1. Introduction
An understanding of atmospheric energy and moisture transport over high latitudes is important for understanding global climate change. These studies require atmospheric wind profiles with reasonable accuracy and sufficient spatial and temporal resolutions to yield reliable flux estimates and to resolve their temporal variation. Atmospheric winds observed from aircraft and the radiosonde network, or derived from the satellite cloud-motion, do not have the spatial and temporal resolutions for accurate flux estimates. In particular, radiosonde stations are sparse over the oceans, aircraft observations are collected on an intermittent basis along fixed flight tracks and heights, and satellite cloud-drift winds only contain high- and low-level winds (e.g., Schmetz et al. 1993). For this reason, many recent energy and moisture flux studies are based on the wind fields from various analyses and reanalyses (e.g., Bromwich et al. 1995; Cullather et al. 1996, 2000; Wang and Paegle 1996; Genthon and Krinner 1998; Groves 2001). The most frequently used products are the European Centre for Medium-Range Weather Forecasts (ECMWF) analysis and reanalysis, National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalysis, and National Meteorological Center (NMC) analysis. These winds have the advantages that they include various routine observations from satellites, ships, buoys, aircraft, and the radiosonde network, etc., combined with analysis/forecast models, which also contain a comprehensive set of physical parameterizations. They are easy to use because they are in gridded formats with global coverage for extended periods of time. Unfortunately, the analyses can contain spurious climate change signals resulting from frequent changes in the analysis models. To ameliorate these effects, reanalysis products were generated using frozen, state-of-the-art analysis/forecast models (Kalnay et al. 1996; Gibson et al. 1997). These reanalysis products, because of their duration, are often used as proxies for climate change studies.
Despite the obvious benefits, studies indicate that the wind fields in these reanalyses have large uncertainties, often producing different, and sometimes even conflicting, climatologies of the heat and moisture fluxes. For instance, Wang and Paegle (1996) showed that the wind differences, measured by a nondimensional root-mean-square ratio, between the NMC and ECMWF analyses are 2 to 3 times larger than their moisture differences. This large wind difference is the main cause of the large uncertainties in their moisture flux estimates over North America and South America. Cullather et al. (1996) show a close relationship between the El Niño–Southern Oscillation (ENSO) and the Antarctic moisture flux convergence calculated using the ECMWF operational analyses as well as the NCEP–NCAR reanalyses. In contrast, Genthon and Krinner (1998) found no convincing correlation between ENSO and the ECMWF reanalysis data. Bromwich et al. (2000) show that the different climate variability of the moisture flux convergence between the ECMWF analyses and reanalyses is caused by the differences in their wind fields.
The wind differences in the various analyses can be attributed to many causes. For instance, differences noted by Wang and Paegle (1996) in the NMC and ECMWF analysis winds were attributed to different physical parameterizations used to simulate low-level jets associated with topography. On the other hand, wind field differences between the ECMWF analysis and reanalysis in Bromwich et al. (2000) can be attributed to a height error in the reanalysis assimilation of the Vostok station on the Antarctic continent. Furthermore, because of the complicated error structure resulting from assimilation of various types of observations and different model parameterization schemes, it is often difficult to identify the specific causes of the errors in the analysis/reanalysis winds. Indeed, Francis and Cermak (2001) find that the ECMWF and NCEP–NCAR reanalyses winds exhibit systematic biases of up to 3 to 4 m s−1 in the middle and upper troposphere of the Arctic region when compared to independent radiosonde observations. The reasons for these errors are not yet clear.
The above difficulties in using the analysis/reanalysis winds call for improved reanalyses. The NCEP–NCAR Reanalysis II dataset has corrected some of the known problems in the NCEP–NCAR reanalysis (Kistler et al. 2001), and may provide improved understanding of the heat and moisture flux climatology in high latitudes. On the other hand, different wind products such as the purely satellite-based geostrophic wind may be used to provide a basic constraint on, and improved understanding of, the high-latitude wind climatology. For this reason, Slonaker and Van Woert (1999) and Zou and Van Woert (2001) attempted to derive a satellite-based, geostrophic-like wind dataset for the moisture flux and net precipitation estimates over the Southern Ocean. In Slonaker and Van Woert (1999), atmospheric meridional wind profiles were derived from the satellite-based surface wind field and the satellite temperature soundings based on the thermal wind relationship. Zou and Van Woert (2001, hereafter ZVW01) extended their wind derivation algorithm by including the mass conservation in a variational procedure. The retrieved annual-mean meridional wind in ZVW01 agrees with the radiosonde observations (raob) at the Macquarie Island, Tasmania station to within ±0.5 m s−1 and its general circulation structure is also compatible with the ECMWF and NCEP–NCAR reanalyses. Compared to reanalysis systems that use primitive equation models, diverse observations, and complicated physical parameterizations, the satellite-retrieved winds have a much simpler error structure. Furthermore, satellite-derived winds are capable of reproducing the basic structure of the atmosphere with reasonable accuracy, and thus provides an attractive alternative for wind-related climate studies over the middle- and high-latitude oceans.
