+ All Categories
Home > Documents > The Journal of Marine Research is an online peer-reviewed journal...

The Journal of Marine Research is an online peer-reviewed journal...

Date post: 10-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
25
Journal of Marine Research, Sears Foundation for Marine Research, Yale University PO Box 208118, New Haven, CT 06520-8118 USA (203) 432-3154 fax (203) 432-5872 [email protected] www.journalofmarineresearch.org The Journal of Marine Research is an online peer-reviewed journal that publishes original research on a broad array of topics in physical, biological, and chemical oceanography. In publication since 1937, it is one of the oldest journals in American marine science and occupies a unique niche within the ocean sciences, with a rich tradition and distinguished history as part of the Sears Foundation for Marine Research at Yale University. Past and current issues are available at journalofmarineresearch.org. Yale University provides access to these materials for educational and research purposes only. Copyright or other proprietary rights to content contained in this document may be held by individuals or entities other than, or in addition to, Yale University. You are solely responsible for determining the ownership of the copyright, and for obtaining permission for your intended use. Yale University makes no warranty that your distribution, reproduction, or other use of these materials will not infringe the rights of third parties. This work is licensed under the Creative Commons Attribution- NonCommercial-ShareAlike 4.0 International License. To view a copy of this license, visit http://creativecommons.org/licenses/by-nc-sa/4.0/ or send a letter to Creative Commons, PO Box 1866, Mountain View, CA 94042, USA.
Transcript
Page 1: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

Journal of Marine Research, Sears Foundation for Marine Research, Yale University PO Box 208118, New Haven, CT 06520-8118 USA

(203) 432-3154 fax (203) 432-5872 [email protected] www.journalofmarineresearch.org

The Journal of Marine Research is an online peer-reviewed journal that publishes original

research on a broad array of topics in physical, biological, and chemical oceanography.

In publication since 1937, it is one of the oldest journals in American marine science and

occupies a unique niche within the ocean sciences, with a rich tradition and distinguished

history as part of the Sears Foundation for Marine Research at Yale University.

Past and current issues are available at journalofmarineresearch.org.

Yale University provides access to these materials for educational and research purposes only. Copyright or other proprietary rights to content contained in this document may be held by

individuals or entities other than, or in addition to, Yale University. You are solely responsible for determining the ownership of the copyright, and for obtaining permission for your intended use.

Yale University makes no warranty that your distribution, reproduction, or other use of these materials will not infringe the rights of third parties.

This work is licensed under the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License. To view a copy of this license, visit http://creativecommons.org/licenses/by-nc-sa/4.0/ or send a letter to Creative Commons, PO Box 1866, Mountain View, CA 94042, USA.

Page 2: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

Evidence for wind-forced circulation in the Gulf of Mexico

by John Blaha1 and Wilton Sturges'

ABSTRACT A study is conducted into the response of sea level and dynamic height to fluctuations of

alongshore wind stress and wind stress curl at periods greater than a few months per cycle. Monthly tide gauge data from Key West to Progreso, Mexico, during 1954 to 1974 are

adjusted to remove the effects of local atmospheric pressure and seasonal steric heating. The adjusted mean monthly sea level elevations are significantly greater from Progreso to Port Isabel than they are elsewhere in the Gulf. This observation remains unchanged after the elevations are reduced for the effect of local alongshore winds. Among the tide gauges in the western Gulf, Galveston is the most coherent with the local alongshore wind forcing at periods greater than 2 mo/ cycle, exhibiting a phase with the winds not significantly different from 'TT.

At the other coastal sites, at least half of the elevation signal remains. This residual signal is presumed to be caused by the geostrophic flu ctuations of an offshore boundary current.

The available wind data from the western half of the Gulf show a negative wind stress curl; the mean is -11 X 10----i dyne/ cm8 and curl is most negative in July. A common feature in the sea level elevations from Progreso to Port Isabel and in curl is the sharp transition from sum• mer to fall. It is suggestive of a seasonal component to the Gulf circulation forced by the wind stress curl. This transiti on occurs from July to September in curl but from August to October in sea level, a one month lag. The observed 17 cm of change in elevation corresponds to 23 X Io-' dyne/ cm' of change in curl.

A mean baroclinic circulation in the northwestern Gulf is evident in which the mean difference in dynamic height (sea surface relative to 700 db) from offshore to inshore regions is about 14 dynamic cm. The total seasonal variation across the flow (after the influence of Loop Current rings has been minimized in the data) is about 5 dynamic cm, which is one--third the above change in sea level attributed to curl.

1. Introduction We explore the correlations among winds, coastal sea level, and dynamic heights

in the western Gulf of Mexico for those periods greater than a few months per cycle. We work from the hypothesis that at the appropriate scales these variables are dynamically related to the circulation of the western Gulf according to the

l. Pacific Marine Environmental Laboratory/NOAA, Seattle, Washington, 98105, U.S.A. 2. Department of Oceanography, Florida State University, Tallahassee, Florida, 32306, U.S.A.

711

Page 3: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

712 Journal of Marine Research [39, 4

30°~ GALVE STON

25°

TA MP ICO

20°N V ERA CR UZ

100°w 95° 90° 85° 90° Figure I. The locations of the tide gauges and wind stations used in this study. The contours

of dynamic height at the surface relative to 1000 db are taken from Nowlin and McLellan (1967) and depict conditions in the Gulf of Mexico during March 1962.

classical (e.g., Stommel, 1948) models of wind-driven ocean circulation. Our ob-jective is to present that evidence for curl-forced circulation in the Gulf of Mexico.

Nowlin and McLellan (1967) and Nowlin (1972) have presented distributions of dynamic height from the western Gulf of Mexico; their dynamic height map of a single winter cruise is reproduced in Figure 1. Sturges and Blaha (1976) proposed that these data were consistent with an anticyclonic gyre which is largely forced by the wind stress curl. They observed that the anticyclonic circulation inferred from the dynamic heights coincided with a strongly negative wind stress curl based upon Hellerman's (1967) compilation of wind stress. Another, and rather different, cal-culation of the curl distribution is shown in Figure 2. Hydrographic data from the western Gulf always indicate an anticyclonic flow (J. Cochrane, personal com-munication). Sturges and Blaha also observed that the spatial contours of dynamic height showed asymmetry towards the west and north, typical of other wind-driven western boundary currents, e.g., the Gulf Stream.

