interannual to decadal
variability, they have limited spatial coverage and have
only indirect information about currents. The availabil-
ity in recent years of satellite altimeter sea level offers
us the opportunity to examine the accuracy of the anal-
ysis essentially globally. In order to make the compar-
ison to altimeter sea level independent we introduce a
new experiment, experiment 8, in which altimeter sea-
level information has been excluded from the updating
procedure. We limit our discussion hereto consider only
the TOPEX/Poseidon altimetry because of its low ob-
servation error.
The error in observed monthly averaged TOPEX/Po-
seidon altimetric sea level has been estimated to be in
the neighborhood of 2 cm in the tropical Pacific (Cheney
et al. 1994; Mitchum 1994). The rms difference between
observed and experiment 8 sea level exceeds the ob-
servation error by 2–4 cm (Fig. 3a). The difference is
lowest in the Tropics and somewhat loweron the eastern
side of the basin than on the western side. In the eastern
tropical basin the rms difference drops below 3 cm. For
the whole tropical belt 158S–158N the rms difference is
around 4.0 cm.
The rms difference increases in regions of high eddy
generation such as the regions of western boundary cur-
rent extensions and in the Antarctic Circumpolar Cur-
rent. A few ‘‘bull’s-eyes’ appear in Fig. 3a. Close ex-
amination shows that these result from mislocated XBTs
that are then inconsistent with altimeter sea level. One
example of a mislocated XBT is at 188N, 1858W in Fig.
3a. The rms difference for the full 628S–628N domain
is 5.2 cm.
When altimetric sea level is used as a constraint on
the analysis temperature and salinity fields (but not pres-
sure since pressure is not a prognostic model variable),
the rms difference is reduced by 1–2 cm (Fig. 3b). In
the tropical Atlantic and Pacific the differences lie in
the range 2–3 cm. The average for the full tropical belt
158S–158N becomes 3.1 cm. In the mesoscale eddy pro-
duction regions of the midlatitudes the rms difference
again increases due to limitations in model physics and
poorer error statistics.
b. WOCE hydrographic transects
The World Ocean Circulation Experiment is an in-
ternational program designed to define the large-scale
316 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 4. Temperature and salinity with depth along WOCE meridional transect P17 near longitude 1348W
in the eastern Pacific Ocean during June–July 1991. The observed transect was made of three sections. Upper
panel shows the difference between observed and control analysis, middle panel shows observations, lower
panel shows analysis. (a) Temperature and (b) salinity.
structure of the ocean. The field program, including a
broad array of observations, has been concentrated dur-
ing the period since 1989. A primary dataset consists
of a series of high quality one-time hydrographic sur-
veys transecting the major oceans with some repeat sec-
tions. With the exceptions noted below, this dataset was
not included in the data archive of Levitus et al. (1994)
and consequently provides us with wonderful indepen-
dent data for comparison.
Not all of the WOCE hydrography is publicly avail-
able. From the more limited set of data available from
the Scripps Institution of Oceanography mirror site of
the WOCE hydrographic program Special Analysis
Center in Hamburg, we have extracted 16 transects list-
ed in Table 3. The temperature and salinity data along
each transect has been linearly interpolated on constant
depth surfaces to the model grid coordinates and then
compared to the monthly averaged temperature and sa-
linity fields from the control analysis. Along each sec-
tion the rms difference between observed and control
analysis temperature and salinity has been computed at
the model levels and averaged in depth and distance
F
EBRUARY
2000 317CARTON ET AL.
F
IG
.4.(Continued)
T
ABLE
4. Rms near-surface anomalous observed velocity compo-
nents and rms differences between anomalous observed and analysis
velocity components. Zonal and meridional anomalous velocity com-
ponents are computed with respect to the 1988–93 monthly clima-
tology and averaged over the Pacific basin between 308S and 408N.
Units: centimeters per second.
Year Rms (u9) Rms (
y
9) Rms (Du9) Rms (D
y
9)
1988
1989
1990
1991
1992
1993
1988–93
25.4
20.4
19.0
21.3
20.6
21.5
20.6
15.2
13.3
12.7
12.1
12.9
12.5
12.4
16.7
17.0
13.5
14.1
13.5
12.1
13.3
11.3
10.1
9.3
9.5
8.9
8.9
8.9
along the cross section. In some cases the transectshave
been decomposed into smaller sections when they span
more than a single month.
The large error in temperature for A11 can be un-
derstood because of its location in the poorly sampled
Southern Hemisphere and the high degree of eddy var-
iability. The errors for PR3 in the subtropical North
Pacific and PR16 in the tropical Pacific are more sur-
prising. The transects with large salinity errors generally
have two kinds of errors. Either the salinity errors are
large in the mixed layer (PR3, AR15, AR4E), indicating
problems with surface fluxes, or the errors are confined
below the mixed layer. Two of the transects with sub-
stantial salinity errors below the mixed layer (P1,
318 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 5. Comparison of anomalous observed and control analysis near-surface annual-averaged anomalous
currents in the tropical Pacific for the year 1991. Upper panel shows observations the control analysis, while
lower panel shows currents from currents based on drifter. Anomalous currents have been computed relative
to the 1988–93 monthly climatology. Observed and analysis tropical currents show a distinctive eastward
anomaly during this year.
PR13J) do not have large corresponding errors in tem-
perature, suggesting that the salinity errors are the result
of errors in horizontal advection.
One particularly interesting transect, labeled P17, cuts
through the eastern Pacific from 198Sto348N. This
transect was broken into several segments that are sep-
arated by vertical lines in Figs. 4a,b. At subtropical
latitudes comparison of observed and analysis temper-
ature fields shows that the most significant error is at
thermocline depths. Analysis temperature is too low by
18–28C suggesting that the thermocline is too shallow
by 10–20 m. The largest error is between 38N and 88N.
This band of latitudes corresponds to the North Equa-
torial Countercurrent trough that separates the northern
extension of the South Equatorial Current from the
North Equatorial Countercurrent. Weakness in the
F
EBRUARY
2000 319CARTON ET AL.
F
IG
. 6. Comparison of observed and control analysis zonal current
time series on the equator with depth at 1408W in the eastern Pacific.
Note the presence of strong interannual variability associated with
El Nin˜o.
trough implies that the transports in these two currents
are weak.