In ZVW01, only meridional wind profiles were derived. In this study, the wind derivation approach of the thermal wind plus mass conservation constraint developed in ZVW01 is applied to derive both zonal and meridional winds. Two forms of mass conservation constraints are presented and examined. The resultant atmospheric general circulation structure is then compared with the ECMWF and NCEP–NCAR reanalyses. To assess the performance of the algorithm, the resultant wind profiles are also compared to the raob at Macquarie Island. Through comparisons with the reanalyses and radiosonde measurements, this study attempts to provide a foundation for further application of these retrieval algorithms and the resultant wind profiles to climate diagnostic studies.
The next section describes the data used in this study. Section 3 describes the algorithms used to infer the wind field. Section 4 provides comparisons of the satellite-derived winds with the ECMWF and NCEP–NCAR reanalyses. Section 5 discusses the differences between the two mass conservation methods and section 6 provides comparisons between the satellite winds and radiosonde winds at Macquarie Island. Section 7 contains a summary.
2. Data
The satellite data used in this study are the same as in ZVW01. The surface wind field is the Special Sensor Microwave Imager–based (SSM/I) variational analysis wind obtained by Atlas et al. (1996). This surface wind is available globally every 6 h since July 1987 with a spatial resolution of 2° latitude by 2.5° longitude. This dataset uses the ECMWF 10-m analysis wind as the background and optimally blends nearly all available surface wind data, including conventional ship and buoy wind vectors, SSM/I wind speeds, and most of the Tropical Atmospheric Ocean (TAO) array of moored buoys into the background field. Over the ocean area, wind velocities are significantly influenced by the SSM/I wind speeds. The SSM/I wind speeds have an accuracy of ±2 m s−1 with better coverage and resolution than in situ measurements. Therefore, the speed and direction of the blended surface wind vectors, referred to as VAM winds, exhibit higher accuracy than the original ECMWF 10-m wind when compared against independent buoy data. This is especially true over the tropical and Northern Hemispheric oceans (Atlas et al. 1996). Near Antarctica and the Southern Ocean, the VAM winds tend to present a better-defined circulation and more smoothly varying features when compared against the Comprehensive Ocean–Atmosphere Dataset (CODAS). Only the Southern Ocean portion of the dataset is used in this study.
The temperature data were taken from the Television and Infrared Observational Satellite (TIROS) Operational Vertical Sounder (TOVS) Pathfinder Path A dataset. Susskind et al. (1997) described the detailed characteristics of this dataset. TOVS Path A data are based on the interactive physical retrieval schemes of the Goddard Laboratory for Atmosphere (GLA). In this interactive retrieval system, the retrieval subsystem depends on the first-guess field, which is provided by the 6-h forecast of the GLA general circulation model; the retrieved satellite soundings, as well as other in situ concurrent measurements, are in turn used as the initial field for the next 6-h forecast. Susskind and Pfaendtner (1989) showed a significant positive improvement in skill of forecasts (and thus the first-guess for retrievals) using this interactive forecast–retrieval–assimilation system compared to those using operationally produced satellite soundings. The retrieved TOVS Path A temperature at 500 hPa has a global bias of 0.13 K and standard deviation of 1.40 K when validated against radiosonde observations (Susskind et al. 1997).
Daily mean temperature retrievals of TOVS Path A are available globally on a 1° latitude by 1° longitude grid. The dataset contains temperature outputs at the standard pressure levels and auxiliary layer-mean virtual temperature (Susskind et al. 1997). In this study we use the layer-mean virtual temperature defined at the layers of 1000–850, 850–700, 700–500, 500–300, and 300–100 hPa. Horizontal temperature gradients, and hence the thermal wind, are calculated from these temperature data. TOVS retrievals are performed only under clear or partially cloudy (up to 80% sky cover) conditions (Susskind 1993), resulting in data-void areas for totally overcast conditions. Within individual missing data areas, the temperature gradients in the longitudinal or latitudinal direction are approximated by the gradient obtained from the two surrounding good data in that direction. ZVW01 indicated that this approximation for the unresolved overcast areas would not significantly affect the annual-mean winds, but would most likely underestimate the eddy winds.
The TOVS Pathfinder Path A daily data have been interpolated by Slonaker and Van Woert (1999) onto a 2° latitude by 2.5° longitude grid with a 6-h interval to match the VAM surface wind data. Only data for 1988 are used in this study.