Unfortunately, the winter satellite images show frequent clouds, and in the sum-mer the sea surface is isothermal. One cloud-free satellite IR image from 18 April 1975 (R. Legeckis, personal communication) shows a feature in the thermal gra-

Page 4: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 713

34o,---------------------~

30°

20°N

100°w 90°

Figure 2. Wind stress curl computed from satellite observations of low-level cloud winds. The data are hourly pictures; cloud winds are reduced to the surface and rotated to the left on the basis of observed surface winds. The average is from 16 June to 23 August 1974. From Krishnamurti and Krishnamurti (1979). Units are 10_, dynes/ cm'. The triangle grid points enclose a region within which the average wind stress curl has been computed based upon FNOC winds.

dients similar to those fronts generally associated with the deep water currents such as the Gulf Stream. One can infer from this image a north-south current at the western boundary extending from south of 22N to roughly 27.5N. Figure 3 shows the depth of the main thermocline in the western Gulf in July 1979. The scale of the flow is greater than 500 km north-south, but no clear pattern of closed contours emerges in the offshore, return flow. Other appearances in both satellite IR images and hydrographic data indicate a smaller north-south scale to the anticyclonic flow. For example, Nowlin and McLellan's late winter data in Figure 1 show the flow separating from the Mexico-Texas coast near 24N. A separate flow cell appears to the north in this and another recent survey by Merrell and Morrison (1981).

The wind stress curl has a large negative mean value over the western Gulf north of 22N. Classical linear theory for a steady wind-driven subtropical gyre specifies that across a line of latitude the transport of the western boundary current balances the Sverdrup, curl-driven, interior transport. We show selected, spatially averaged, dynamic heights from which mean transport in the western boundary current can be estimated. However, intensely baroclinic rings detach from the Loop Current (the strong ridge of geopotential seen in the eastern Gulf in Figure 1) and sub-sequently drift into the western Gulf, as described by Elliott (1979). The intensity of these rings produces a serious signal-to-noise problem for the study of the back-

Page 5: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

714 Journal of Marine Research [39, 4

28°N

26°

24°

22°

14 JULY 79, t4 • C ISOTHERM DEPTHS, IN METERS

20•1---~9.6-0 W---9~4.---9·2------'~9'0_0 ___,

Figure 3. Depth of 14 'C isotherm in western Gulf of Mexico from XBT and AXBT data, July 1979. Dots show AXBT data points from 14 July; line shows the cruise track of R.V. Longhorn, 24 July-I August. The duration of the AXBT flight was about 10 hours.

ground wind-driven flow. The hydrographic sampling is too sparse to yield reliable estimates of the seasonal (average January, average February, etc.) variations in tho dynamic heights. In lieu of dynamic heights, we have investigated the response of the sea level elevations to the fluctuations in wind stress curl.

A spectral multiple regression method is employed to determine the joint re-sponse of the elevations at a tide gauge to local alongshore wind stress and to wind stress curl. At periods greater than 2 mo/ cycle, the effect of bottom friction on the shelf dominates over the local acceleration. The Ekman (1905) model implies a phase of nearly zero or 1r between sea level and wind stress depending upon the orientation of the water and shoreline. When geostrophic currents offshore of the continental shelf do not mask locally-driven shelf currents, one generally finds cor-relation between the alongshore shelf flow and sea level. Previous investigations, however, have concentrated at higher frequencies, as for example, Kundu et al. (1975), and Beardsley and Butman (1974).

We find that a large signal in seasonal sea level (the residual elevations) remains after the effects of local alongshore winds are removed. We suggest that the residual elevations of sea level are an effect of the geostrophic fluctuations in the flow of an offshore boundary current. In a comparison which we do not show here. we have

Page 6: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 715

found that the coastal elevations of sea level at Key West (after adjustment for local atmospheric pressure and seasonal steric heating) agree with the elevation dif-ferences Key West minus Havana across the Florida Current. Reid and Mantyla ( 197 6) have found that from southern California to Washington, the seasonal dynamic heights which arise from an integration across the bottom, from 1000 db to the tide gauge, correlate with the elevations at the tide gauge. In these two cases the coastal elevations reflect the fluctuations in the cross-stream pressure gradient of an alongshore flow which is in total or in part offshore of the continental shelf. At the seasonal time scales, we assume that the surface geostrophic flow is cor-related with the flow at depth and that the boundary flow along the continental slope in the western Gulf is an effect of the interior transport which may be driven by the wind stress curl. We will show that there is agreement between the residual elevations and the seasonal fluctuations in curl over the western Gulf.

Finally, we use the observed phase between sea level and curl to compute the frictional damping parameter in an analytical model by Pedlosky (1965) of a closed ocean basin under periodic wind forcing. Pedlosky ( 1965) treats the case of a square basin of homogeneous water on a beta-plane. The vertically integrated, non-dimensional vorticity equation is

V' cf, + cf,,+ R,(cf,, V2cf,, -cf,, V'cf,,) = T(x,y,t) -8V 2cf,, (1)

where cf, is the strearnfunction and T is the wind stress curl. The time scale is chosen as (/3L)-1, in which L is the meridional scale of the basin. The frictional damping parameter 8 is equal to a constant coefficient of frictional drag divided by {3L. (This form of friction is independent of scale and does not allow for increased frictional damping along the boundaries of the basin.) The zonal dimension is denoted by x, meridional by y; the boundary conditions are:

cf, = 0 on x = 0, x = 1 and y = 0, y = 1 . (2)

Pedlosky assumes weakly nonlinear flow, a small Rossby number R,, and solves the problem to the first nonlinear correction by means of a regular expansion in R,.

The dependence of the forced solution on frequency is of central importance. Pedlosky applies the forcing

T = cos(kx -wt) sin(mry) (3)

At the higher frequencies, w > > 8, the circulation response is dominated by the resonant modes. We have neglected the solution in this frequency range since we felt that it would be especially sensitive to the actual configuration of the basin and that its comparison to actual data would be of doubtful value. Pedlosky shows that as the forcing period approaches the frictional time scale, w ~ 0(8), friction at the western boundary is effectively trapping the westward propagating energy; western

Page 7: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

716 Journal of Marine Research [39, 4

intensification occurs. If the forcing is reduced to a zonally uniform standing wave, k = 0, we find that at the western boundary the alongshore flow is proportional to

ii<f, - [(1 8 ) cos wt - w sin wt ] sin mry. (4) iix,-o - 62 + w' 62 + w'

Equation ( 4) represents the cross-stream change in sea level nearest to the coast; we will compute a decay time based upon it. We begin the analysis with a discussion of the available wind stress data.