Comparison of observed and analysis salinity fields
also shows that significant error is present at thermocline
depths as well as in the mixed layer (Fig. 4b). South of
the equator the observed transect shows evidence of
subduction and equatorward transport of high salinity
subtropical water. The observed salinity maximum is
narrowly confined in depth within a few degrees of the
equator. In contrast, the analysis salinity shows a broad-
er, weaker subsurface salinity maximum. This difference
indicates that the analysis is subducting subtropical wa-
ter, but not as rapidly as is observed. One factor may
be that the mixed layer salinity reaches a maximum of
36.2 psu, 0.2 psu lower than observed. In contrast, the
analyzed mixed layer salinity between 68and 128Nis
more than 0.2 psu higher than observed. In summary,
the major sources of error include those associated with
errors in the mixed layer and subduction processes,
large-scale bias, and unresolved mesoscale variability.
c. Surface drifters
The most extensive spatial coverage of velocity mea-
surements is provided by the WOCE/TOGA surface
drifter velocity program (Niiler et al. 1999, manuscript
submitted to J. Phys. Oceanogr.). Although some mea-
surements were collected in the early 1980s, extensive
coverage is only available since 1988. We haveobtained
the data presented here from the Atlantic Oceanographic
Marine Laboratory/NOAA, where the drifter data has
been converted to Eulerian velocities and averaged into
8832831 month bins. We compute velocity anomalies
by removing the 1988–93 monthly climatology from the
observations and analyses. Table 4 shows the spatial
average of the rms velocity anomaly components, as
well as the rms differences between observed and anal-
ysis anomalous velocity components for each year. We
approach this comparison with trepidation since near-
surface velocity is a difficult field to simulate. The 6-
yr averages show that the rms anomalous zonal velocity
difference (13.3 cm s
21
) is 30% less than the rms zonal
velocity itself (20.6 cm s
21
). Some similarity in ob-
served and analysis variability is present in almost every
year in both zonal and meridional components.
The spatial pattern of anomalous velocity for 1991,
the year of the beginning of a strong El Nin˜o, is shown
in Fig. 5. The surface velocity during this year is char-
acterized by a strong 10 cm s
21
eastward surge along
and just north of the equator in response to therelaxation
of the trade winds [see Frankignoul et al. (1996) for
discussion of the surge]. South of the equator and north
of 108N the velocity is weakly westward. In the extra-
tropics the observed velocity components are confused.
The control analysis velocity also shows an eastward
surge of water of somewhat higher amplitude than ob-
served, extending not quite as far toward the coast of
South America. North and south of this surge westward
return currents are apparent, as observed. The analysis
velocity in the extratropics is of lower amplitude than
observed. We think that the reduced amplitude of an-
alyzed velocity at the oceanic mesoscale is the result of
dissipation of mesoscale eddies by the forecast model.
d. Equatorial Pacific moored current
The anomalous eastward surface velocity at the equator
during 1991 has corresponding subsurface changes. InFig.
6 we compare observed current from the Tropical Ocean–
Atmosphere mooring maintained by the Pacific Marine
Environmental Laboratory at 1408W. At this longitude,
close to the longitude of the section P17 shown in Fig. 4,
the westward South Equatorial Current is confined to the
upper 40 m where its annual average speed rarely exceeds
220 cm s
21
. During 1991 the South Equatorial Current
shoals to 20 m and actually disappears early in the year.
The analysis velocity at the same longitude as the mooring
shows a significant westward bias (Fig. 4). Thus, the South
Equatorial Current is too strong and the Equatorial Un-
320 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 7. Annual cycle of (a) kinetic energy and (b) sea level from the control analysis. The annual cycle
is defined as the annual harmonic of the Fourier series. Sea level has strong annual amplitude in the high
variability regions of the western boundary current. The maximum amplitude in these regions approaches
15 cm. Units are m
2
s
22
and cm.
dercurrent velocity is too weak at this location. The control
analysis surface South Equatorial Current is relatively
weak during 1991, while the Equatorial Undercurrent does
not weaken until 1992. From these results and additional
experiments we conclude that better meridional resolution
and stronger trade winds are required to improve the anal-
ysis of tropical currents.
4. Global statistics
The strongest signal in the mass and momentum fields
is the annual shift of heat and mass in response to shift-
ing winds and surface heat flux. We introduce our sta-
tistical analysis by discussing the annual cycle of two
key quantities, surface kinetic energy (u
2
1
y
2
)/2 and
sea level based on the 46-yr control analysis (Figs. 7a,b).
The annual cycle of a third variable, heat content, will
be discussed separately. The basin-scale structure of the
annual cycle of sea level is dominated by a pattern of
rising level in the summer hemisphere as a result of the
antisymmetry about the equator of winds and solar heat-
ing. Smaller-scale variations are also evident. Western
boundary current regions of the North Atlantic and
North Pacific have seasonal amplitudes approaching 15
cm (these results closely resemble those computed from
short 1-yr-long altimeter records by Cheney et al. 1994).
The Tropics also have distinct smaller-scale features.
The tropical Atlantic and Pacific show 4–6 cm varia-
tions in zonal bands resulting from seasonal changes in
the North Equatorial Countercurrent. The corresponding
amplitude of heat content in these regions (not shown)
is 2008Cm.
In contrast to sea level the annual cycle of kinetic
energy (Fig. 7b) is largest in the Tropics, reflecting the
seasonal changes in the tropical current system. The
annual amplitude of these currents exceeds 20 cm s
21
.
In the tropical Pacific the intensity of the current is
reduced somewhat near 1408W. The annual kinetic en-
F
EBRUARY
2000 321CARTON ET AL.
F
IG
.7.(Continued)
ergy in the Indian Ocean is elevated throughout, with
highest values along the western boundary in the region
of the seasonal Somali Current.
Seasonal variations in local heat storage, the time rate
of change of heat content, reflect the difference between
net surface heating and horizontal divergence of heat
transport. Because the ocean basins are bounded to the
east and west at most latitudes, the zonal average of
heat storage, shown in Figs. 8a–c for the three basins,
is the difference between net surface heating and the
meridional divergence of heat transport. Here storage is
computed by taking the center difference from succes-
sive monthly averages. These results can be directly
compared with the results of Hsiung et al. (1989). Heat
storage to a shallower 275-m depth, but using the ex-
panded WOA-94 dataset, is presented in Levitus and
Antonov (1997). Heat storage in the Pacific follows the
cycle of solar radiation with a maximum in the Northern
Hemisphere in June–August. The seasonal maximum is
somewhat lower than in either of the two previous stud-
ies. Maximum heat loss from the Southern Hemisphere
occurs perhaps a few weeks earlier and is of significantly
lower amplitude than Levitus and Antonov. The pattern
of heat storage near the equator resembles that of Lev-
itus and Antonov with a complex set of zonal bands of
heat gain and loss. The period July–October is one in
which the region from 108S–08is storing heat rapidly
as a result of the appearance of the cold tongue in the
eastern Pacific, while the region from 08–108N is losing
heat rapidly.
Heat storage in the Atlantic (Fig. 8b) generally re-
sembles heat storage in the Pacific. In the North Atlantic
maximum heat storage occurs somewhat earlier (May–
July). Heat is being exported mainly during early boreal
summer. In the tropical Atlantic rapid heat storage is
limited to the months June–September. South of the
equator storage occurs during September–February.
Heat storage in the tropical Indian Ocean (Fig. 8c) is
quite complicated. Between 208S and 58S the southern
Indian Ocean is gaining heat during June–October, a
period in which the sun is actually in the NorthernHemi-
sphere. This region is losing heat during November–
March, a period in which the sun has crossed over into
the Southern Hemisphere! These counterintuitive results
are generally consistent with those of Hsiung et al.