3. Derivation of satellite wind
Note that in variable surface pressure conditions, the vertical integral of the divergence is related to the surface pressure tendency plus the surface pressure advection (Kasahara 1974; Haltiner and Williams 1980). The constant surface pressure assumption results in the approximate mass conservation equation (4). This approximation greatly simplifies the retrieval algorithms. However, constant surface pressure assumption also introduces errors into the retrieved wind. In particular, over the Southern Ocean area that is the focus of this study, annual-mean surface pressure ranges from 980 to 1005 hPa (Fig. 1) with small seasonal variations (±5 hPa) (Peixóto and Oort 1992). Therefore, the 1000-hPa surface pressure assumption results in a height assignment error of the surface wind of ∼20 hPa under lower surface pressure conditions. Using Eqs. (1) and (2) and the typical middle- and high-latitude values of synoptic systems, ΔT = 10 ∼ 15 K, ΔL = 1000 km, and f = 1.2 × 10−4 s−1, produces surface wind shear estimates of |∂u/∂p| = |∂υ/∂p| = 0.02 ∼ 0.04 m s−1 hPa−1. Therefore, a 20-hPa height error would translate into a wind speed error of 0.5∼0.7 m s−1. To eliminate this error, the surface wind can be converted to 1000-hPa wind by using some assumed surface wind profiles (e.g., Boutin and Etcheto 1996). Because this error does not influence our discussions on the atmospheric general circulation structure and comparisons with radiosonde observations at Macquarie Island (ZVW01), we did not perform such a conversion. However, over plateau and mountain regions, such as the Antarctic continent, surface pressure can reach as low as 600-hPa (Fig. 1). In these regions, large surface wind errors could occur as a result of the constant surface pressure assumption.
a. Deriving zonal and meridional winds separately
The previous procedure gives a complete solution for the mass-conserved meridional and zonal winds provided the temperature soundings and surface wind fields are known.
b. Deriving zonal and meridional winds simultaneously
In the actual calculation, the fast Fourier transform is first performed on C to obtain χm at each latitude and then Eq. (29) is solved for Λm. After Λm is obtained, an inverse fast Fourier transform is used to obtain the solution for λ2 at each grid point. The mass-conserved winds can be then obtained from (24) and (25).
Using the satellite layer-mean virtual temperature profiles and surface wind observations, the mass-conserved and nonmass-conserved winds are obtained at the levels of 850, 700, 500, 300, and 100 hPa. The retrieval region extends from 50° to 76°S, but this study analyzes only the Southern Ocean region from 50° to 65°S. There, the surface pressure is closer to the 1000 hPa, as discussed before. For convenience, uc1 and υc1 are used to refer to the zonal and meridional winds obtained by the first mass conservation constraints and uc2 and υc2 to the second mass conservation constraint.
4. Monthly mean zonal-mean winds
For climate studies, the satellite-derived winds need to provide a reasonable representation of the atmospheric general circulation structure and also provide reasonable accuracy compared to radiosonde observations. In spite of the limitations inherent in the ECMWF and NCEP–NCAR reanalysis winds for climate variability studies, as discussed earlier, their basic atmospheric general circulation structure over short time periods can still provide a useful reference for satellite wind comparisons. Figures 2a–d and Figs. 3a–d show the satellite-derived, monthly averaged, Southern Ocean zonal-mean zonal wind, compared with the ECMWF and NCEP–NCAR reanalyses for January and July 1988, respectively. In the figures, both the ECMWF and NCEP–NCAR reanalysis plots use only the six vertical levels that correspond to the satellite data levels for a parallel comparison. The ECMWF and NCEP–NCAR reanalyses show almost identical zonal-mean zonal winds for the 2 months. Note that the mass conservation correction to the nonmass-conserved wind ũ depends on the longitudinal derivatives of the Lagrange multipliers in both (16) and (24). Because the zonal integrals of the correction terms are zero after using the periodic boundary conditions, the zonal-means of uc1 and uc2 are both equal to the zonal-mean of ũ. Therefore, Fig. 2a and Fig. 3a represent the satellite-derived zonal-mean zonal winds for both mass-conserved and nonmass-conserved formations.
The main feature in the zonal-mean zonal wind for the 2 months selected is the seasonal march of the jet stream positions associated with the seasonal variation of the atmospheric thermal structure. A complete analysis of the Southern Hemispheric ECMWF and NCEP–NCAR reanalysis data (not shown) indicates that the Southern Hemispheric jet stream was located near 45°S with an amplitude of approximately 30 m s−1 during January 1988. The jet stream moved from about 45°S in January to about 30°S in July. Figures 2 and 3 show that the satellite-derived zonal wind basically captures the strength and position of the jet stream. In January, an almost-closed jet stream is observed near 300 hPa equatorward of 50°S with a strength of about 30 m s−1. This is similar to the jet stream position observed in the ECMWF and NCEP–NCAR reanalyses. Note that because there is no 200-hPa level in the plots, the jet stream position that generally appears near 200 hPa in January (e.g., Van Loon 1972) now appears at 300 hPa. During July, the satellite-derived zonal wind shows much weaker horizontal shear near 50°S, indicating that it is far removed from the jet stream position. This is also similar to the reanalysis data.