2. Surface wind stress

Aside from the satellite data, the historical wind data over the ocean are from: (a) ship observations and (b) geostrophic computations using atmospheric pressure data. Hastenrath and Lamb (1977) and Elliott (1979) show one degree compilations of quarterly wind stress over the Gulf based upon the ship observations. From the historical sea level pressure field Fleet Numerical Oceanography Center (FNOC) can provide computed surface wind stress on about a 3 degree grid resolution at ten-day averages. We have interpolated the FNOC wind stress to the grid points shown in Figure 2, which lie over the region of anticyclonic circulation. We have computed an average curl by integrating the wind stress around the region enclosed by the grid points and applying the Stokes integral theorem. All sets of wind stress data show a region in the western Gulf with negative wind stress curl.

A comparison of the seasonal variations in curl appears in Figure 4. Considering that Elliott 's (1979) values represent averages over the entire western Gulf, the agreement in phase is satisfactory. However, Elliott's maps indicate positive curl over a region roughly 600 km across in the southwestern corner of the Gulf. This

feature is not adequately resolved in the FNOC data, but a cyclonic circulation cell beneath it has been noted in the hydrographic data by Vasquez (1975).

Comparisons of the alongshore stress along the western boundary also appear in Figure 4 (see Appendix for further comparisons). There are differences in detail to which aliasing from a lack of ship observations may be contributing. Over much of the western Gulf there are usually 5 or fewer sampled days per 2 ° box in any particular month. Offshore of Tampico, in the region 21-23N, 96-98W, the median and mean number of observations per month from 1952-1972 are 2 and 4.1 , re-

Figure 4. Seasonal comparisons of FNOC winds and ship observations of wind in wind stress curl (upper) and in meridional wind stress off shore of Tampico Oower). The tong-term av-erages of curl based upon ship observations (right band side scale) are taken from Elliott (1979), bis Figure IS, and are averaged over the entire Gulf. Our (FNOC) curl Oeft band side scale) is computed within the region shown in Figure 2. The drag coefficients are 1.5 X 10-•. The record lengths are 19S2-1974 for the FNOC winds and 1952-1972 for the ship observations.

Page 8: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

FNOC 0 24 - 27°N, 96-90°W

0 • SHIP (ELLIOTT )

-4 10

,,., L

6 u -s ' w z "' >- 0 0 -12 2

"' >< 0 ><

-16 -2

-20 -6

SHIP

• 21-25°N, 96-9S• w 21-23°N, 96-98° W

.4 FNOC 0 22.4°N, 96.7"W

.2 \

\

\ 0 I

C\J I

L I I u 0

' If w -.2 ·Z \

I >- I 0 I I 0

-.4 I 0 I I I

I I

- 6 I I I I \ I ~-

-.8 J M M J s N J

MONTH

Page 9: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

718 Journal of Marine Research [39, 4

spectively. An expanded region 21-25N, 96-98W increases the median number of samples per month to only 5. During winter, this sampling rate would not ade-quately resolve the 3 to 4 synoptic fronts passing over the Gulf per month. Where time series can be constructed from adequate ship observations adjacent to the tide gauges, the ship winds are used in the subsequent comparisons between sea level and wind stress. Otherwise, when airport winds are not available, we have chosen. the better temporal resolution provided by the FNOC estimates of wind stress. Measurements of surface winds (twice per day) were obtained at Brownsville and Merida. For purposes of determining stress, the anemometer heights Oess than 11 m) were assumed to be within a logarithmic wind profile and Zo was arbitrarily chosen as 2.0 cm. Unless explicitly stated, stresses are computed at the sampling interval and then averaged over the month.

3. Dynamic height

Rings detached from the Loop Current represent a large amplitude perturbation on the background flow field. Typically, the historical data base contains only a few hydrographic casts in each one degree square and the variability within the data is large. We have attempted to eliminate hydrographic data clearly within rings and to improve upon the signal-to-noise ratio in the western Gulf by pooling the avail-able data into three regions. We define region I as the northern cell above 26N, which appears in Figure 1 as a relative high with respect to region II that is directly to the east: 93-90W, 26-28N. Region III occupies the main anticyclonic gyre: 97-92W, 22-25N. The data base consists of all hydrographic data west of 90W that were available from NODC for the years through 1972 and from two recent Coast Guard surveys (1975, 1976). These data were first intercalibrated according to the deep temperature, salinity properties that were assumed to be common to all the data during this period. By examining in detail the distribution of data in each month we attempted to (a) account for meanders in the high shear zone, the current which separates regions I and III ; and (b) to eliminate casts clearly in Loop Current rings. The latter was achieved by examining the subsurface maximum in salinity,. ~ 150 m, for salinities greater than 36.55%0 that were accompanied by high dynam-· ic heights. From 213 casts deeper than 700 db, 17 were clearly judged to be within rings, but the marginal casts remain in the data pool. ' The dynamic heights were, averaged within the regions for each month having data. These averages were then again averaged into the mean monthly values shown in Figure 5. The 125 db· surface has been chosen to be the bottom of the surface layer, above which there, are large steric effects from seasonal heating. The reference surface of 700 db was·.

3 We hoped to suppress the large "n oise" effect of rings by this technique, because the rings are not sampled well. However, thi s will bias the results toward lower dynamic heights. Although the scatter-in the data will be reduced, the results will underestimate the value of the central high and of the~ cross-str eam difference in dynamic height.

Page 10: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 719

120 .~ ......

,,, "' ~

t\ 100 I • ···· ·• ····•/ . ....... )" __ o---o-,-'\\ . • ·. / b

90 ,,,.. -<f 11 o,---0...... .,,.,.a'"'

'tr

80 Ill

1251700 DB 50

M M s N

3 2 4 4 3 II 3 6 5 3 2 5 4 I 3

Ill 5 4 2 4 2 2 3

Figure 5. Mean monthly dynamic heights 0/700 db and 125/700 db for regions I, II, and m (sec text). Each monthly sample represents an individual year. The standard error of tho mean is shown when there arc sufficient data. Region Ill is tho primary dynamic high and region II is the low.

chosen because it nearly always extends to the bottom of the thermocline, yet it eliminates fewer casts than would a deeper choice.

