1. Introduction
Important information on the transport and overall flow of mass in the stratosphere can be obtained by analysis of long-lived chemical constituents. Two such constituents are methane (CH4) and water vapor (H2O). Methane is produced by biotic activity near the earth’s surface, is transported into the stratosphere in the Tropics, and chemically destroyed (oxidized) above 35 km. The photochemical lifetime of CH4 below 40 km is >100 days, so the stratospheric distribution is determined mainly by the circulation. Methane measurements from satellites have been used in a number of studies to deduce stratospheric circulation features (e.g., Jones and Pyle 1984; Solomon et al. 1986; Gray and Pyle 1986; Holton and Choi 1988; Stanford et al. 1993;Russell et al. 1993a; Kumer et al. 1993; Bithell et al. 1994; Schoeberl et al. 1995). Likewise, water vapor has been used to deduce the global stratospheric circulation (Brewer 1949) and examine details of tropical (Mote et al. 1995, 1996) and global (Lahoz et al. 1996a; Rosenlof et al. 1997, hereafter RO97) transport. The ability to correctly simulate CH4 and H2O is a useful benchmark for numerical models of the middle atmosphere (e.g., Prather and Remsberg 1992).
The objective of this study is to analyze a long record (1991–97) of satellite CH4 and H2O measurements in order to quantify the seasonal cycle and interannual variability in stratospheric transport. The primary data are from the Halogen Occultation Experiment (HALOE) on the Upper Atmosphere Research Satellite (UARS). To fill in polar regions not sampled by HALOE, we use CH4 measurements from the Cryogenic Limb Array Etalon Spectrometer (CLAES) and H2O data from the Microwave Limb Sounder (MLS) instruments on UARS obtained during 1992–93. These combined data provide a complete global sample of the seasonal cycles of CH4 and H2O. Interannual variations are studied based on HALOE data alone.
There are two regions where H2O data provide details of transport not seen in CH4. First, a strong seasonal cycle is imparted to H2O at the tropical tropopause from seasonal temperature variations, and this annual cycle is observed to propagate vertically in the tropical lower- middle stratosphere (Mote et al. 1995, 1996). Second, strong dehydration is observed in the Antarctic polar vortex during winter and spring (due to extreme cold temperatures), and this signal provides an opportunity to study transport across the edge of the vortex (Russell et al. 1993a; Tuck et al. 1993; Pierce et al. 1994; RO97). The seasonal cycle H2O data presented here provide novel detail and global perspective to these features.
2. Data and analyses
a. HALOE data and equivalent latitude mapping
The primary data analyzed here are HALOE V18 vertical profile measurements of CH4 and H2O. The HALOE instrument is described in Russell et al.(1993b), and the CH4 and H2O data are discussed in detail in Park et al. (1996) and Harries et al. (1996b), respectively. We use HALOE level 3a data, which are available on 16 standard UARS pressure levels spanning 100–0.32 mb (approximately 16–56 km), with a vertical spacing of about 2.5 km. The time period analyzed here covers November 1991–March 1997.
HALOE is a solar occultation instrument that obtains 15 sunrise and 15 sunset measurements per day, each near the same latitude but spaced ∼24° apart in longitude. The latitudinal sampling progresses in time, so that much of the latitude range 60°N–S is sampled in 1 month. One method to analyze the latitude–height structure of these data is to construct zonal means by simple longitudinal averaging. However, this has the disadvantage of averaging airmasses that may be physically distinct, in particular those inside versus outside the polar vortex [for cases when the vortex is elongated or not centered over the pole; see the example Plate 1 of Schoeberl et al. (1995)].
An alternative method is to average data along potential vorticity (PV) contours, which acts as an approximate vortex-following coordinate (McIntyre and Palmer 1983; Buchart and Remsberg 1986; Schoeberl et al. 1989, 1992, 1995; Norton 1994; Manney et al. 1995). For the analyses presented here we have derived PV from daily wind and temperature fields output from the United Kingdom Meteorological Office (UKMO) stratospheric data assimilation system (Swinbank and O’Neill 1994). The utility of PV versus zonal averaging is illustrated in Fig. 1, showing HALOE CH4 measurements at 10 mb during September–October 1993 plotted versus latitude and versus PV. (Here and throughout the rest of this work the PV coordinate is expressed in terms of “equivalent latitude,” i.e., the latitude of an equivalent PV distribution arranged symmetrically about the pole.) The distribution of CH4 versus latitude in Fig. 1 shows a wide scatter over high southern latitudes, due to sampling inside (low CH4) and outside (high CH4) of the polar vortex. Conversely, the CH4 shows strong correlation with PV, with pronounced separation of vortex interior data (see also Bithell et al. 1994; Schoeberl et al. 1995). In a similar manner, HALOE H2O data are found to be correlated with the polar vortex (Pierce et al. 1994; RO97) and associated PV fields.