Figures 2d and 3d show the difference fields between the ECMWF reanalysis and the satellite-derived zonal-mean zonal wind. The differences between the satellite-derived and the NCEP–NCAR reanalyses are similar to these plots and, thus, are not shown here. The figures show that the differences are typically within 2.5 m s−1 throughout the troposphere for both months except near the tropopause. There the differences reach 5 m s−1. These values are close to the differences between the ECMWF reanalysis and an ensemble mean from 31 atmospheric general circulation models in the Atmospheric Model Intercomparison Project (AMIP) (Gates et al. 1999) over the Southern Ocean (the ensemble has a bias of 2–2.5 m s−1 compared to the ECMWF reanalysis and a standard deviation of 3–5 m s−1 in the troposphere over the Southern Ocean). This indicates that the satellite algorithms yield a zonal-mean zonal wind structure similar to the reanalyses and most of the other general circulation models.
Figures 4a–g show the satellite-derived zonal-means of
Figures 4b and 4c show that the two mass conservation schemes for the satellite wind derivation yield similar structure and magnitudes for the zonal-mean meridional winds during the selected month. The zonal-mean differences between υc1 and υc2 are less than 0.2 m s−1 for January and 0.1 m s−1 for July over the ocean region. The basic structure of the satellite meridional winds is similar to the reanalysis winds except at the top level of the model atmosphere. There, both the ECMWF and NCEP–NCAR reanalyses show weak northward zonal-mean wind between 50° and 63°S while the satellite winds are southward. ZVW01 suggest that more observations are needed to resolve this difference. In the lower troposphere, the satellite-derived thickness of the southerly wind regime near the planetary boundary layer (lower branch of the Ferrel cell) is intermediate to the ECMWF and NCEP–NCAR reanalyses. Also, the amplitudes of the satellite surface winds are intermediate to the ECMWF and NCEP–NCAR products, with peak values that are 25% weaker than the NCEP–NCAR data, but 50% stronger than the ECMWF data for both months. The amplitude differences between the satellite-derived and reanalysis winds can also be seen in the difference fields, Figs. 4f and 4g, where the differences reach 0.5 m s−1 near the ocean surface between 50° and 55°S.
It is of interest to compare the surface zonal wind differences between various products with the differences of the meridional components. Such a comparison for July 1988 is shown in Fig. 5. It is seen that the differences between the VAM, ECMWF, and NCEP–NCAR reanalyses surface winds are on the order of 0.5 to 1 m sec−1 for both the zonal and meridional components over the entire Southern Ocean. Since the magnitude of the zonal-mean meridional wind is small, these differences yield a larger relative error (50%–100%) in the meridional circulation among the various products. For the zonal component, however, the relative difference is smaller (10%–20%) when the surface wind is large (e.g., near 50°S) and larger when the surface wind is small (e.g., near 65°S). Part of these differences could reflect errors inherited in the satellite wind measurements. Specifically, the VAM wind has a bias on the order of 0.5 m sec−1 (Atlas et al. 1996). On the other hand, these differences could reflect errors in the model-simulated surface wind, since the satellite-observed surface wind was not assimilated into the ECMWF and NCEP–NCAR reanalyses. Within the framework of the middle- and high-latitude zonal-mean circulation theory, the zonal surface wind can be determined from the approximate balance between the zonal surface stress and the vertically integrated convergence of the eddy momentum flux (e.g., Green 1970). In contrast, the zonal-mean meridional circulation is determined by the imbalance between the convergence of the eddy heat and momentum transport and the atmospheric heating and friction (including the planetary boundary layer friction and cumulus friction) (Lorenz 1967; Schneider and Lindzen 1977; Held and Hou 1980). This imbalance requires a weak Ferrel cell to transport additional energy and momentum. The uncertainties in the parameterizations of some of those processes (e.g., the surface and cumulus friction) will influence the accuracy of the model-produced surface wind, especially for the zonal-mean meridional component. Indeed, errors of up to 50% to 100% exist in the zonal-mean meridional circulation among various models (e.g., Gates et al. 1999).
Despite the magnitude of the differences, all products show the same zonal-mean surface wind directions for both the zonal and meridional components over the Southern Ocean. Future assimilation of the satellite surface winds in the reanalyses would most likely reduce the magnitude of these differences.