Region II (the low southeast of Galveston in Figure 1) has the lowest overall mean value of dynamic height at 125 db, relative to 700 db. No months show a significant difference from a mean value of 60 dynamic cm. The mean difference between regions III and II is 9.1 :!: 1.3 dynamic cm at 125 db and 14.3 :!: 1.4 dynamic cm at the sea surface, relative to 700 db. (The contribution of the upper 125 db to the difference between regions ill and II is 5 :!: 2 dynamic cm and is nearly uniform throughout the year.)

Values of curl were extracted from those months during which hydrographic data were obtained. The mean value is - 16 x 10-• dynes/ cm', which for a nominal basin width of 1200 km yields a Sverdrup transport of 9 X 10• m'/ s. The ge-ostrophic baroclinic transport within the upper 700 db, between regions II and III,

Page 11: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

720 Journal of Marine Research [39, 4

was computed to be 8 x 10• m'/s. A deeper reference level would yield a greater transport.

In Figure 5 the dynamic heights of region I vary between those of regions II and III. This variation could result from the latitudinal movement of the boundary current of the main anticyclonic gyre (region nn . The flow could leave the coast at about 25N, separating I and III, as in Figure 1 or continue along the coast as in Figure 3. Examples of these two flow patterns have also been observed in the satellite IR images. The low dynamic height of region II would appear to be un-affected by these shifts in the boundary flow pattern. However, the sparseness of the data makes the significance of the seasonal variation of the baroclinic field from 700 to 125 db uncertain.

4. Coastal sea level

a. Adjustments for static effects

Monthly averages of hourly heights were obtained from tide gauges at Key West, St. Petersburg, Pensacola, Galveston, Port Isabel, Tampico, Veracruz, and Pro-greso. Data in the U.S. were provided by the National Ocean Survey; data in Mexico were provided by the Instituto de Geofisica, Mexico. The common time interval is January 1954 to December 1974, 21 years, although some comparisons will be longer or shorter to match the wind records.

We adjusted sea level to uniform atmospheric pressure using a barometer factor of 1 cm/mb. The barometric factor is the response of sea level, in cm, directly attributable to fluctuations in atmospheric pressure, in mb. Over the continental shelf and in the presence of friction, Crepon (I 976) calculates a barometric factor equal to 1 cm/mb after sufficient time has elapsed for the barotropic "spin-down." Thls time is estimated to be about 2 weeks in the Gulf if the bottom boundary layer on the shelf is 10 m thick and the average water depth is 100 m. This is sub-stantially shorter than the periods of greater than 2 mo/ cycle that are under study here. Observations of bottom pressure, e.g., Beardsley et al. (l 977), would imply even shorter adjustment times in sea level toward the "inverted barometer." The monthly values of sea level pressure are based upon hourly data at nearby airports, when available, or the monthly FNOC pressures. Some examples of the seasonal range are 6.5 mb at Brownsville and 5 mb at Tampa. The additional remaining rms fluctuation is about 5 mb.

In the central and northern Gulf, where the bathythermograph (BT) data are adequate to determine the steric effect of seasonal heating, Blaha (1978) estimates a variation across the Gulf of about 1 cm in a 14 cm signal at the surface relative to 120 db. Such a large, but nearly static, contribution to sea level ought to be removed prior to making comparisons with the winds. The eddy noise appearing in

Page 12: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Srurges: Gulf of Mexico wind-forced circulation

J M M J S N MONTH

J M M J S N MONTH

J M M J S N

MONTH

721

Figure 6. Mean monthly elevations of sea level, 1954-1974, adjusted to uniform atmospheric pressure, and from which the elevations at St. Petersburg have been subtracted. At St. Peters-burg the actual sea level elevations are shown and compared with the surface steric heights relative to 120 db, which are based upon BT data from the northern central Gulf (Blaha, 1978).

these BT data is great; it does not allow construction of a reliable time series of

steric heights. Instead, we have used the sea level elevations at St. Petersburg as a "bench mark" series which has been subtracted from the other time series of elevations around the Gulf. The mean monthly differences from the St. Petersburg elevations appear in Figure 6; the original St. Petersburg signal also appears. The steric heights which Blaha (1978) attributed to seasonal heating have been com-pared to the elevations at St. Petersburg in this figure as well. If allowances are made for local heating below 120 db, perhaps 1 to 2 dynamic cm, the agreement indicates that nearly all of the signal at St. Petersburg is in response to seasonal heating. Of course, some error may be introduced to the elevation differences be-cause of the effect of local winds at St. Petersburg. Let us define an anomaly as the difference between a monthly value and the average value for that month. From 1965 to 1974, when monthly resultant winds at the Tampa airport are routinely published, the correlation between anomalies of alongshore wind stress and sea level at St. Petersburg is .22 and maximum at zero lag. The rms sea level anomaly at St. Petersburg over the 10 years is 4.5 cm of which 22%, or 1 cm, can be attributed to local winds. A 1 cm error in the rms anomaly at Tampico, 7.2 cm, amounts to 14% but is acceptable. In the remaining text, all reference to sea level

Page 13: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

722 Journal of Marine Research [39, 4

C/MON 1 1 10" 70 '

, ' 7

CURL

il. u ' -~

r<l 70 :rE u ....... z >-0

0

~o

SEA LEVEL

Cl. u NQ

'-N

u

I 10.00

MON ·1 , O(?

Figure 7. Power spectra of anomalies of wind stress curl (FNOC) and of sea level at Progreso (- - - - -), Tampico ( __ ), and Port Isabel ( .. .. . ). Sea level trends over the period 1952-1972 have been removed. A 95% confidence interval for both spectra is shown and the effective frequency resolution is .067 cycle/ mo. Some bias is occurring at the very lowest frequencies in the curl spectrum . For comparison, at the 12, 6, and 4 mo/ cycle harmonics (arrows), the periodogram estimates of Tampico elevation are: 21, 4, and 1 x 1()1 cm1/

(cycle/ mo); and of curl are: 47, 7, and 5 X 10->< (dyn/ cm')'/(cycle/ mo), respectively.

elevations refers to the diff erences from the St. Petersburg elevations. This shift in

nomenclature applies to the mean monthly and the anomaly signals.

Page 14: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 723

8 EOF MODE 2

Fa. / \ \ I q I q I I

0 I \ I I u p 'o,P I a. o-cr 'b

-s

24 ?