(1989) and Levitus and Antonov (1997).
322 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 8. Zonal average of monthly climatological 0/500 m heat
storage for the three ocean basins. Positive values indicate that the
ocean is gaining heat. Units are 8Cmmo
21
.
At periods longer than the annual cycle the distri-
bution of variance changes. In Fig. 9 we present a crude
decomposition of the spectral characteristics of the heat
content field by separating the variability into an inter-
annual band (1–5 yr) and a decadal band (5–25 yr). At
interannual periods the variability along the equator in
the Pacific is enhanced and shifted eastward, while off-
equatorial variability occurs in the western side of the
basin. The western tropical Indian Ocean has significant
variability, as do the eddy production regions of the
western boundary currents and Antarctic Circumpolar
Current. At decadal periods the Tropics become less
prominent relative to the subtropical and midlatitude
gyres. The shift of variability from the Tropics toward
midlatitude as the frequency decreases reflects the fun-
damental dynamical properties of the ocean.
Much of the heat content variability at interannual
and decadal periods in Fig. 9 is spatially incoherent. In
order to explore the sources of just that part of the heat
content signal that has basin scales we define three focal
areas, one in the central North Pacific (averaged, 308
458N, 1808–2108W), a second in the eastern tropical
Pacific (averaged 58S–58N, 1508–908W), and a third in
the central North Atlantic (averaged 258–408N, 708
308W). Note that the focal area in the North Atlantic is
smaller than the others, reflecting the smaller spatial
scales of variability there. The time series of area-av-
eraged heat content anomaly are presented in Fig. 10
(a linear trend has been removed, consistent with the
previous analysis). In order to define the spatial structure
of these modes we correlate the three heat content anom-
aly time series with heat content anomaly throughout
the global ocean. The three resulting correlation maps
are shown in Fig. 11.
The time series and spatial pattern of correlation with
North Pacific heat content is shown in the upper panels
of Figs. 10 and 11. The variability is substantially de-
cadal. The most prominent feature of the time series is
a cooling during the 1980s following a relatively warm
1970s. Heat content variability in the North Pacific has
been related to fluctuations in the Pacific–North Amer-
ica pattern of wind variation (Trenberth and Hurrell
1994; Graham 1994). Changes in the wind patterns in
1976–77 led to a cooling in the central basin that was
referred to as a climate shift by Miller et al. (1994).
This climate shift is evident in Fig. 10 but is followed
11 years later by a return to warm conditions, also noted
by Levitus and Antonov (1995).
By correlating the North Pacific time series with heat
content anomaly time series for the rest of the oceans
we can identify the spatial pattern of oceanic thermal
variability. The region that is most highly correlated
with the midlatitude North Pacific, interestingly, is the
western tropical and southern Pacific. Much of the cor-
relation results from a drop in heat content in the 1970s
that approximately coincides with the drop in the central
North Pacific. Between these two regions of positive
correlation is a region of negative correlation. This pat-
tern of alternating positive and negative correlation is
reminiscent of ideas of a slow advection of thermal
anomalies throughout the North Pacific (e.g., Latif and
Barnett 1994; Deser et al. 1996).
Next we turn our attention to the tropical Pacific. The
variability here is strongly interannual, reflecting the
importance of El Nin˜o. Individual peaks in the time
series can be identified with individual El Nin˜os. Ex-
amination of the time series shows that the amplitude
F
EBRUARY
2000 323CARTON ET AL.
F
IG
. 9. Root-mean-square 0/500 m heat content variability in the (a) interannual band (1–5 yr)
and (b) decadal band (5–25 yr). Units are 8Cm.
and frequency of heat content variability has been fairly
regular during the past five decades. Decadal variability
is also evident in the time series. The late 1970s were
a period of gradual warming coinciding with a warming
of SST (Wang 1995).
Examination of the spatial pattern of correlation with
tropical Pacific heat content shows the spatial structure
of the oceanic expression of El Nin˜o. For example, we
find that on either side of the zonal band of high cor-
relation near the equator are bands of negative corre-
lation. The band of negative correlation in the Northern
Hemisphere is most coherent, extending west-south-
westward from the west coast of North America. The
existence of off-equatorial bands of negative correlation
is a key feature of the delayed-oscillator theories of the
periodicity of El Nin˜o. The zone of negative correlation
extends into the eastern Indian Ocean, suggesting a con-
nection between the western Pacific and eastern Indian
Ocean. Examination of the lagged correlation shows that
the band of positive correlation near the equator shifts
eastward and poleward with time, while the off-equa-
torial bands of negative correlation shift westward.
The impact of El Nin˜o on the North and tropical
Atlantic has been examine in a number of studies (e.g.,
Carton and Huang 1994; Enfield and Meyer 1997). In
the tropical Atlantic Carton and Huang have proposed
that the changes in western tropical Atlantic winds as-
sociated with warm SST in the eastern tropical Pacific
leads to a buildup of heat in the western tropical At-
lantic. This anomalous heat, which occurs in the form
of anomalous deepening of the thermocline, eventually
leads to a deepening of the equatorial thermocline and
contributes to warming in the eastern Atlantic. Some
confirmation for these ideas is apparent in the weakly
positive correlation (0.2) between eastern tropical Pa-
cific heat content and western tropical Atlantic heat con-
tent.
The final time series in the lower panel of Fig. 10
represents the central North Atlantic Ocean. Heat con-
tent variability in the North Atlantic is also associated
with variability in the sector winds, in this case the North
Atlantic Oscillation pattern Atlantic winds (Hurrell and
van Loon 1997). The time series of heat content shows
rich decadal variability. The 1950s are relatively warm,
followed by a series of cool events in the early 1960s,
the late 1960s and early 1970s, and again in the late
1970s and early 1980s. The second of these cool events
has been documented by S. Levitus (Levitus 1990).
5. Conclusions
In a companion to this study Carton et al. (2000)
present a five-decade-long historical analysis of global
324 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
F
IG
. 10. Time series of heat content anomaly from the seasonal
cycle for three regions, (a) North Pacific (averaged, 308–458N, 1808
2108W), (b) Nino3 region of the eastern tropical Pacific (averaged
58S–58N, 1508–908W), and (c) North Atlantic (averaged 258–408N,
708–308W). The regions are indicated in the corresponding panels of
Fig. 11. Units are 8Cm.
upper-ocean temperature, salinity, sea level, and cur-
rents. The purpose of the analysis is to provide a uni-
formly gridded historical dataset for use in studies of
the ocean’s role in climate. Two ways to quantify the
accuracy of the analysis are by direct comparison to
independent observations and by examining identifiable
climate features such as the annual cycle, El Nin˜o, and
the Pacific–North American anomaly. Here wetake both
approaches.