Figure 2 shows an example of the PV mapping of HALOE CH4 for data during January 1994. The PV mapping is done on each UARS pressure level separately (one can choose to perform the PV mapping using potential temperature as a vertical coordinate, but this involves vertical interpolations before and after mapping, and gives nearly identical results to mapping on pressure levels used here). Data are analyzed between 80°N and 80°S (as available for each month) on a 4° equivalent latitude grid. There are two advantages to the PV mapping: 1) data that are physically distinct remain separated; and 2) there is an increase in the effective latitude range of the data, as measurements inside the vortexcorrespond to high equivalent latitude (note the NH data shown in Fig. 2). The equivalent latitude data (over 80°N–S) are binned into monthly samples prior to further analyses; we note these are not true monthly means, as data at different latitudes are observed at different times during each month. We require at least three data points at a particular altitude and equivalent latitude or else the corresponding monthly sample is left blank.
b. Seasonal cycle fits
Seasonal cycles are estimated from the monthly binned data by the fitting of seasonal harmonics at each (equivalent) latitude and pressure altitude. This method works well for the irregularly sampled HALOE measurements [similar analyses are used by McCormick et al. (1993)]. Examples are shown in Fig. 3 for CH4 and H2O data at 10 mb and 28°S. These show the monthly sampled equivalent latitude data (as “plus” signs), together with the harmonically varying seasonal cycles (the repeating smooth curves). The seasonal cycle at each location is determined using a time mean, plus annual and semiannual harmonics fit by least squares regression. The linear trend (determined over January 1993–December 1996) is removed prior to these seasonal fits, and it is also omitted from the interannual analyses discussed below. Note the opposing seasonality (and interannual variations) in the H2O and CH4 data in Fig. 3; this is a feature observed throughout the stratosphere in general, due to conservation of the quantity D = H2O + 2 × CH4 (as discussed above).
Over high latitudes of both hemispheres, HALOE data are not available during several months in winter (the equivalent latitude PV mapping helps fill in this void in the NH, but less so during SH winter when the vortex is typically centered over the pole). For such high (equivalent) latitudes where data are unavailable each year, the seasonal cycle is only fit over months that data are available, and the remaining months are left blank [specifically, we require at least two years of data (out of five) for each individual month]. An example is shown in Fig. 4 for CH4 data at 72°S, where no seasonal cycle is fit during April–August.
c. Addition of CLAES CH4 and MLS H2O data
In order to fill in the high latitude voids in the seasonal cycle estimates of CH4, we include CH4 observations from the CLAES instrument on UARS (Kumer et al. 1993; Roche et al. 1996). These data span a little more than one year (January 1992–April 1993), but offer more continuous sampling than the HALOE data, in particular observing the polar winter regions. The data used here are from the version 7 (V7) retrieval. The CLAES CH4 data are mapped into equivalent latitude space and monthly averaged. Direct comparison of the CLAES and HALOE data shows reasonable agreement in the latitudinal structures, although there is an offset (bias), with CLAES approximately 20% higher than HALOE (as discussed in Roche et al. 1996). We remove this bias for the analyses here by simply adjusting the CLAES data (using a latitude-independent multiplicative factor) to match the HALOE measurements over a region adjacent to the missing HALOE data. This matching is done using a seasonal cycle fit of the CLAES data (over 1992–93) together with the seasonal cycle HALOE data (1991–97), in order to obtain smooth transitions. The adjusted CLAES data and fit seasonal cycle are included in the time series at 72°S shown in Fig. 4. Note the CLAES data are used only in the seasonal cycle estimates, since the short CLAES record precludes interannual analyses.