EOF MODE I / \

16 I \ 8 I b

I ' \ 0

8 4 I "' I

0 I 0 u u I '-w

I z >-I 0 -s I -4 ())

I 0 I ><

-16 I -s I q I \ I

-24 \ ,P -12 ti

J M M J s N J

MONTH Figure 8. The time amplitudes of the seasonal empirical orthogonal functions of sea level

around the Gulf (· - - - -). Mean monthly wind stress curl C--), 1954-1974. The mean curl is -11 X 10-dynes/cm' .

Selected power spectra of the anomalies of sea level and of curl appear in Figure 7. The anomaly power at Tampico is the largest in the Gulf but all the elevation spectra are red. At periods greater than 10 mo/ cycle, power at Galveston (not shown) falls below that at Port Isabel. The anomalies of curl show a consistent increase in power at periods greater than 10 mo/ cycle.

Page 15: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

724 Journal of Marine Research [39, 4

Table I. The empirical orthogonal functions of sea level, 1954-1974, at tide gauges around the Gulf of Mexico from which the signal at St. Petersburg has been subtracted. The percentage of variance represented by the modes appears in parentheses. The seasonal functions refer to the average January, average February , etc. The anomalies refer to the differences be. tween the long.term averages at each month and the individual monthly values.

Seasonal Anomalies

1st Mode(86%) 2nd Mode(8%) !st Mode(57%) 2nd Mode(l7%)

Progreso -.50 .24 -.40 .51

Veracruz -.43 .16 - .36 .13

Tampico -.46 -.12 -.62 .23

Pt. Isabel -.45 -.88 - .39 -.29

Galveston -.26 -.61 - .33 -.66

Pensacola -.07 - .76 -.15 -.37

Key West - .22 .30 - .09 -.02

b. Empirical orthogonal functions (EOF)

In Figure 6 the seasonal variation in adjusted sea level is noticeably greater in the western and southern Gulf, e.g., Progreso and Tampico vs Pensacola and Key West. This is reflected in the first two empirical orthogonal functions (EOF) of sea level appearing in Table 1. Generally, one expects the EOF analysis, through a partitioning of the variance, to separate unique local effects from effects correlated on basin-wide scales (see Davis, 1976, for a discussion of the EOF). In Figure 8, upwards of 85 % of the seasonal (mean monthly) variance is represented by the time dependent amplitudes of the fir st seasonal EOF, to which the signals of the western and southern Gulf contribute greatly. These time amplitudes compare well with the seasonal variations in curl if one accounts for a phase lag in sea level of 1 mo. The variations in both curl and sea level imply a more intensely anticyclonic circulation in summer. The time amplitudes of the second seasonal EOF represent the semiannual signal which dominates in the northern Gulf, as found in sea level by Chew (1964) and as appears in the temperature and salinity analyses of Arm-strong (1976).

In Figure 9 the time amplitudes of the first EOF of the elevation anomalies are not significantly correlated (.06) with the anomalies of the curl. The coherency spectrum, not shown, indicates that this mode and the anomalies of curl are not coherent. It is evident that many of the events clearly appearing in the anomalies of sea level elevations are not distinguishable in the anomalies of curl. Aside from a very low frequency trend in curl much of the variability appears composed of month-to-month fluctuations. However, we do not find a corresponding fluctuation in sea level for the large event in curl that does appear during 1969-1970.

In Figure 9, the elevation anomalies at Tampico and the time amplitudes of the first EOF are quite similar. Indeed, there is significant correlation (-.31) at the

Page 16: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation

' ' ' ' '

:i~\·}i\ ',, ' 'th; , \

"') • . '

54 56 58 60 62 64 66 68 70 72 74 76

YEA R

725

Figure 9. The time amplitudes of the first EOF of anomaly elevations, see Table I. For com-parison, the anomalies of sea level and alongshore wind stress at Tampico and of curl are shown. The mean has been removed from curl. Each tick mark denotes January of that year. All serie s are low passed with a Lanczos filt er; half power point is 1/ 6 cycle/ mo.

95% level between the fir st EOF and the anomalies of alongshore wind stress near Tampico. Therefore, in the following, we have attempted to determine whether the effect of local alongshore winds could be contributing to the lack of correlation between sea level and curl.

c. Separation of .the effect of local winds

Method. At the seasonal harmonics of 12 and 6 mo/ cycle, where relatively greater variance is concentrated within a narrow frequency band, curl and local alongshore wind stress may be coherent. We have used a multiple input spectral analysis (see Bendat and Piersol, 1971) at selected gauges (Galveston, Port Isabel,

Page 17: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

726 Journal of Marine Research

I 10°1

-~_.__------+=- --~

r ~ 1o.r~-"'-~-,-------,

B:1------1'4-'--"l-1'u--11--1

I .~----+..:----\

; +-----+--------<

a

C/H CN

"'

-~-'--------+=-- -~

PT.ISABEL

,. • 1 11

b

[39, 4

Figure IO. The gain, partial coherency, and phase spectra of the response of sea level to local alongshore wind stress based upon a multiple input analysis of the anomaly series which included wind stress curl. The 95% significance level is shown for zero coherency (Groves and Hannan, 1968). The 95% confidence intervals for gain appear in Table 2; the effective frequency resolution is .067 cycle/ mo. The record lengths are determined by the lengths of the local wind series: Progreso vs Merida winds 1957-1974; Tampico vs FNOC winds (same winds as in Figure 4) 1952-1974; Port Isabel vs Brownsville winds 1961-1974; and Galveston vs ship observations (27-29N, 91-93W) 1952-1972.

Tampico, Progreso) to separate the relative contributions of the two forcing var-iables and to compute their individual frequency response functions (FRF). We apply this analysis to the anomalies of sea level, local wind stress, and curl. In a power spectrum of the full signal, the anomalies provide the background power which, unlike at the seasonal harmonics, can be greatly smoothed without severe spectral bias. To maximize the number of degrees of freedom, we have not in-

Page 18: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 727

,o .. CIHON CI HON

" 10' 10·• 10"' ul

;; ~o ,,-u

z ' _,, .:1 1:;

, .. 0

i'l -

'l"~

;a TAMPICO PROGRESO

8: u.;

"' ~-,.. f.i

!

: ,.

10' ,·· i ,.. . ' " ..

C d

creased the number of inputs beyond curl and alongshore wind stress. Limitations on record length are imposed by the local wind data which is as few as 14 years at Brownsville (Port Isabel).