We begin with a comparison to time series with lim-
ited spatial coverage and to global observation sets with
limited temporal coverage. The island tide-gauge time
series suggests that the analysis explains 25% of the
observed sea level variance at longer than annual fre-
quencies and more than 30% in the frequency band
between 5 and 25 years. The explained variance in-
creases in the tropical Pacific. Comparison to satellite
altimeter sea level shows a root-mean-square difference
of 4.0 cm in the Tropics 158S–158N and 5.2 cm globally
when this dataset is not assimilated. When altimetry is
assimilated, the rms difference in the tropical sea level
decreases to 3.1 cm. The comparison to a series of global
hydrographic sections shows average temperature and
salinity errors in the upper 500 m of 0.708C and 0.092
psu. The temperature errors are mainly concentrated at
thermocline depths. Salinity errors are distributed
through the mixed layer and pycnocline. A comparison
to moored and surface drifter velocity shows the anal-
ysis reproduces the qualitative behavior of surface ve-
locity in the Tropics.
We begin the discussion of climate features by con-
sideration of the annual cycle. At the annual period both
hemispheres show a strong response to seasonal forcing
of surface winds, heat, and freshwater. In the subtropics
and midlatitudes sea level varies in phase with seasonal
changes in solar radiation, except in the southern Indian
Ocean. Closer to the equator the phase of sea level un-
dergoes rapid reversals as the ocean responds to strong
annual variations in winds. The implied rates of heat
storage are consistent with previous studies.
The strongest basin-scale signal at interannual periods
is associated with El Nin˜o. To explore this pattern of
variability we examine the correlation of global heat
content with the heat content time series from a region
of the eastern equatorial Pacific (the Nino3 region) that
is itself considered an index of El Nin˜o. Our exami-
nation of the zero-lag correlation of global heat content
shows the eastern and western tropical Pacific to be out
of phase (correlation 20.4 to 20.6). The eastern Indian
Ocean is in phase with the western Pacific and thus is
out of phase with the eastern Pacific. The North Pacific
has a weak positive correlation with the eastern equa-
torial Pacific. Correlations between eastern Pacific heat
content and Atlantic heat content are modest.
At longer decadal periods the Pacific–North America
wind pattern leads to broad patterns of correlation in
heat content variability. Increases in heat content in the
central North Pacific are associated with decreases in
heat content in the subtropical Pacific and increases in
the western tropical Pacific. Atlantic heat content ispos-
itively correlated with the central North Pacific. The
Atlantic–Pacific relationship is confirmed by correlating
the heat content anomaly in the central North Atlantic
with global heat content.
In several respects the analysis is clearly inconsistent
with the observations. Major problems include:
1) The inability of the forecast model to produce ther-
mocline water masses in sufficient volume. Two ex-
amples of this problem are identified, the subtropical
mode water of the North Atlantic and the high sa-
linity subtropical water that enters the equatorial
thermocline from the south in the Pacific and At-
lantic. Interannual changes in these water masses
change the covariances of temperature and salinity
errors, and thus violates an assumption of our anal-
ysis. The major causes of insufficient thermocline
water-mass production are still not clear. Adjoint or
streamline assimilation techniques together with is-
F
EBRUARY
2000 325CARTON ET AL.
F
IG
. 11. Correlation of three heat content anomaly time series shown in Fig. 10 with global heat content
anomaly. (a) North Pacific, (b) tropical Pacific, and (c) North Atlantic.
opycnal coordinate forecast models may prove help-
ful.
2) The inability of the forecast model to produce re-
alistic mixed layer salinity. Our lack of knowledge
of historical surface salinity limits any estimate of
interannual fluctuations of salinity in the mixed layer
and contributes to the problems associated with in-
sufficient thermocline water-mass formation. Im-
provements in estimate of historical rainfall may help
address this problem.
3) The inability of the forecast model to simulate the
effects of important narrow topographic features.
Our lack of resolution limits the ability of the fore-
cast model to resolve features such as the Strait of
Gibraltar, the Windward Islands, or the Florida
Straits in the North Atlantic, all of which have pro-
326 V
OLUME
30JOURNAL OF PHYSICAL OCEANOGRAPHY
found effects on the general circulation. Lack of res-
olution limits mesoscale eddy production and alters
the dynamical balances in narrow currents such as
the Gulf Stream and the equatorial current systems.
On the other hand, increasing resolution will increase
production of mesoscale eddies. These eddies must
be treated as noise since they are not resolved by
the observation set prior to the availability of satellite
altimetry.
4) The errors in the Indian Ocean and the Southern
Hemisphere generally. The adequacy of the historical
subsurface data varies with location. Generally
speaking, the temperature of the North Atlantic and
major shipping routes of the North Pacific is rea-
sonably well sampled throughout most of the last
five decades. Sampling in the tropical Atlantic and
Pacific Oceans is intermittent, while the Indian
Ocean and the southern oceans are always inade-
quately sampled. Sampling of overlying atmospheric
variables is similarly limited. More salinity obser-
vations are also badly needed, as are measurements
below 500 m. Upgrades to WOA-94 are expected to
help fill in some of these data voids (S. Levitus 1997,
personal communication).
Despite these problems the authors feel greatly en-
couraged by the potential of the analysis for upper-ocean
climate studies on interannual to decadal timescales.
Acknowledgments. We are very grateful to a number
of people who have given us access to their datasets.
Mark Swenson and Zengxi Zhou of the Atlantic Ocean-
ographic Marine Laboratory have provided the monthly
averaged surface drifters, and Sydney Levitus and Rob-
ert Cheney and their colleagues at the National Ocean-
ographic Data Center/NOAA have provided access to
the hydrographic and altimeter data. We have benefited
from the datasets collected by principal investigators of
the Tropical Ocean Global Atmosphere/World Ocean-
ographic Circulation Experiments. Finally, we want to
express our gratitude for support from the Office of
Global Programs/NOAA under Grant NA66GP0269 and
the National Science Foundation under Grant
OCE9416894.
REFERENCES
Bennett, A. F., 1990: Inverse Methods in Physical Oceanography.
Cambridge University Press, 346 pp.
Carton, J. A., and B. Huang, 1994: Warm events in the tropical
Atlantic. J. Phys. Oceanogr., 24, 888–903.
, G. Chepurin, X. Cao, and B. S. Giese, 2000: A simple ocean
data assimilation analysis of the global upper ocean 1950–95.
Part I: Method. J. Phys. Oceanogr., 30, 294–309.
Cheney, R. E., L. Miller, and J. Lillibridge, 1994: Topex/Poseidon:
The 2 cm solution. J. Geophys. Res., 99, 24, 555–563.
Chepurin, G. A., and J. A. Carton, 1999: Comparison of retrospective
analyses of the global ocean heat content. Dyn. Atmos. Oceans,
submitted.
da Silva, A. M., C. C. Young, and S. Levitus, 1994: Atlas of Surface
Marine Data 1994, Vol. 1: Algorithms and Procedures. NOAA
Atlas NESDIS 6, U.S. Department of Commerce, NOAA, NES-
DIS, 83 pp.