In a similar manner, we augment the HALOE seasonal cycle H2O results using MLS data in polar regions. Stratospheric H2O data are derived from MLS 183-GHz measurements, as described in Lahoz et al. (1996b). We use the version 4 (V4) retrieval for the analyses here, with time period spanning November 1991–April 1993. The seasonal cycle fit of the equivalent latitude MLS data are adjusted (by a small amount, ∼5%) to smoothly match the HALOE seasonal cycle fit.
d. Calculation of residual mean circulation
3. Seasonal variability
The overall temporal variance of the HALOE CH4 data during 1991–97 is shown in Fig. 5, along with the component attributable to the harmonic seasonal cycle, and the residual (i.e., that attributable to interannual variance). Only HALOE data are used in these calculations, which span 60°N–S. Variance in all components peaks in the middle-upper stratosphere (over 30–50 km) over low latitudes. The overall variance is split fairly evenly between the seasonal and interannual components, but the spatial patterns are distinct (note the seasonal variance has a minimum over the equator, while the interannual variance has a maximum there). A majority of the interannual variance in Fig. 5 is associated with the QBO, as discussed below. Variance patterns in H2O data (not shown) reveal upper stratospheric maxima similar to that in Fig. 5, with additional maxima in the Tropics over 20–30 km due to both seasonal and interannual components.
a. Global seasonality
Figure 6 shows the global structure of CH4 in January, April, July, and October, derived from the seasonal cycle fits. The overall structures are similar to those shown previously (e.g., Luo et al. 1995; Park et al. 1996), but the synthesis of many years of data here allows the seasonality to be studied in some detail. Contours bulge upward in the Tropics and move latitudinally with season in the upper stratosphere, reflecting variations in the upward transport circulation in the Tropics (as shown below). There is a double-peaked pattern in the upper stratosphere in April related to the seasonal subtropical upwelling and thetropical semiannual oscillation (SAO); a similar double-peaked structure is not seen in October (the SAO is discussed in more detail below). Relatively strong horizontal gradients are seen in the subtropical lower stratosphere near 20°–30° N–S throughout much of the year, and strong gradients are also associated with the polar vortex in both hemispheres during winter–spring. These regions of enhanced gradients are associated with differential vertical advection (upward in the Tropics and downward in polar winter) and relative minima in horizontal mixing (so-called mixing barriers). The SH winter–spring polar vortex is particularly robust in these data, due to strong downward flow inside the vortex and small transport across the vortex edge (Russell et al. 1993a; Schoeberl et al. 1995; Fisher and O’Neill 1993; Manney et al. 1994; Sutton 1994). In contrast, the CH4 contours are relatively flat over midlatitudes during winter–spring, showing evidence of more rapid mixing associated with the planetary waves (McIntyre and Palmer 1983, 1984; Leovy et al. 1985). There are relatively weak vertical gradients in high latitudes over 20–30 km during SH summer (January), possibly a remnant of very low (vortex) CH4 following the SH final warming.
The corresponding seasonal cycle patterns in H2O are shown in Fig. 7. The overall patterns are approximate mirror images of the CH4 data in Fig. 6, including the seasonal movement of the upper-stratospheric maxima and appearance of a double peak in April. These similar patterns reflect the fact that the H2O increase with altitude is due to CH4 oxidation, so that total hydrogen is conserved (see references in introduction). This is demonstrated with the data here by calculation of the quantity 2CH4 + H2O, shown for the October distribution in Fig. 8. Note the contour interval is identical between Figs. 7 and 8, so that the relative lack of spatial structure in 2CH4 + H2O (cf. H2O or CH4 alone) demonstrates that to first order 2CH4 + H2O ≈ constant for these seasonal cycles. Two locations where this approximation does not holdare 1) in the tropical lower stratosphere, and 2) over Antarctica in winter–spring, as discussed in the introduction.
The latitude–time evolution of CH4 at 68, 10, and 2.2 mb is shown in Fig. 9. There is a slight latitudinal shifting of the tropical maximum region in the lower stratosphere (68 mb) toward the respective summer hemisphere. Low CH4 is observed in the springtime high latitudes of both hemispheres; these low polar values in the NH have received less attention than those in the SH (e.g., Russell et al. 1993a; Schoeberl et al. 1995), but appear very similar in these equivalent latitude data. These overall variations of CH4 are in reasonable agreement with seasonality in the mean vertical circulation at 68 mb shown in Fig. 10 (calculated as discussed in section 2d). The patterns of high tropical CH4 follow the upward velocity variations over low latitudes; note the weak tropical maximum during NH winter and the similar latitudinal movement of the contours (the
The CH4 seasonal cycle in the middle stratosphere (10 mb) shows weakening of the midlatitude gradients during winter–spring in both hemispheres; this effect is more obvious in the SH. Similar midlatitude evolution at 10 mb is seen in the seasonal variation of nitrous oxide (N2O) measured by CLAES (Randel et al. 1994). At 10 mb strong CH4 gradients are seen across both polar vortices beginning in autumn, and the larger and stronger SH vortex is echoed in these CH4 measurements.