5. Longitudinal distribution
In the previous section, it was shown that the two mass conservation schemes give the same zonal means of uc1 and uc2 and similar zonal means of υc1 and υc2. This section examines the impact of the different mass conservation schemes on the zonal distributions of the satellite winds. Again, the zonal structure of the satellite winds is first compared with reanalyses. Figures 6a–d and Figs. 7a–d show the zonal distributions of the monthly mean uc1 and uc2 compared with the ECMWF and NCEP–NCAR reanalyses at 54°S for January and July 1988. The zonal distribution of ũ is not shown because it is almost identical to uc2. Figures 6 and Fig. 7 show that both uc1 and uc2 exhibit structural variations during January and July that are similar to the ECMWF and NCEP–NCAR reanalysis data. In January, both satellite and reanalysis winds show a four-wave-like structure with four jet cores along the zonal direction near 300 hPa. The longitudinal locations of the jet cores are similar in the satellite and reanalysis winds. The satellite winds have slightly weaker (∼5 m s−1) jet amplitudes than the reanalysis data and weaker vertical shear in the lower troposphere. Near the tropopause, the satellite winds have steeper vertical shear than the reanalysis winds. In July, both satellite winds and reanalysis winds are characterized by a dominant wave number 1 structure with a strong wind regime over the Indian Ocean (around 60°E) and a weaker wind regime over the eastern South Pacific Ocean (around 100°W).
Figures 8a–e show the longitudinal distributions for January 1988 of the monthly mean
Though the two satellite winds have similar zonal structure and are also similar to the reanalysis winds, there are subtle differences between the satellite and reanalysis winds and between the satellite winds themselves. There are many reasons for the differences between the satellite and reanalysis winds. For instance, as mentioned before, the reanalyses have more sophisticated dynamics and data assimilation schemes and include other types of observations. It is not the intent of this study to analyze these differences in detail. Rather, the focus is mainly on the differences between the two satellite wind products.
Figures 9a–b show the difference fields (uc2 − uc1) for January and July 1988 at 54°S. These figures depict a phase shift between uc1 and uc2 characterized by a dominant wavenumber 1 oscillation plus a minor wavenumber 2 oscillation in the zonal direction. Fourier analysis at 500 hPa indicates that the wave 1 variance is about 66% and 88% of the total variance in January and July, respectively; wave 2 variance is about 10% of the total variance in both months (wave variance is defined as the square of a wave amplitude and the total variance is the summation of all the wave's variance). The amplitudes of the oscillation are about 1–2 m s−1 in January 1988 and reach 5–6 m s−1 in July. Similar wave structure also occurs in the meridional wind difference field (υc2 minus υc1) for January and July 1988 at 54°S as seen in Figs. 10a–b. Figure 10 shows that υc2 has a larger equatorward component than υc1 from 90°E to 180° (eastern Indian Ocean to western South Pacific Ocean) in January 1988 and from 20° to 170°E (Indian Ocean to western South Pacific Ocean) in July.
Based on the previous analysis, it is found that ũ and uc2 are essentially identical winds. Because of this, the wave structure in Fig. 9 reflects the behavior of the correction term ∂λ1/∂θ. Because there is no latitudinal boundary restriction on ∂λ1/∂θ, it yields a larger correction to ũ. This wave structure is added to the ũ field and causes differences in the wave amplitudes and phases between uc1 and uc2. For instance, Fourier analyses of 500-hPa wind fields indicate that for January, uc1 has a total variance of 13.2 m2 s−2 and wavenumber 1 variance of 6.8 m2 s−2, which are similar to the ECMWF and NCEP–NCAR reanalysis values. However, ũ and uc2 have total variances of only 4.6 m2 s−2 and a negligible wavenumber 1 variance of 0.7 m2 s−2.
The previous analysis illustrates significant differences between the two methods. The second method is essentially a boundary value problem—its zonal wind is strongly constrained by the latitudinal boundary condition. In contrast, the first method does not require the latitudinal boundary conditions. However, the procedure requires obtaining the meridional wind first and then secondly, the zonal component. Any meridional wind input can force the zonal wind to satisfy the mass conservation (5), but not all inputs give reasonable zonal winds. Errors in the input meridional wind are linearly transferred to the zonal component. In particular, assuming an error of Δυk = δυ cosθ [which must satisfy the mass flux conservation (10)] is introduced into the input meridional wind, where δυ is a constant, then the zonal wind error is found to be Δuk = (fk
Another difference between the two methods is that the first method satisfies both constraints (5) and (10), while the second satisfies (5) and its zonal integral ∂/∂φ
Though different, both methods yield reasonable atmospheric general circulation structure. Therefore, their validity for climate studies can only be assessed through validation against real observations such as radiosonde measurements.