As a practical matter, we generate the periodograms of the anomaly series, and then interpolate across the seasonal harmonics before smoothing with a running average over 17 bands.

Results. The partial coherency between the anomalies of curl and sea level is not different from zero at the 95% probable significance level. We conclude that at the level of anomaly power, there is variability in sea level which cannot be attributed to our input data series. Some additional forcing function is needed among the inputs, the quality of the curl data is insufficient, or there is incoherent variability in the offshore flow which reflects in the sea level elevations.

Page 19: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

728 Journal of Marine Research [39, 4

Table 2. The frequency response of sea level at selected gauges to local alongshore wind stress. The gains are estimated from Figure IO which is the result of a multiple input analysis which includes wind stress curl. The variability in the gain over frequency bands of significant coherence appears as an uncertainty adjacent to the estimated gain. The 95% confidence intervals, see Benda! and Piersol (1971), are based upon 30 degrees of freedom.

9 5 % Confidence Intervals

Estimated Gain Gain Phase cm/(dyne/ cm') cm/(dyne/cm') Degrees

Galveston 24 ± 4 ± II ± 27°

Pt. Isabel 5.5 ± .5 ± 4 ± 45•

Tampico 17 ± 4 ± 13 ± so· Progreso 30 ± 5 ± 30 ± 90°

In contrast, in Figure 10, bands of significant coherence between alongshore wind stress and sea level generally can be found among the four selected gauges. In particular, sea level at Galveston clearly shows a quasi-steady response to local winds in which the phase is nearly a constant value of 1r. Consistent with the classical Ekman model of coastal circulation, the value of 7T refers to sea level "set-up" occurring under negative wind stress. Based upon the high coherence at Galveston and the clear phase agreement with Ekman theory, we will take Galves-ton as most representative of the effect of local winds on sea level at these periods. It is probable that the difficulty in reproducing the results at Galveston at the other locations is caused by an increased independent variability in sea level or by un-representative wind data. For example, we are unable to determine whether the winds measured at Merida, some 40 km inland, are fully representative of the local wind forcing at Progreso. On the other hand, one could speculate that the increased level of anomaly power (over Galveston) indicates that a large portion of the variability in sea level is dynamically independent of the alongshore wind forcing.

In Table 2, we have selected gains averaged over the bands in which the coher-ence is relatively high. In the following, these gains are approximated as constant and are assumed to be associated with constant phase (1r or, at Progreso, zero). The variability of the gain within the coherent bands are in all cases smaller than the 95% confidence intervals shown in Table 2.

The seasonal effect of local wind forcing on sea level has been computed by multiplying the gains in Table 2 by the mean monthly values of alongshore wind stress. These computed elevations have been subtracted from the mean monthly elevations of sea level as shown in Figure 11; the differences are hereafter referred to as the residual elevations. The estimated error in the residual elevations that can be attributed to the uncertainty in the gains is 27 % at Galveston, 15 % at Tampico, and less than 10% at Port Isabel and Progreso. The residuals at Port Isabel now take on a more semiannual trend similar to that at Galveston and to the dynamic

Page 20: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981]

::; u

Blaha & Sturges: Gulf of Mexico wind-forced circulation

10 GALVESTON

8

6

4

2

0

- 2

-4

-6

-9

-10

18 TAMPICO

16

14

12

10

8

6 .. -•··

,,_

_ .. -~

..

·,

' I . 0

A / \ -2 P.' 1 'q I \f ~--4 /l \1 1, · -6 •

-9

-10

ii ··-.. ,-•-..

J M M MONTH

S N J

.30 12 PORT ISA8EL

10

.20 8

6

.10 4

2

0 0

-2

-.10

-6

..

-.20 -s •

- .25 -10

-12

-14

10

.40 8

6

.20 4

2

0 0

-2

-.20 -4

-6

-.40 -9

-10

-12

PROGRESO

M M MONTH

.8

.4

0

··-.•. -·-·~ -.4

·-. -.a

.20

.10

0

-.10

-.20

S N J

729

',e

w z >-0

Figure 11. The mean monthl y elevations of sea level (- - • - -), alongshore wind str ess ( .. . . ), and residual elevations (- - o - -). The effect of alongshore wind stress has been removed from the residual elevations using the gain values in Table 2. The mean value has been removed from the wind stress. Positi ve wind str ess is towards the north and toward s the east.

heights from region I. It is unclear why the relative summer low at Tampico dis-appears whereas at Progreso there is only minor change. Progreso retains a close similarity to curl.

Page 21: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

730 Journal of Marine Research [39, 4

The similarity between the mean monthly signals of wind stress curl and sea level provides the most supportive evidence for the notion of flow within the west-ern Gulf responding to curl. The sharp transition from summer to fall is a common feature of both variables. However, the transition in sea level, from August to

October, lags the transition in curl by about one month. If we assume that this lag is a phenomenon of the wind-driven circulation of the basin and the effective frictional decay of Rossby waves is less than a year, the observed phase lag and equation (4) can be used to compute a frictional decay time. At the annual harmonic and assuming that the forcing is zonally uniform, a 1 month phase lag corresponds to a decay time of about 33 days. However, this result is dependent upon the zonal scale of the curl at the annual harmonic which is unknown. If the zonal wavelength, k, of the forcing in equation (3) is twice the width of the Gulf, we obtain 100 days.

A computation for the lag can be found in the initial value problem in which the lag is the time needed for the barotropic circulation to " spin-up." Anderson and Gill (1975) have shown that the spin-up time of the basin is the time it takes a Rossby wave to propagate across the basin. Suppose that this time is determined by the longest barotropic free wave capable of appearing in the basin. (Baroclinic waves are not capable of traversing the Gulf in one year.) If the zonal width of the Gulf is 1200 km, the longest wavelength should be twice this dimension. In addi-tion, if the meridional wavelength is twice 600 km, we can compute a group velocity of -35 cm/ s, which over 1200 km implies a lag of 40 days. This lag is not significantly different from the observed 1 mo lag.

5, Conclusions

The mean difference of dynamic height (0/700 db) from offshore to inshore regions (III vs II) is 14 dynamic cm. In his Figure 1-38b, Nowlin (1972) shows a difference of about 20 dynamic cm at the surface relative to 1000 db. The mean anticyclonic flow which is suggested by this baroclinic shear is consistent with the observed anticyclonic curl, which we have estimated as -11 x 10-0 dyne/ cm' above the area of the gyre.