Seasonality in the upper stratosphere (2.2 mb) is very different from the lower levels, showing a strong semiannual oscillation (SAO) with distinct maxima in the respective summer subtropics. These maxima are very similar to the patterns of vertical velocity at 2.2 mb (shown in Fig. 10), with the CH4 maxima lagging theupward motion by approximately 1 month. A double- peaked latitudinal structure in CH4 is seen over April–June.
Figure 11 shows the altitude–time variation of CH4 at 76°N and S; note the time axes have been shifted in order to directly compare midwinter variations over both poles. The overall evolution is similar between hemispheres, with regular downward movement of the isolines during autumn and winter, followed by rapid elevation of the contours in spring (associated with breakup of the polar vortices and mixing in of midlatitude air). Somewhat lower CH4 values are observed in the SH middle and upper stratosphere in winter, and the springtime transition is more dramatic and somewhat later in the SH; these aspects are due to the larger and more isolated SH polar vortex. The 0.3–0.8 isolines show consistent downward movement in the SH middle stratosphere during autumn and winter with a rate of −1.0 to −1.5 km month−1; there is little downwardmovement in the lower stratosphere until spring (September–October), when values near −1.5 km month−1 (−0.6 mm s−1) are inferred. These latter values are in excellent agreement with the downward velocities derived from the HALOE CH4 data by Schoeberl et al. (1995), and with the TEM vertical velocities shown in Fig. 10 (see also Rosenfield et al. 1994). The NH middle stratosphere data (e.g., the 0.5 contour) yields descent rates near −1.5 km month−1, similar to the SH. In the NH lower stratosphere the data in Fig. 11 shows a relative maximum in November–December (see also the lower panel of Fig. 9) that is inconsistent with the downward circulation in Fig. 10. This isolated feature is entirely due to the use of CLAES data during one year (1992), and its reality is suspect.
Both NH and SH data in Fig. 11 show CH4 minima (below 0.1 ppmv) near the stratopause during summer (see also the top panel in Fig. 9). These minima are due to rapid photochemical destruction of CH4 over polar regions during summer (see Fig. 17 below); the calculated photochemical lifetime of CH4 in this region is near 1 month [see Fig. 1 of Solomon et al. (1986)].
b. Tropical variations
The altitude–time variations of both CH4 and H2O over the equator are shown in Fig. 14, which also includes diagrams with the respective time means removed at each altitude to highlight seasonality. There is a strong upward propagating annual variation in H2O in the lower stratosphere, associated with the seasonal variation of tropical tropopause temperatures and the freezing out of H2O in the tropical upwelling region [see extensive discussions in Mote et al. (1995) and (1996)]. A nearly identical pattern is seen in the total hydrogen 2CH4 + H2O, as illustrated in a different manner in Fig. 15. Here the individual zonal means over 4° N–S are plotted, along with the associated harmonic fits,for pressure levels 100–10 mb (16–32 km). A strong annual cycle that progresses upward with time is observed over 100–32 mb; this cycle can be traced to 22 mb using the harmonic analysis, although there is substantially more interannual variability at this level. At 15 and 10 mb (29–32 km), the annual harmonic is relatively weak, and a semiannual component dominates at these levels (and above). The upward propagation of the annual signal over 100–32 mb (16–24 km) takes approximately 11 months; this corresponds to a vertical velocity of ∼0.7 km month−1 (∼0.25 mm s−1), in reasonable agreement with the tropical TEM vertical velocities in Fig. 10.