6. Comparison at Macquarie Island
To further assess the validity of the two mass conservation methods, the satellite winds are compared with the twice-daily wind observations at Macquarie Island, Tasmania (54.5°S, 158.9°E) for 1988. ZVW01 indicated that the 1988 annual-mean surface pressure at Macquarie Island was 997 hPa with a standard deviation of 5 hPa. This is very close to the 1000-hPa surface pressure assumption in the retrieval algorithm. Therefore, Macquarie Island is a fairly ideal station to test the performance of the satellite algorithm.
a. Zonal wind
Table 3 lists the statistics of the comparisons between the raob zonal winds and the satellite winds. The satellite winds at Macquarie Island were obtained by interpolating the values from the four surrounding grid points. Table 3 shows that the comparison is most favorable at the surface where the satellite wind is taken to be the VAM wind. At the surface, the correlation between the satellite and raob winds is 0.89, the root-mean-square (rms) error is 2.6 m s−1, and the bias is 0.76 m s−1. Atlas et al. (1996) indicated that the VAM surface wind speed has a rms error <2.4 m s−1 and bias within ±0.5 m s−1 of buoy winds over most of the tropical and Northern Hemispheric oceans. The rms error and bias at Macquarie Island are consistent with their results.
From Table 3, it is seen that the differences between the raob wind and ũ are similar to the differences between the raob wind and uc2. This again shows that ũ is very similar to uc2. Above the surface, the annual-mean ũ has a typical bias of −3.5 to −4.5 m s−1 compared to the raob data and the bias of uc2 is nearly the same as ũ. These biases are larger compared to the VAM surface wind. This vertical inhomogeneity in the bias would most likely yield unrealistic vertical wind shear near the planetary boundary layer (assuming the raob data are the truth). Francis and Cermak (2001) found similar biases between the ECMWF (as well as NCEP–NCAR) reanalyses and independent radiosonde observations (except that it was positive—reanalysis westerly winds were too strong) in the middle and upper troposphere over the Arctic. In contrast, except at 100 hPa, uc1 has a bias of approximately −1 m s−1 throughout the troposphere. This bias is close to the bias range of the VAM surface wind. The surface wind bias represents the lower limit on the upper-level wind, since the upper-level wind is directly added to the surface wind.
For instantaneous winds, both uc1 and uc2 yield rms errors comparable to the standard deviations of the raob wind. This is much larger than the WMO (1996) required accuracy (the required rms accuracy is at least 5 m s−1 in the troposphere and 10 m s−1 at the tropopause). There are many reasons for the large rms errors. Some of these are the spatial and temporal differences between the satellite and raob data. In particular, the TOVS data are spatial- and temporal-averaged data with a spatial resolution of 1° × 1° and time period of 24 h, while the radiosonde measurements are point data. Kitchen (1987) showed that observations separated in space and time would result in large errors when those observations are compared with each other. For instance, two wind observations in the Northern Hemispheric midlatitudes separated by 12 h at the same location would have rms errors varying from 6 m s−1 near the surface to 16 m s−1 near 300 hPa (Kitchen 1987). Considering these factors, the rms errors in Table 3 appear to be acceptable.
The above comparisons demonstrate that uc1 and uc2 behave quite differently—uc1 has smaller biases while uc2 has better instantaneous winds. Since it is essential for the biases to be minimal for climate applications, uc1 appears to be more suitable for climate studies. In order for uc2 to be suitable for climate studies, its large bias must be removed or reduced. Since uc2 is nearly the same as ũ, improvements in uc2 must be made by improving ũ. As seen from its construction, ũ lacks ageostrophic dynamics—the only ageostrophic component comes from the surface. However, other ageostrophic winds introduced by the geostrophic-momentum approximation or the ageostrophic acceleration itself could be very important in certain situations (e.g., Shapiro and Kennedy 1981). Including other dynamic processes might be one way of reducing the bias in ũ.
b. Meridional wind
ZVW01 have compared
7. Summary
The thermal wind relationship plus a variational technique for conserving mass has been developed in this study to retrieve the three–dimensional horizontal wind field from satellite temperature soundings and the surface wind field over the middle- and high-latitude oceans. In this technique the thermal wind derived from the satellite temperature profiles plus the surface wind is used as a first-guess wind profile, then a Lagrange multiplier is introduced in a variational formalism to constrain the first-guess wind to conserve mass. Two mass conservation schemes have been proposed. One is to use the meridional mass transport conservation equation as a constraint to retrieve the meridional wind first, and then the vertically integrated mass conservation equation is used to infer the zonal wind. The meridional and zonal winds are obtained sequentially in this approach. The second scheme is to use the vertically integrated mass conservation as a constraint to retrieve the zonal and meridional winds simultaneously from the first-guess wind. Comparisons of the two wind fields with the ECMWF and NCEP–NCAR reanalyses over the Southern Ocean and radiosonde observations at Macquarie Island, Tasmania lead to the following conclusions:
Both mass conservation schemes result in significant improvements in the monthly mean meridional wind compared to
. Both υc1 and υc2 have atmospheric general circulation structure similar to the ECMWF and NCEP–NCAR reanalysis winds. However, υc1 satisfies more physical constraints than υc2 and it has a smaller bias than υc2 compared to the radiosonde observations. In addition, quantitative comparisons suggest that υc1 is closer to the reanalysis winds than υc2 (Table 2).υ̃ Because of the restriction of the latitudinal boundary conditions, the second mass conservation scheme results in uc2 being essentially identical to ũ. The first-guess wind ũ has an annual-mean bias of approximately −4 m s−1 at Macquarie Island. The first method reduces this bias to approximately −1 m s−1, while uc2 has approximately the same bias as ũ. In contrast, ũ and uc2 have smaller rms errors than uc1 compared to radiosonde data.