The seasonal variation in dynamic height, about 5 cm, is only a fourth the change in sea level from summer to fall . A small part of this lack of agreement may be the fault of a poor data base, but there are three kinds of alongshore flow variations that may be seen at the tide gauge: (a) large-scale, baroclinic variations that appear in the dynamic height signals, (b) large-scale barotropic variations that are seen at a nearby tide gauge but not in the dynamic heights; and (c) the nearshore flow on the shelf that responds to the local forcing of longshore winds. The data do not allow separation of the fir st two effects on the tide gauge, but we have attempted to remove the third effect. The remaining signal is large (a seasonal range of about

Page 22: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 731

I 5 cm) and its agreement with the seasonal (mean monthly) variation in curl is a new result. The similarity between the seasonal variations of sea level and curl is suggestive of a broad-scale, curl-forced circulation in the Gulf.

Questions that may challenge this interpretation remain. Why is the signal at Progreso as large as that at Tampico? Is this an indication of some flow emanating from the Yucatan Strait? Over the Cayman and Caribbean Seas the August to October transition in curl lags the transition in the Gulf by one month. Could this contribute to the observed lag in sea level in the Gulf-also one month?

The lack of coherence between the anomalies of curl and sea level may simply indicate that we have a poor data set in curl, in regard to the analysis of the anomalies. There are some additional inputs that seem appropriate, but the data are not presently available. One such set would be a time series of curl on scales smaller than can be resolved on the FNOC grid. Another would be the fluctuations in sea level difference from the Yucatan Peninsula to Cuba (Yucatan Strait) which might indicate the effect that fluctuations originating upstream of the Gulf have on the circulation in the western Gulf.

It seems appropriate to mention here the work of Elliott (1979) whose calcula-tions suggest that the work done by the annual wind stress on the western Gulf of Mexico is no more than twice the energy contained in a single ring pinched off the Loop Current. These rings are generated approximately on an annual basis (Vuk-ovitch et al., 1978) and propagate into the western Gulf. Elliott's suggestion is essentially that the western anticyclone represents an accumulation of old rings; the wind is an additional source of energy. It might be shown that each new ring drifting to the west merely displaces the remnant of the old one. However, the great reg-ularity of the observed tide gauge response, coupled with the apparent irregularity of the formation of rings (Vukovitch et al., 1978), would seem to argue against this hypothesis.

A more basic objection to the idea that the anticyclonic circulation in the western Gulf represents an accumulation of old rings, comes from our understanding of the cascades of energy and vorticity in the ocean. Rhines (1977) suggests that the cascade of energy from baroclinic eddies is such that the scale of motion, if initially larger than the Rossby radius, which the western gyre clearly is, decreases and that the motion tends toward barotropic rather than baroclinic flow. The barotropic flow of the Loop Current rings has not yet been observed. These comments are appropriate for a flat-bottom ocean model; with the inclusion of bottom topography, the scattering to smaller baroclinic scales still holds. At the western boundary, the reflection of long Rossby waves into short ones again decreases the scale. The in-fluence of both the western wall and bottom topography, therefore, suggests that the scale of the baroclinic energy contributed by pinched-off rings will decrease, rather than increase to form a baroclinic western gyre.

Page 23: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

732 Journal of Marine Research (39; 4

Finally, the contribution of wind stress curl to the average vorticity of the western Gulf is likely to be larger than the vorticity brought in by rings (about one ring every ten months). We approximate the average vorticity in a circular ring by the integral f fk • (V X v)dA = fv · di = 2Trr v(r) multiplied by the effective depth of the ring, H. We select an outer ring speed v(r = 200 km) of 4 cm/s. Because of the strong baroclinic shear in rings, the greater part of the ring circulation is contained within the upper 1000 m; we set H = 1000 m. In comparison, an average curl, of -10 x 10-• dyne/cm', is taken to act over an area in the western Gulf of 750 km by 500 km. The ratio of curl to ring vorticity over 300 days is 19.

Acknowledgments. A large number of individuals are gratefully acknowledged for their assistance in obtaining the various data. Among them we thank Dr. B. Elliott, Mr. A. Bakun, Lt Col. C. Nelson (U.S. NMFS), Drs. I. Galindo and I. Emilsson (lostituto de Geofisica, Mexico), and Mr. H. Odum (NODC). We wish to thank Dr. B. Elliott and Dr. D. Chelton for several helpful discussions. We also wish to thank Mr. P. Oppenheimer for assistance and innovative computer programs. Support has been provided under grants from the National Science Foundation (Grant OCE76-21072), the Office of Naval Research (Contract N0OOl4-7S-C.0201) and a traineeship from the Public Health Service, HEW (I T32 ES07011 TOX). Support for the fir st author under the National Research Council Associateship program and the Pacific Marine Environmental Laboratory, NOAA is also acknowledged. Contribution Number 516 from the NOAA / ERL Pacific Marine Environmental Laboratory.

APPENDIX The 4° box 25-29N, 91-95W encloses a shipping lane from which ship observations are

available every day. Within tltis 4° box there are about 95 dail y averages of wind stress per month which, over the period 1967-1972, have been averaged to 72 monthly values. At the center of the box, 27N, 93W, monthly averages of 6-hourly computations of wind stress were provided by A. Bakun. The data base was the historical FNOC sea level pressure field. The computed geostrophic winds were rot ated by 15 ° toward lower pressure and contracted by 30% . (In Figures 4 and 5 and in the main text , we have used a more recent recalculation of stress by FNOC. We have no reason to believe that it is any worse than the set presently under discussion.) For ship observations, we use Co = 1.5 X 10-1 and for computed winds Co = 1.1 X 10__,.

If a ship observation is assigned a typical uncertainty of 3 m/ s (about I Beaufort number at moderate winds), a monthly wind stress uncertainty in thi s region is ± .02 dyne/ cm\ not accounting for fair weather bias. Typical standard deviations are larger, between .04 and .10 dyne/ cm2

• Therefore, we assume that the variability in the ship observations is real. The mean errors are defined as the wind stre ss computed from atmospheric pressure minus

the wind stress from ship observations. The mean error in the meridional component is .09 ± .02 dyne/cm' . The mean error in the zonal component is -.12 ± .03 dyne/ cm'. Monthly zonal ro;tre ss is always negative, that is, toward s the west. Therefore, the wind stress computed from atmospheric pressure would tend to be too large in thi s component. The results in the merid• ional component are not clear.