Figure 16 shows the space–time structure of both H2O and 2CH4 + H2O at 100, 68, and 46 mb. While the annual cycles at 68 and 46 mb are approximately centered over the equator, variations at 100 mb are maximum near 15°–30°N. Lowest values are observed near 15°N in February–March, although the minimum 100- mb temperatures occur near the equator (e.g., Newell and Gould-Stewart 1981; Tuck et al. 1993). The maximum during NH summer (August–September) is even more asymmetrical, with maximum near 20°–30°N; this structure has little relation to zonal-mean temperatures. There is also a seasonal variation in meridional transport at 100 mb inferred in Fig. 16, with relatively little spreading of the low subtropical values into the NH during NH winter (note the strong gradients near 30°N), but stronger propagation of the NH summer maximum into high latitudes (with a time lag of order three months between 30° and 60°N). There is much less spreading of this maximum into SH midlatitudes. Based on analyses of these same HALOE H2O data, RO97 discuss the asymmetry seen in Fig. 16 and note the meridional propagation of maximum values into NH midlatitudes. The asymmetry of these patterns (particularly the NH summer maximum) contribute to the higher values of H2O observed in the NH versus SH lower stratosphere, as reported by Kelly et al. (1990). These data suggest the NH–SH asymmetry in the lower stratosphere is notdue to asymmetric transport of equatorially centered maxima, but to dynamical processes in the NH subtropics (such as the Indian monsoon during NH summer). The enhanced meridional speading of the H2O and 2CH4 + H2O at 100 mb compared to levels at and above 68 mb agrees with the results of Trepte et al. (1993), who present evidence for rapid meridional flow between Tropics and midlatitudes below 20 km, but not above this altitude. Similar results were inferred from the aerosol observations in Grant et al. (1994). This behavior is consistent with the tropical vertical velocity profile discussed by RO97: there is a minimum in
The H2O data at 68 and 46 mb in Fig. 16 show minimum values in the Tropics and relatively strong horizontal gradients near both 30°N and 30°S, similar (but opposite) to the CH4 latitudinal structure seen in Fig. 9. The lack of such gradients in the quantity 2CH4 + H2O (i.e., the 46-mb plot in Fig. 16) implies that it is the downward transport of photochemically aged air in the extratropics, plus a lack of rapid meridional mixing with the Tropics, that is responsible for these latitudinal H2O gradients. Seasonal variations in both H2O and 2CH4 + H2O show that the upward propagating tropical annual cycle is confined to ∼30°N–S.
The upper stratosphere signal in Fig. 14 is dominated by a semiannual oscillation (SAO), with a much stronger amplitude during the first half of the year. This is a well- known aspect of the SAO (e.g., Dunkerton and Delisi 1988), related to stronger extratropical wave driving in the NH winter (as compared to the SH winter). Variations in H2O echo those in CH4, such that the total hydrogen 2CH4 + H2O is nearly constant throughout the seasonal cycle. Dynamics of the SAO is discussed further in the next section.
c. Global budgets and estimates of eddy transport
Figure 17 shows the structure of the individual terms ∂
Figure 18 shows altitude–time plots of several CH4 seasonal budget terms over the equator [−
4. Interannual variability and the QBO
a. Observations
This section focuses on interannual variations in CH4 and H2O over 1991–97, defined as deviations from the seasonal cycles defined above (with linear trends also removed). Figure 19 shows altitude–time sections of CH4 anomalies over the equator (average over 10°N–10°S); here data gaps have been interpolated and the monthly anomalies smoothed slightly in time. Largest CH4 anomalies are observed in the upper stratosphere, over 35–45 km, with an approximate 2-yr periodicity; these values represent approximately ±10% deviations from the time mean. The approximate 2-yr periodicity in CH4 anomalies suggests a link with the equatorial quasibiennial oscillation (QBO), and this is confirmed by comparison with the UKMO equatorial zonal wind anomalies shown in Fig. 20a (with the CH4 anomalies superimposed). Specifically, the CH4 anomalies over 35–45 km are approximately in phase with the zonal wind anomalies near 30 km.
Figure 21 shows the H2O interannual anomalies over the equator. Maxima are observed in the upper stratosphere with very similar (but oppositely signed) patterns to the CH4 anomalies, reflecting conservation of 2CH4 + H2O on interannual (as well as seasonal) timescales. There are also significant H2O anomalies in the lower and middle stratosphere; there are relative maxima over 27–33 km with an approximate 2-yr periodicity, but these are not observed in the CH4 data.