Moreover, ũ, uc1, and uc2 all have the same zonal-mean structure and it is similar to the ECMWF and NCEP–NCAR reanalyses. However, there is a phase shift between uc1 and uc2 in the zonal direction, caused by the zonal wave structure in the correction term, ∂λ1/∂θ. This wave structure leads to differences of uc1 and uc2 in their zonal wavenumber spectra.
The overall comparisons suggest that the monthly mean uc1 and υc1 are probably more suitable for climate studies because their biases are small.
These conclusions are based on the results over the Southern Ocean and limited radiosonde comparisons. More validations using radiosonde observations are needed to further assess the validity of each method. In particular, it would be interesting to know whether similar results can be obtained over the Northern Hemispheric oceans.
Acknowledgments
Joel Susskind and Paul Piraino provided the TOVS Path A data. Robert Atlas and Joe Ardizzone provided the SSM/I-based surface wind fields. The ECMWF analysis and reanalysis data are supplied by NCAR. Yuejian Zhu provided the NCEP–NCAR reanalysis data. Their efforts in producing these datasets are greatly appreciated. We also appreciate Dr. Milijia Zupanski and Dr. Peitao Peng for their comments and discussions on solving the Euler–Lagrange equations. Comments from the three anonymous reviewers greatly helped to improve the manuscript. This study was supported by NASA Grant W18,795.
REFERENCES
Atlas, R., R. N. Hoffman, S. C. Bloom, J. C. Jusem, and J. Ardizzone, 1996: A multi-year global surface wind velocity dataset using SSM/I wind observations. Bull. Amer. Meteor. Soc., 77 , 869–882.
Boutin, J., and J. Etcheto, 1996: Consistency of Geosat, SSM/I, and ERS-1 global surface wind speeds—Comparison with in situ data. J. Atmos. Oceanic Technol., 13 , 183–197.
Bromwich, D. H., F. M. Robasky, R. I. Cullather, and M. L. Van Woert, 1995: The atmospheric hydrologic cycle over the Southern Ocean and Antarctica from operational numerical analyses. Mon. Wea. Rev., 123 , 3518–3538.
Bromwich, D. H., A. N. Rogers, P. Kållberg, R. I. Cullather, J. W. C. White, and K. J. Kreutz, 2000: ECMWF analyses and reanalyses depiction of ENSO signal in Antarctic precipitation. J. Climate, 13 , 1406–1420.
Cullather, R. I., D. H. Bromwich, and M. L. Van Woert, 1996: Interannual variations in Antarctic precipitation related to El Niño–Southern Oscillation. J. Geophys. Res., 101 , 19109–19118.
Cullather, R. I., and M. C. Serreze, 2000: The atmospheric hydrologic cycle over the Arctic basin from reanalyses. Part I: Comparison with observations and previous studies. J. Climate, 13 , 923–937.
Daley, R., 1991: Atmospheric Data Analysis. Cambridge University Press, 457 pp.
Francis, J. A., and R. Cermak, 2001: Satellite-observed changes in energy advection within the arctic. Preprints, Sixth Conf. on Polar Meteorology and Oceanography, San Diego, CA, Amer. Meteor. Soc., 45–48.
Gates, W. L., and Coauthors. 1999: An overview of the results of the Atmospheric Model Intercomparison Project (AMIP I). Bull. Amer. Meteor. Soc., 80 , 29–55.
Genthon, C., and G. Krinner, 1998: Convergence and disposal of energy and moisture on the Antarctic polar cap from ECMWF reanalyses and forecasts. J. Climate, 11 , 1703–1716.
Gibson, R., P. Kallberg, S. Uppala, A. Hernandez, A. Nomura, and E. Serrano, 1997: ECMWF re-analysis project report series: 1. ERA description. European Centre for Medium-Range Weather Forecasts, 72 pp.
Green, J. S. A., 1970: Transfer properties of the large-scale eddies and the general circulation of the atmosphere. Quart. J. Roy. Meteor. Soc., 96 , 157–185.
Groves, D. G., 2001: The moisture budget of the Arctic atmosphere from TOVS satellite data. M. S. thesis, Dept. of Atmospheric Sciences, University of Washington, Seattle, 93 pp.