There is a correlation in the magnitude of the error with season; it is lower duxing the summer when the trade winds persist than during winter when the synoptic fronts dominate

Page 24: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

1981] Blaha & Sturges: Gulf of Mexico wind-forced circulation 733

the weather. However, no statistical dependence of the error on the sample standard deviation was found.

About 45% of the monthly stresses from atmospheric pressure fall outside 2 standard de-viations of the monthly stresses from ship observations. Nevertheless, the linear regression between the two sets of wind stress shows signific ant correlations of .83 and .65 in the merid-ional and zonal components, respectively. The meridional component slope and intercept (stress from pressure as a function of str ess from ship observations) are 1.04 and - .07 dyne/cm2 and the zonal component slope and intercept are .93 and - .15 dyne/cm2

REFERENCES Anderson, D . L. T. and A. E. Gill. 1975. Spin-up of a stratified ocean, with applications to

upwelling. Deep-Sea Res., 22, 583-596. Armstrong, R. S. 1976. Hi storical physical oceanography: Seasonal cycle of temperature,

salinity, and circulation, in Environmental Studies of the South Texas Outer Continental Shelf, 1975, Vol. II , Physical Oceanography, J. W. Angelovic, ed. A report to the U.S. Bureau of Land Management prepared by NOAA, NMFS, and NOS, 290 pp.

Beardsley, R. C. and B. Butman. 1974. Circulation on the New England continental shelf: Response to strong winter storms. Geophys. Res. Letters, J, 181-184.

Beardsley, R. C., H . Mofield, M. Wimbush, C. N. Flagg and J. A. Vermersch, Jr. 1977. Ocean tides and weather-induced bottom pressure fluctuation s in the Middle-Atlantic Bight. J. Geophys. Res., 82, 3175-3182.

Benda!, J. S. and A. G. Piersol. 1971. Random Data: Analysis and Measurement Procedures. John Wiley & Sons, Inc., New York, 408 pp.

Blaha, J. P. 1978. Evidence for wind forced circulation in the Gulf of Mexico. Ph.D. Disserta-tion, Dept. of Oceanography, Florida State Univ., Tall ahassee, Florida, 134 pp.

Chew, F. 1964. Sea-level changes along the northern coast of the Gulf of Mexico. Trans. Arner. Geophys. U ., 45, 272-280.

Crepon, M. R. 1976. Sea level bottom pressure and geostropbic adjustment. Memoires Societe Royale des Sciences de Liege, 6, 43-60.

Davis, R. E. 1976. Predictability of sea surface temperature and sea level pressure anomalies over the North Paciiic Ocean. J. Phys. Oceanogr., 6, 249-266.

Ekman, V. 1905. On the influence of the earth's rotation on ocean cur rents. Arkiv. f. Matern . Astron. och Fysik, 2, 53 pp.

Elliott, B. A. 1979. Anticyclonic rings and the energetics of the circulation of the Gulf of Mexico. Ph.D. Dissertation, Texas A & M Univ., College Station, Texas, 188 pp.

Groves, G. W. and E. J. Hannan. 1968. Times series regression of sea level on weather. Re-views of Geophys., 6, 129-174.

Hastenrath, S. and P. J. Lamb. I 977. Climatic Atlas of the Tropical Atlantic and Eastern Pacific Oceans. Univ. of Wisconsin Press, Madison, Wisconsin.

HeIJerman, S. I 967. An updated estimate of the wind stress on the world ocean. Mon. Weather Rev., 95, 607; also correction, ibid 1968, 96, 63.

Kri shnamurti, T. N. and R. Kri shnamurti. 1979. Surface meteorology of the A-scale during one-hundred days of GA TE. Deep-Sea Res., Gate suppl. II to 26, 29-62.

Kundu, P. K., J. S. Allen, and R. L. Smith. 1975. Modal decomposition of the velocity field near the Oregon coast. J. Phys. Oceanogr., 5, 683-704.

Merrell, W. J., Jr. and J . M. Morrison. 1981. On the circulation of the western Gulf of Mexico with observations from April 1978. J. Geophys. Res., 86 (5), 4181-4185.

Page 25: The Journal of Marine Research is an online peer-reviewed journal …images.peabody.yale.edu/publications/jmr/jmr39-04-06.pdf · 2019-05-18 · 1981] Blaha & Sturges: Gulf of Mexico

734 Journal of Marine Research [39, 4

Nowlin, W. D., Jr. 1972. Winter circulation patterns and property distributions, in Contribu-tions on the Physical Oceanography of the Gulf of Mexico, Vol. II, L. R. A. Capurro and J. L. Reid, eds., Gulf Publishing Co., Houston, Texas, 3-51.

Nowlin, W. D., Jr. and H.J. McLellan. 1967. A characterization of the Gulf of Mexico waters in winter. J. Mar. Res., 25, 29-59.

Pedlosky, J. 1965. A study of the time dependent ocean circulation. J. Alm. Sci., 22, 267-272. Reid, I. L. and A. W. Mantyla. 1976. The effect of the geostrophic flow upon coastal sea

elevation in the northern North Pacific Ocean. J. Geophys. Res., 81, 3100-3160. Rhines, P. B. 1977. The dynamics of unsteady currents, in The Sea, Vol. 6, E. D. Goldberg,

I. N. McCave, J. J. O'Brien, and H. H. Steele, eds., Wiley lnterscience, New York, 189-318. Stammel, H. 1948. The westward intensification of wind-driven ocean currents. Trans Amer.

Geophys. U., 28, 202-206. Sturges, W. and J.P. Blaha. 1976. A western boundary current in the Gulf of Mexico. Science,

191, 367-369. Vasquez, A. M. 1975. Currents and waters of the upper 1200 m of the southwestern Gulf of

Mexico. M.S. Thesis, Texas A & M Univ., College Station, Texas. Unpublished document, 99 pp.

Vukovitch, F. M., B. W. Crissman, M. Bushnell and W. J. King. 1978. Sea-surface temperature variability analysis of potential OTEC sites utilizing satellite data: Final report: U.S. Dep~ of Energy, Div. of Solar Energy, DOE Contract No. EG-77-C-05-5444.

Received: 2S March, 1980; revised: Z Sep/ember, 1981.


Recommended