Space–time structure of coherent QBO variations in CH4 may be isolated in an optimal manner using singular value decomposition (SVD) (Randel and Wu 1996). This involves manipulation of the covariance matrix between CH4 anomalies (as a function of latitude and altitude) and zonal wind anomalies (over 10–70 mb), and yields modal spatial structures that explain optimum fractions of covariance between these fields. Figure 22 shows the spatial structure of the first two modes derived from this analysis (termed SVD1 and SVD2); patterns show the location (and phase) of anomalies in CH4 that are coherent with the QBO zonal windanomalies. (Two such modes are typical for a propagating oscillation, analogous to a harmonic sine and cosine Fourier decomposition.) The spatial structure of SVD1 shows a maximum in the upper stratosphere over 35–45 km, centered in the Tropics over approximately 30°N–S. This structure corresponds to the equatorial anomalies seen in Fig. 19. The structure of SVD2 shows a narrow maximum over the equator at 40–45 km, and several maxima over ∼10°–40° in both hemispheres [note these modes are spatially (and temporally) orthogonal by construction]. These extratropical patterns are qualitatively similar to QBO variations observed in column and profile ozone and NO2 data (Bowman 1989;Randel and Wu 1996), and deduced from CLAES N2O data by O’Sullivan and Dunkerton (1997). The QBO patterns in Fig. 22 are largest near gradients in the background CH4 structure (included in Fig. 22), as anticipated for anomalies due to transport variations. Note the similarity of these QBO patterns with the total interannual variance map shown in Fig. 5.
Temporal evolution of the QBO patterns in CH4 is revealed by projection of the full monthly CH4 anomalies onto the spatial patterns in Fig. 22. Figure 23 shows a phase-space plot of the projections of SVD1 versus SVD2 over the entire time record 1991–97. Nearly regular counterclockwise phase progression is observed as the QBO anomalies project successively onto the patterns in Fig. 22. This nearly regular QBO phase variation in CH4 data is analogous to the harmonic QBO variations in dynamical quantities discussed by Wallace et al. (1993), and the ozone variations shown in Randel and Wu (1996).
Figure 24 shows CH4 data observed during April for four years (1993–96); these time periods are indicated in Fig. 23 and correspond to opposing projections of SVD1 and SVD2. These plots demonstrate that there is a strong effect of the QBO in modulating the double- peaked structure in upper-stratospheric CH4: data during April 1993 and 1995 (westerly QBO winds in Fig. 20a) exhibit a pronounced double peak, whereas April 1994 and 1996 (easterly QBO winds) do not.
Figure 25 shows latitude–time series at 3 mb of the full CH4 field and the detrended anomalies (which exhibit a clear QBO dependence). Time periods of enhanced double-peak structure are also indicated in Fig. 25. The QBO anomalies show maxima centered near the equator, with a suggestion of propagation into the subtropics, particularly into the NH. This latitudinal propagation near 3 mb is associated with successive projections of SVD1 and SVD2 (Fig. 22). The timing of these anomalies is such that large NH subtropical maxima occur during NH winter–spring; this is the time of seasonal maximum in SH subtropics, and this combination leads to double-peaked structure for the positive QBO anomalies in NH spring 1993 and 1995. The details of this behavior clearly depend on the phase relationship between the QBO and the seasonal cycle, and this is something that will change on a decadal timescale (Gray and Dunkerton 1990). We note that relatively small anomalies are observed over the equator during these times; the double-peak structure (or its absence) in this time sample is mostly due to the presence of subtropical anomalies, particularly in the NH (corresponding to a positive projection of SVD1 + SVD2, i.e., Fig. 23).
Figure 26 shows the latitude–time evolution of CH4 at 10 mb throughout 1991–97, for both the full-field and interannual anomalies. The anomalies at this altitude are very small at the equator but large over the subtropics of each hemisphere (somewhat larger in the NH), with a strong QBO time dependence. These interannualanomalies are associated with a latitudinal shifting of the tropical maximum region in CH4, as seen in the upper panel of Fig. 26. The maximum is shifted toward the NH during QBO easterlies (near 30 mb) enhancing NH subtropical gradients, and toward the SH for QBO westerlies (decreasing these gradients). Similar CH4 patterns are observed over approximately 22–10 mb (27–32 km), but not below 30 mb; these anomalies are associated with strong projection onto SVD2 in Fig. 22. Note the clear effect of these anomalies in the cross sections in Fig. 24, especially on the vertical gradient structure in NH midlatitudes. This is remarkable evidence of QBO modulation of the latitudinal position of the tropical maximum region (Trepte and Hitchman 1992) and the interaction with extratropics.