Daley, R., 1991: Atmospheric Data Analysis. Cambridge University
Press, 457 pp.
Deser, C., M. A. Alexander, and M. S. Timlin, 1996: Upper-ocean
thermal variations in the North Pacific during 1970–91. J. Cli-
mate, 9, 1840–1855.
Enfield, D. B., and D. A. Meyer, 1997: Tropical Atlantic sea surface
temperature variability and its relation to El Nin˜o–Southern Os-
cillation. J. Geophys. Res., 102, 929–945.
Frankignoul, C., F. Bonjean, and G. Reverdin, 1996: Interannual var-
iability of surface currents in the tropical Pacific during 1987–
1993. J. Geophys. Res., 101, 3629–3647.
Graham, N. E., 1994: Decadal-scale climate variability in thetropical
and North Pacific during the 1970s and 1980s: Observations and
model results. Climate Dyn., 10, 135–162.
Hsiung, J., R. E. Newell, and T. Houghtby, 1989: The annual cycle
of oceanic heat storage and oceanic meridional heat transport.
Quart. J. Roy. Meteor. Soc., 115, 1–28.
Hurrell, J. W., and H. van Loon, 1997: Decadal variations in climate
associated with the North Atlantic Oscillation. Climatic Change,
36, 301–326.
Ji, M., A. Leetmaa, and J. Derber, 1995: An ocean analysis system
for seasonal to interannual climate studies. Mon. Wea. Rev., 123,
460–481.
Latif, M., and T. Barnett, 1994: Causes of decadal climate variability
over the North Pacific and North America. Science, 266, 634–
637.
Levitus, S., 1990: Interpentadal variability of steric sea level and
geopotential thickness of the North Atlantic Ocean, 1970–1974
versus 1955–1959. J. Geophys. Res., 95, 5233–5238.
, and J. Antonov, 1995: Observational evidence of interannual
to decadal-scale variability of the subsurface temperature–salin-
ity structure of the world ocean. Climatic Change, 31, 495–514.
, and , 1997: Climatological and Interannual Variability of
Temperature, Heat Storage, and Rate of Heat Storage in the
Upper Ocean. NESDIS Atlas Series 16, NOAA, 6 pp.
, T. P. Boyer, and J. Antonov, 1994: World Ocean Atlas, 1994.
Vol. 5: Interannual Variability of Upper Ocean Thermal Struc-
ture, NESDIS Atlas Series, NOAA, 176 pp.
Malanotte-Rizzoli, P., 1996: Modern Approaches to Data Assimilation
in Ocean Modeling. Elsevier, 455 pp.
Miller, A. J., D. R. Cayan, and T. P. Barnett, 1994: The 1976–77
climate shift of the Pacific Ocean. Oceanography, 7, 21–26.
Miller, R. N., and M. Cane, 1996: Tropical data assimilation: The-
oretical aspects. Modern Approaches to Data Assimilation in
Ocean Modeling, P. Malanotte-Rizzoli, Ed., Elsevier, 207–234.
Mitchum, G., 1994: Comparison of TOPEX sea surface heights and
tide gauge sea levels. J. Geophys. Res., 99, 24 541–24 553.
Reynolds, R. W., and T. M. Smith, 1994: Improved globalsea surface
temperature analysis using optimum interpolation. J. Climate, 7,
929–948.
Trenberth, K. E., and J. W. Hurrell, 1994: Decadal atmosphere–ocean
variations in the Pacific. Climate Dyn., 9, 303–319.
Wang, B., 1995: Interdecadal changes in El Nin˜o onset in the last
four decades. J. Climate, 8, 267–285.
Wunsch, C., 1996: The Ocean Circulation Inverse Problem. Cam-
bridge University Press, 442 pp.
... Ocean reanalyses originate from the efforts of meteorologists in the 1990s to reconstruct the time evolution of the global atmosphere state [1]. One application of these atmosphere reanalyses is to provide surface forcing fields which the oceanographers use to drive ocean models [2][3][4][5]. Similar to atmosphere reanalyses, ocean reanalyses, which combine different types of observations with ocean models, have the potential to provide dynamically consistent and accurate estimates of the ocean states [6]. ...
... Ocean reanalyses at the global and regional scales have already become available and been an established activity in some research and operational centres with enhancements of ocean observing systems such as satellite data, Argo floats, and other observing platforms. The SODA (simple ocean data assimilation) ocean reanalysis was produced by the optimal interpolation method in the global oceans [2,3,7]. The ocean reanalysis systems developed at the European Centre for Medium-Range Weather Forecasts (ECMWF) used different assimilation methods including optimal interpolation and variational assimilation NEMOVAR in different versions [8,9]. ...
Article
Full-text available
This paper describes an ocean reanalysis system in the Indian and Pacific oceans (IPORA) and evaluates its quality in detail. The assimilation schemes based on ensemble optimal interpolation are employed in the hybrid coordinate ocean model to conduct a long-time reanalysis experiment during the period of 1993–2020. Different metrics including comparisons with satellite sea surface temperature, altimetry data, observed currents, as well as other reanalyses such as ECCO and SODA are used to validate the performance of IPORA. Compared with the control experiment without assimilation, IPORA greatly reduces the errors of temperature, salinity, sea level anomaly, and current fields, and improves the interannual variability. In contrast to ECCO and SODA products, IPORA captures the strong signals of SLA variability and reproduces the linear trend of SLA very well. Meanwhile, IPORA also shows a good consistence with observed currents, as indicated by an improved correlation and a reduced error.
... To improve the representation of physical processes, the outer domain solution is based on a data-assimilative reanalysis of the California Current circulation (Neveu et al., 2016). For the outer domain, ROMS is forced at the surface with the 0.25°resolution Cross-Calibrated Multi-Platform winds of Atlas et al. (2011) and open boundary conditions are derived from the Simple Ocean Data Assimilation reanalysis of Carton et al. (2000). For the inner domain, boundary conditions are updated daily using the outer domain solution and surface atmospheric forcing is imposed as 6-hourly momentum, heat, and freshwater fluxes at 1/10°corrected by data assimilation to minimized model-data error in the outer domain. ...
Article
Full-text available
Coastal upwelling variability in the California Current region, one of the four main eastern boundary current upwelling systems, is controlled by processes acting over a wide range of spatial and temporal scales. While the ensuing ecosystem response depends strongly on upwelled water properties, determining their exact physical and biogeochemical characteristics is notoriously difficult as it requires tracking water masses backward in space and time from the moment they upwell near the coast to their subsurface origin. Adjoint model simulations have been used successfully to track water masses in coastal upwelling systems and the work presented here extends these applications to determining the co‐variability of physical and biogeochemical properties of source waters at spatial scales that resolve the known alongshore variability of coastal upwelling in the region. Notably, the results identify that the modulation of coastal upwelling efficiency by onshore/offshore geostrophic meanders is the dominant mechanism explaining alongshore variability in source depth and properties of upwelled waters. The simulations also reveal that source water properties vary seasonally in response to different balances between coastal upwelling intensity and biogeochemical processes. During spring, interannual variability of physical and biogeochemical properties is directly tied to the intensity of upwelling‐favorable alongshore winds, whereas, during summer, biogeochemical properties respond more strongly to biological activity and subsequent organic matter remineralization at depth. Overall, the present work provides important insight into the mechanisms responsible for the alongshore mosaic and seasonal variation of upwelled source water properties in the central California Current region.