Gruber, A., and J. J. O'Brien, 1968: An objective analysis of wind data for energy budget studies. J. Appl. Meteor., 7 , 333–338.
Haltiner, G. J., and R. T. Williams, 1980: Numerical Prediction and Dynamic Meteorology. John Wiley and Sons, 477 pp.
Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37 , 515–533.
Kalnay, E., and Coauthors. 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77 , 437–471.
Kasahara, A., 1974: Various vertical coordinate systems used for numerical weather prediction. Mon. Wea. Rev., 102 , 509–522.
Kistler, R., and Coauthors. 2001: The NCEP–NCAR 50-Year Reanalysis: Monthly means CD-ROM and documentation. Bull. Amer. Meteor. Soc., 82 , 247–267.
Kitchen, M., 1987: Representativeness errors for radiosonde observations. Radiosonde Data Quality and Monitoring, ECMWF/WMO Workshop, European Centre for Medium-Range Weather Forecasts, Reading, United Kingdom, 43–55.
Lindzen, R. S., and H-L. Kuo, 1969: A reliable method for the numerical integration of a large class of ordinary and partial differential equations. Mon. Wea. Rev., 97 , 732–734.
Lorenz, E. N., 1967: The nature and theory of the general circulation of the atmosphere. WMO Publ. 218, TP115, World Meteorological Organization, 161 pp.
Oort, A. H., and J. P. Peixóto, 1983: Global angular momentum and energy balance requirements from observations. Advances in Geophysics, Vol. 25, Academic Press, 355–490.
Peixóto, J. P., and A. H. Oort, 1992: Physics of Climate. American Institute of Physics, 520 pp.
Sasaki, Y. K., P. S. Ray, J. S. Georss, and P. Soliz, 1979: Inconsistent finite differencing errors in the variational adjustment of horizontal wind components. J. Meteor. Soc. Japan, 57 , 88–92.
Schmetz, J., K. Holmlund, J. Hoffman, B. Strauss, B. Mason, V. Gaertner, A. Koch, and L. Van De Berg, 1993: Operational cloud-motion winds from Meteosat infrared images. J. Appl. Meteor., 32 , 1206–1225.
Schneider, E. K., and R. S. Lindzen, 1977: Axially symmetric steady-state models of the basic state for instability and climate studies. Part I. Linearized calculations. J. Atmos. Sci., 34 , 263–279.
Shapiro, M. A., and P. J. Kennedy, 1981: Research aircraft measurements of jet stream geostrophic and ageostrophic winds. J. Atmos. Sci., 38 , 2642–2652.
Slonaker, R. L., and M. L. Van Woert, 1999: Atmospheric moisture transport across the Southern Ocean via satellite observations. J. Geophys. Res., 104 , 9229–9249.
Susskind, J., 1993: Water vapor and temperature. Atlas of Satellite Observations Related to Global Change, R. J. Gurney, J. L. Foster, and C. L. Parkinson, Eds., Cambridge University Press, 89–128.
Susskind, J., and J. Pfaendtner, 1989: Impact of Interactive Physical Retrievals NWP. Report on the Joint ECMWF/EUMETSAT Workshop on the Use of Satellite Data in Operational Weather Prediction: 1989–1993. Vol. 1, T. Hollingsworth, Ed., ECMWF, 245–270.
Susskind, J., P. Piraino, L. Rokke, L. Iredell, and A. Mehta, 1997: Characteristics of the TOVS Path A dataset. Bull. Amer. Meteor. Soc., 78 , 1449–1472.
Van Loon, H., 1972: Wind in the Southern Hemisphere. Meteorology of the Southern Hemisphere,. Meteor. Monogr., No. 35, Amer. Meteor. Soc., 87–99.
Wang, M., and J. Paegle, 1996: Impact of analysis uncertainty upon regional atmospheric moisture flux. J. Geophys. Res., 101 , 7291–7303.
WMO, 1996: Guide to Meteorological Instruments and Methods of Observation. 6th ed. WMO-8, Secretariat of the World Meteorological Organization.
Zou, C-Z., and M. L. Van Woert, 2001: The role of conservation of mass in the satellite-derived poleward moisture transport over the Southern Oceans. J. Climate, 14 , 997–1016.
Table 1a. Avg and max differences between different zonal wind fields at 54°S for Jan 1988. The upper-right side above the diagonal line of zeros is the avg difference between the two corresponding winds indicated by the column and row titles. The lower-left side below the diagonal line of zeros is the max difference between the two corresponding winds. Units are m s−1
Table 1b. Same as in Table 1a except for Jul 1998
Table 2a. Same as in Table 1a except for the meridional wind
Table 2b. Same as in Table 2a except for Jul 1998
Statistics between raob and satellite-derived zonal winds at Macquarie Island for 1998
Same as in Table 3 except for the meridional winds