Figure 27 shows the latitudinal structure of H2O at 10 mb throughout 1991–97, for both the full field and the interannual anomalies. Large anomalies are observed in the NH subtropics, mirroring the CH4 patterns in Fig. 26. Unlike the CH4 data, the H2O anomalies extend across the equator, giving rise to the equatorial anomalies centered near 10 mb seen in Fig. 21. The overall structure of H2O at 10 mb shows latitudinal movement of the tropical minimum region and modulation of subtropical gradients on the QBO timescale, very similar to that seen for CH4 in Fig. 26.
b. QBO budget calculations
The observed QBO anomalies in CH4 and H2O likely result from transport variations associated with the QBO. Although a detailed budget analysis of QBO transport is beyond the scope of this work, there are some aspects of coupling with QBO circulation anomalies that deserve mention. Figure 20b shows an altitude–time section of the QBO anomalies in residual mean vertical velocity (
The observed space–time coherence between QBO anomalies in
Figure 28 shows a latitude–time diagram of the QBO anomalies in
5. Summary and discussion
High quality observations of stratospheric CH4 and H2O from HALOE provide a near-continuous global record of seasonal and interannual variations in stratospheric transport. Motivated by the observed strong correlation between these tracers and derived potential vorticity (PV) fields, we have mapped the HALOE data into equivalent latitude(PV) coordinates. This has the dual advantages of separating physically distinct measurements and increasing the effective latitude sampling of the data (see Fig. 2). Seasonal cycles are then constructed by harmonic regression analysis of monthly binned data. To obtain complete global seasonal cycles, we have included CLAES CH4 and MLS H2O data over polar winter regions (with smooth transitions obtained by matching the data over midlatitudes). These global seasonal cycle data allow novel study of details in the seasonal cycle of stratospheric transport. These data will also be useful for initializing or studying transport characteristics in stratospheric models.1
The seasonal cycle of CH4 in the lower stratosphere shows characteristics that have similar space–time structure to vertical velocity fields derived from meteorological analyses. This is a reasonable finding, as CH4 is strongly vertically stratified. The enhanced subtropical gradients observed in CH4 coincide with the region of weak vertical motions, that is, where the tropical upwelling changes to middle–high-latitude downward motion. Furthermore, a seasonal latitudinal shifting of both the tropical CH4 and vertical velocity fields is observed, with isolines displaced slightly into the respective summer hemispheres. This behavior suggests that tropical upwelling plays a large role in formation of the subtropical gradients for tropospheric source gases; winter hemisphere wave activity is also likely to be important (Polvani et al. 1995; Waugh 1996).
Polar regions in the lower stratosphere of both hemispheres show very similar seasonality in CH4, with very low values during spring (March–April in the NH and October–November in the SH). Such low values in the SH vortex have been discussed by Russell et al. (1993a) and Schoeberl et al. (1995) and provide evidence of relatively unmixed descent from high altitudes. The similar low values observed in the NH data here are due to having organized the observations according to the polar vortex structure (i.e., zonal means do not show such low values). Good agreement with vertical transport calculations (Fig. 12) confirms isolation of the SH polar vortex and demonstrates that the vertical velocities (derived primarily from radiative heating codes) are reasonably correct. Poor agreement in the NH (Fig. 13) suggests less isolation of the NH polar vortex.
The harmonic seasonal cycle fits of the HALOE watervapor data provide a clear view of the upward-propagating annual cycle in the Tropics (Figs. 15–17). The annual cycle is dominant over 100–22 mb (up to ∼27 km); above this altitude the seasonal variation is primarily semiannual. The lack of propagation of the annual cycle above 27 km suggests there is somewhat stronger horizontal mixing above this altitude, with less isolation of tropical air (i.e., this is the upper level of the tropical “reservoir”). This is consistent with enhanced tropical mixing above 10 mb noted by Dunkerton and O’Sullivan (1996). Two mechanisms for enhanced mixing with a semiannual timescale are 1) mean circulation advection, and 2) the tropical extension of winter hemisphere planetary wave effects.