... To investigate the subsurface temperature anomaly before, during, and after the period of ESMHWs detected from the OISST data, we first identified best-performing reanalysis among four products using potential temperature, practical salinity, potential density, and current velocity data from: 1) Simple Ocean Data Assimilation (SODA 3.3.1) at a resolution of 1/4 • for 1990-2015 (Carton et al., 2000a;Carton et al., 2000b); 2) Estimating the Circulation and Climate of the Ocean (ECCOv4r4) at a resolution of 1/2 • for 1992-2018 (Forget et al., 2015;ECCO Consortium et al., 2020a;2020b); 3) Hybrid Coordinate Ocean Model (HYCOM) at a resolution of 1/12 • for 1994-2015 (Cummings, 2005;Cummings and Smedstad, 2013;Metzger et al., 2014); 4) Global Ocean Reanalysis product (GLORYS) at a resolution of 1/12 • for 1993-2020 (Jean-Michel et al., 2021). Latent heat flux, sensible heat flux, upward longwave radiation, downward solar radiation, surface net longwave radiation, surface net solar radiation and wind speed were derived from the hourly NCEP Climate Forecast System Reanalysis (CFSR) and Climate Forecast System Version 2 (CFSv2) data that are used as the surface forcing for the HYCOM (Cummings, 2005;Cummings and Smedstad, 2013;Helber et al., 2013). ...
... The minimum depth of the model was set to 10 m near the shore. The open boundary of this domain was obtained from the monthly average data reanalyzed using Simple Ocean Data Assimilation (SODA) [48]. The initialization of POMgcs was a two-step process: In the first step, a cold start, we integrated the model using the December climate state from the World Ocean Atlas 2005 (WOA05) [49] as the initial temperature and salinity fields. ...
Article
Full-text available
The uncertainty in the initial condition seriously affects the forecasting skill of numerical models. Targeted observations play an important role in reducing uncertainty in numerical prediction. The conditional nonlinear optimal perturbation (CNOP) method is a useful tool for studying adaptive observation. However, the traditional CNOP method highly relies on the adjoint model, and it is difficult to find the global optimal solution. In this paper, a pre-screening and ensemble CNOP hybrid method called PECNOP is proposed to identify optimal sensitive areas in targeted observations. PECNOP is an adjoint-free method that captures global CNOP with high probability, which can effectively solve the two major problems faced by traditional CNOP methods. We evaluated the performance of PECNOP by building an observation simulation system consisting of an ocean model and data assimilation. One of the assimilation experiments was dedicated to evaluating the stability and effectiveness of PECNOP in extreme events. The results show that, compared with traditional methods, PECNOP can stably capture the global CNOP. Extra observations and assimilation in the optimal sensitive areas identified by PECNOP can effectively improve forecasting by about 20% within 30 days. Therefore, PECNOP has potential to reduce the initial error of numerical models, which is important for improving forecasting.
... W) at 1/30° (~ 3 km) resolution forced by an outer domain for the broader CCS (30-48° N, 115-134° W) at 1/10° (~ 10 km) resolution with physical data assimilation. The physical circulation is forced at the surface by the 0.25° resolution Cross-Calibrated Multi-Platform (CCMP) winds 33 , and physical initial and boundary conditions for the outer domain are derived from the Simple Ocean Data Assimilation (SODA) reanalysis 34 . A complete description of the physical model configuration can be found in Fiechter et al. 18 . ...
Article
Full-text available
Ocean acidification is progressing rapidly in the California Current System (CCS), a region already susceptible to reduced aragonite saturation state due to seasonal coastal upwelling. Results from a high-resolution (~ 3 km), coupled physical-biogeochemical model highlight that the intensity, duration, and severity of undersaturation events exhibit high interannual variability along the central CCS shelfbreak. Variability in dissolved inorganic carbon (DIC) along the bottom of the 100-m isobath explains 70–90% of event severity variance over the range of latitudes where most severe conditions occur. An empirical orthogonal function (EOF) analysis further reveals that interannual event variability is explained by a combination coastal upwelling intensity and DIC content in upwelled source waters. Simulated regional DIC exhibits low frequency temporal variability resembling that of the Pacific Decadal Oscillation, and is explained by changes to water mass composition in the CCS. While regional DIC concentrations and upwelling intensity individually explain 9 and 43% of year-to-year variability in undersaturation event severity, their combined influence accounts for 66% of the variance. The mechanistic description of exposure to undersaturated conditions presented here provides important context for monitoring the progression of ocean acidification in the CCS and identifies conditions leading to increased vulnerability for ecologically and commercially important species.
... datasets at a water depth of 5 m and a spatial resolution of 0.5 • × 0.5 • (Carton et al., 2018). Although the SODA SSS data are inaccurate before the 1980s (Carton et al., 2000), previous studies have confirmed its continued reliability at seasonal and lower frequency trends in the Indo-Pacific region (Du and Qu, 2010). Monthly precipitation information was retrieved from the Global Precipitation Climatology Project version 2.3 (GPCP v2.3) datasets at a 2.5 • × 2.5 • spatial resolution (Adler et al., 2003). ...
Article
As a gauge for ocean precipitation conditions, sea surface salinity (SSS) is suspected to be an excellent indicator for variations in the global hydrological cycle. The SSS conditions in the South China Sea (SCS) are closely related to precipitation changes of the East Asian monsoon. However, the characteristics and driving mechanisms of the meridional pattern of East Asian monsoon rainfall during the warm periods of the late Holocene are not well understood because of a lack of high-resolution proxies from the SCS. Here, we present three precise chronologies of monthly stable coral skeleton oxygen isotopes (δ¹⁸O) and pair them with established composite strontium/calcium (Sr/Ca) records from the Xisha Islands to obtain a continuous reconstruction of the residual δ¹⁸O (Δδ¹⁸O, regarded as seawater δ¹⁸O) during 1980–2007 CE (Common Era), 1149–1205 CE, and 2070–2011 a BP (years before 1950 CE). The results confirm that coral Δδ¹⁸O is a good tracer for reconstructing SSS and precipitation changes in the northern SCS on seasonal and interannual timescales, with higher Δδ¹⁸O values implying more saline conditions and less precipitation. The mean value of fossil coral Δδ¹⁸O for 1149–1205 CE was higher than that of modern coral, suggesting the existence of salty surface waters and notable dry climate episodes in the northern SCS during the late Medieval Climate Anomaly (MCA). In contrast, the average coral Δδ¹⁸O for 2070–2011 a BP was significantly lighter by about 0.404‰ compared to that of recent decades, indicating that less saline and more humid conditions occurred in the northern SCS during the early Roman Warm Period (RWP). Synthesizing our coral records with other published precipitation reconstructions from eastern China suggested a meridional dipole spatial pattern of moisture variation over East Asia during the historical warm periods of the Holocene. During the MCA, drier conditions generally prevailed in the region south of the Yangtze River, and more humid conditions in the north. This may be closely related to the contemporaneous strong El Niño activity occurring at this time, which probably enhanced the East Asian summer monsoon (EASM) and caused a northward shift of the intertropical convergence zone (ITCZ), possibly inducing increased water vapor transport from the SCS to North China. During the RWP, the relatively weak El Niño–Southern Oscillation (ENSO) variability may weaken the intensity of the EASM and shift the ITCZ southward, possibly resulting in a spatial pattern of “wet south and dry north” in East Asia.
... . The model was initialized starting in 1980, using the Simple Ocean Data Assimilation (SODA) analysis of the global upper ocean63 . The 8-year spin-up period is sufficiently long to mitigate any dependence of the model solution on the initial conditions, typical of shorter-term predictions of chaotic ...
Article
Full-text available
In California offshore waters, sustained northwesterly winds have been identified as a key resource that can contribute substantially to renewable energy goals. However, the development of large-scale offshore wind farms can reduce the wind stress at the sea surface, which could affect wind-driven upwelling, nutrient delivery, and ecosystem dynamics. Here we examine changes to upwelling using atmospheric and ocean circulation numerical models together with a hypothetical upper bound buildout scenario of 877 turbines spread across three areas of interest. Wind speed changes are found to reduce upwelling on the inshore side of windfarms and increase upwelling on the offshore side. These changes, when expressed in terms of widely used metrics for upwelling volume transport and nutrient delivery, show that while the net upwelling in a wide coastal band changes relatively little, the spatial structure of upwelling within this coastal region can be shifted outside the bounds of natural variability.
... The surface forcing fields for the WC12/WC15 domains consist of WRF fields from the outer/inner WRF domains, linearly interpolated onto the ROMS domains. Boundary and initial conditions for the WC12 model Offshore wind upwelling are derived from a data assimilative reanalysis of a previous version of the WC12 model[21], which in turn was initialized by the Simple Ocean Data Assimilation (SODA) analysis of the global upper ocean[56].Both ROMS models (WC12 and WC15) have been utilized extensively in physical oceanographic and ecosystem studies of the California Current System, in addition to forming the basis of operational modeling at the Central California Ocean Observing System. The validated WC12 model ...
Preprint
Full-text available
In California offshore waters, sustained northwesterly winds have been identified as a key energy resource that could contribute substantially to California’s renewable energy goals. However, the development of large-scale offshore wind farms has the potential to reduce the wind stress at the sea surface, which could have implications on wind-driven upwelling, nutrient delivery, and ecosystem dynamics. Using atmospheric and ocean circulation models together with a hypothetical upper boundary buildout scenario of 877 turbines spread across three wind energy areas of interest, wind speed reductions in the lee of the wind energy areas of interest are found to enhance horizontal wind stress shear, leading to reduced upwelling on the inshore side of windfarms and increased upwelling on the offshore side. These changes, when expressed in terms of widely used metrics for upwelling volume transport and nutrient delivery, show that while the net upwelling in a wide coastal band changes relatively little, the spatial structure of upwelling within this coastal region can be shifted outside the bounds of natural variability.
Article
Full-text available
Past analyses of tropical Atlantic sea surface temperature variability have suggested a dipole behavior between the northern and southern tropics, across the Intertropical Convergence Zone (ITCZ). By analyzing an improved 43-year (1950-1992) record of SST [Smith et al., 1996] and other data derived from the Comprehensive Ocean-Atmosphere Data Set (COADS), it is shown that the regions north and south of the ITCZ are statistically independent of each other at the seasonal to interannual timescales dominating the data, confirming the conclusions of Houghton and Tourre [1992]. Some dipole behavior does develop weakly during the boreal spring season, when there is a tendency for SST anomaly west of Angola to be opposite of that in the tropical North Atlantic. It is further shown that tropical Atlantic SST variability is correlated with Pacific El Niño-Southern Oscillation (ENSO) variability in several regions. The major region affected is the North Atlantic area of NE trades west of 40°W along 10°N-20°N and extending into the Caribbean. There, about 50-80% of the anomalous SST variability is associated with the Pacific ENSO, with Atlantic warmings occurring 4-5 months after the mature phases of Pacific warm events. An analysis of local surface flux fields derived from COADS data shows that the ENSO-related Atlantic warmings occur as a result of reductions in the surface NE trade wind speeds, which in turn reduce latent and sensible heat losses over the region in question, as well as cooling due to entrainment. This ENSO connection is best developed during the boreal spring following the most frequent season of maximum ENSO anomalies in the Pacific. A region of secondary covariability with ENSO occurs along the northern edge of the mean ITCZ position and appears to be associated with northward migrations of the ITCZ when the North Atlantic warmings occur. Although easterly winds are intensified in the western equatorial Atlantic in response to Pacific warm events, they do not produce strong local changes in SST. Contrary to expectations from studies based on equatorial dynamics, these teleconnected wind anomalies do not give rise to significant correlations of SST in the Gulf of Guinea with the Pacific ENSO. As the teleconnection sequence matures, strong SE trades at low southern latitudes follow the development of the North Atlantic SST anomaly and precede by several months the appearance of weak negative SST anomalies off Angola and stronger positive anomalies extending eastward from southern Brazil along 15°-30°S.
Article
Full-text available
A newly available, extensive compilation of upper-ocean temperature profiles was used to study the vertical structure of thermal anomalies between the surface and 400-m depth in the North Pacific during 1970-1991. A prominent decade-long perturbation in climate occurred during this time period: surface waters cooled by 1°C in the central and western North Pacific and warmed by about the same amount along the west coast of North America from late 1976 to 1988. Comparison with data from COADS suggests that the relatively sparse sampling of the subsurface data is adequate for describing the climate anomaly.The vertical structure of seasonal thermal anomalies in the central North Pacific shows a series of cold pulses beginning in the fall of 1976 and continuing until late 1988 that appear to originate at the surface and descend with time into the main thermocline to at least 400-m depth. Individual cold events descend rapidly (100 m yr1), superimposed upon a slower cooling (15 m yr1). The interdecadal climate change, while evident at the surface, is most prominent below 150 m where interannual variations are small. Unlike the central North Pacific, the temperature changes along the west coast of North America appear to be confined to approximately the upper 200-250 m. The structure of the interdecadal thermal variations in the eastern and central North Pacific appears to be consistent with the dynamics of the ventilated thermocline. In the western North Pacific, strong cooling is observed along the axis of the Kuroshio Current Extension below 200 m depth during the 1980s.Changes in mixed layer depth accompany the SST variations, but their spatial distribution is not identical to the pattern of SST change. In particular, the decade-long cool period in the central North Pacific was accompanied by a 20 m deepening of the mixed layer in winter, but no significant changes in mixed layer depth were found along the west coast of North America. It is suggested that other factors such as stratification beneath the mixed layer and synoptic wind forcing may play a role in determining the distribution of mixed layer depth anomalies.
Article
Full-text available
The authors describe a 46-year global retrospective analysis of upper-ocean temperature, salinity, and currents. The analysis is an application of the Simple Ocean Data Assimilation (SODA) package. SODA uses an ocean model based on Geophysical Fluid Dynamics Laboratory MOM2 physics. Assimilated data includes temperature and salinity profiles from the World Ocean Atlas-94 (MBT, XBT, CTD, and station data), as well as additional hydrography, sea surface temperature, and altimeter sea level. After reviewing the basic methodology the authors present experiments to examine the impact of trends in the wind field and model forecast bias (referred to in the engineering literature as 'colored noise'). The authors believe these to be the major sources of error in the retrospective analysis. With detrended winds the analysis shows a pattern of warming in the subtropics and cooling in the Tropics and at high latitudes. Model forecast bias results partly from errors in surface forcing and partly from limitations of the model. Bias is of great concern in regions of thermocline water-mass formation. In the examples discussed here, the data assimilation has the effect of increasing production of these water masses and thus reducing bias. Additional experiments examine the relative importance of winds versus subsurface updating. These experiments show that in the Tropics both winds and subsurface updating contribute to analysis temperature, while in midlatitudes the variability results mainly from the effects of subsurface updating.
Article
In this chapter, some of the theoretical issues underlying the application of optimized methods of data assimilation to the tropical oceans are discussed. By “optimized” methods of data assimilation, we mean methods which minimize some objective measure of error. Methods formulated in this way are cast in terms of statistical hypotheses, which can be tested by standard statistical methods.The efficacy of simple models of the tropical ocean has been a major advantage in the practice of data assimilation for this region. We discuss physical reasons for the effectiveness of these simple models, but also remind the reader that much of this apparent simplicity stems from the nature of the agenda in tropical oceanography. Since the focus in the community is on phenomena relevant to ocean-atmosphere interaction and climate prediction, the highest priority is large scale, low frequency low latitude motions. More complex models are necessary for reasonably accurate descriptions of the dynamics of the tropical ocean on shorter spatial or temporal scales, or more than about 10° from the equator.We discuss some of the theory of the data assimilation methods as such, and conclude that the crucial research issues revolve around the prior error estimates that largely determine the product of any practical data assimilation method.
Article
Buoy drifts and current meter records between January 1987 and December 1993 are used to investigate the interannual variability of the equatorial Pacific currents at a depth of 15 m. The sampling is coarse until mid-1988 but more complete afterward, so that the large-scale features of the anomaly currents can be documented on the seasonal to yearly timescale. Using objective analysis, bimonthly current anomalies are mapped between 20°N and 20°S on a 1°×5° grid, and the error covariance matrix of the analyzed fields are estimated. The current anomalies are primarily zonal, with largest amplitudes within about 8° from the equator, and they are largely linked to the El Niño-Southern Oscillation phenomenon. In particular, broad, basin-wide westward anomaly currents were encountered during the 1988 La Niña, and strong eastward currents persisted from July-August 1991 to January-February 1992, followed by westward currents from May-June to July-August 1992. An empirical orthogonal function (EOF) analysis shows that the first EOF of zonal current anomaly is largely uniform in the equatorial band, while the next two EOFs describe large-scale currents of opposite sign across the equator and across 160°W, respectively. The EOFs are rather smooth and the errors on the principal component time series relatively small, which indicates that the sampling is adequate to describe the large-scale, low-frequency zonal current fluctuations. As the dominant EOFs of meridional current are noisy and the relative errors on the principal components larger, the meridional current fluctuations are not as well captured by the data set. Correlation analysis and a singular value decomposition are used to investigate the influence of advection by the large-scale, low-frequency currents on sea surface temperature (SST) anomalies during 1987-1993. Although the data set is noisy and other terms play an important role in the SST anomaly equation, the effect of zonal and, to a lesser extent, meridional advection is seen in much of the central and eastern equatorial Pacific. The dominant terms are the anomalous zonal advection of mean SST, the mean zonal, and, intermittently, meridional advection of SST anomalies.
Article
The unprecedented accuracy of TOPEX/POSEIDON (T/P) altimeter data warrants a new evaluation of the methods typically used to form time series of sea level change. Whereas explicit removal of orbit error has always been required as a first step in altimeter data processing, the T/P analysis presented here is based simply on unadjusted, monthly averages. This approach has the advantage of retaining the large-scale ocean signal, which would be distorted by orbit adjustment. Using 16 months of data, we have evaluated the T/P monthly means on spatial scales ranging from mesoscale to global. In the tropical Pacific, comparisons with 17 island tide gauge records and dynamic height derived from 36 thermistor moorings show that the altimeter data have an accuracy of approximately 2 cm when averaged over spatial scales a few hundred kilometers. On basin scales in the northern hemisphere, similar agreement is found between the T/P data and the dynamic height climatology of Levitus (1982). These new altimeter observations are thus providing the first reliable view of global sea level changes on seasonal-to-interannual timescales.
Article
We have composited and objectively analyzed hiostorical hydrographic observations for the North Atlantic Ocean for two periods, 1955-1959 and 1970-1974. Difference fields of steric sea level and geopotential thickness for the North Atlantic Ocean in the 0- to 1500-m depth range between the 1955-1959 and 1970-1974 pentads were computed. We find that in the central portion of the subtropical gyre, steric sea level decreased by up to 17.5 dyn cm, whereas in the western subarctic gyre (north of the Gulf Stream) steric sea level increased by an amount up to 7.5 dyn cm from earlier to the later pentad. A decrease in steric sea level (~5 dyn cm) occurred along the eastern boundary of the North Atlantic.
Article
This book presents the newest, innovative applications combining the most sophisticated data assimilation methods with the most complex ocean circulation models. First, it reviews models and observations from the data assimilation perspective. Second, for each oceanographic application, from the global to the regional scale, it offers reviews and new results of fundamental assimilation methodologies and strategies as well as of the state-of-the-art of operational ocean nowcasting/forecasting. Finally, the last chapter presents an example of interdisciplinary modeling with data assimilation components. The 16 chapters contained in the book are abstracted separately in Oceanographic Literature Review. The book is available from Elsevier Science, P.O. Box 211, 1000 AE Amsterdam, The Netherlands, or, in the USA/Canada, from Elsevier Science, P.O. Box 945, Madison Square Station, New York, NY 10160-0757, USA.