The annual cycle in water vapor at and above 68 mb is centered over the equator, whereas seasonality at 100mb shows relative maxima in the NH subtropics. Regions of high H2O at 100 mb during NH summer are particularly shifted toward the NH, and there is evidence of strong transport into NH middle and high latitudes (Fig. 16). The asymmetry of these patterns (both in location of the tropical extrema and the latitudinal propagation) contributes to a significantly wetter lower stratosphere in the NH compared to the SH (Kelly et al. 1990; RO97). There is little meridional spreading of the tropical annual cycle above 68 mb beyond 30°N–S; the slow weakening of the tropical annual cycle with altitude over 68–22 mb is further evidence of tropical isolation, and these data infer an exchange time of order 15 months for mixing of midlatitude air into the Tropics. These data also demonstrate that there is little influence of Antarctic dehydration in SH midlatitudes following vortex breakup (as also noted from the analyses of RO97).
The data here show strong seasonality in the tropical upper stratosphere, for both CH4 and H2O, with the characteristics of a semiannual cycle forced by alternating upwelling in the respective summer subtropics. This semiannual signal in H2O was also discussed by McCormick et al. (1993) and Eluszkiewicz et al. (1996). A double-peaked latitudinal structure is evident in the climatology during April–June, but this is strongly modulated by the phase of the QBO.
A global budget analysis of the seasonal cycle CH4 data using UKMO residual mean circulation estimates and parameterized photochemical loss rates showed large residuals, which we attribute (at least in part) to eddy transport effects. The patterns of these eddy transport tendencies show a north–south dipole pattern, with positive tendencies in the winter stratosphere and negative tendencies in low latitudes (extending across theequator). These eddy transports, which are derived as a residual here, are very similar to the explicit eddy transport calculations performed using general circulation model output in Randel et al. (1994). Other studies have also suggested that the effects of planetary waves reach low latitudes (Polvani et al. 1995; Waugh et al. 1996; O’Sullivan 1997). There is a clear semiannual variation of the eddy transports (residuals) in the Tropics, and this is consistent with the amplification of large-scale wave transports in the respective winter hemispheres.
A number of model and diagnostic studies have been used in an attempt to explain the double-peaked constituent patterns in the tropical upper stratosphere (e.g., the April patterns in Figs. 6 and 7) (Gray and Pyle 1986, 1987; Hamilton and Mahlman 1988; Choi and Holton 1991; Sassi et al. 1993; Kennaugh et al. 1997). The mechanism for the double peak in these studies is downward motion over the equator (associated with the westerly shear zone of the dynamical SAO during equinox) coupled with upward motion in subtropics. A common feature of these studies, however, is an underestimation of the strength of the double peak, which is usually attributed to a lack of resolved downward motion over the equator. However, the balance of terms in the budget analyses here suggests that the vertical mean circulation does not primarily determine the tropical behavior. Transport by the
A majority of the interannual variance over the 1991–97 time sample analyzed here is found to be coherent with the QBO; aspects of QBO variations in HALOE CH4 have also been shown in Luo et al. (1997) and Ruth et al. (1997). The structure of the QBO signal in CH4 exhibits a maximum in the tropical upper stratosphere over 35–45 km. This is somewhat higher than the QBO signals observed in ozone (∼20–38 km) and nitrogen dioxide (∼28–35 km) (Zawodny and McCormick 1991). These differing vertical structures are attributable to differences in the respective background means (e.g., Chipperfield and Gray 1992); the strongest background gradients in CH4 are observed over 35–45 km.
We find that the QBO exerts a strong influence on appearance of the double-peaked “rabbit ears” structure in the tropical upper stratosphere during NH spring: the double peak occurs primarily when QBO westerlies are present near 10 mb. The double-peaked structure in the mean seasonal cycle (Figs. 6 and 7) represents a time average over both phases of the QBO. This has implications for numerical simulation of the stratopause SAO and its effects on tracer distributions (e.g., Sassi et al.1993), suggesting that details of the winds in the lower stratosphere may be important. Kennaugh et al. (1997) have recently presented a modeling study of this QBO influence on the double peak.
The observed equatorial CH4 (and H2O) QBO anomalies over 35–45 km are observed to be strongly correlated with anomalies in the residual mean vertical circulation
These constituent data also show evidence of a latitudinal shifting of the tropical maximum region and modulation of subtropical gradients in the middle stratosphere (∼27–32 km) associated with the QBO. There is a shift of approximately 10°–15° lat correlated with the QBO winds near 30 mb, with movement toward the NH during QBO westerlies. The sharpening of the NH subtropical gradients during QBO easterlies is similar to the patterns observed in stratospheric aerosol by Hitchman et al. (1994) and Grant et al. (1996). Examination of QBO anomalies in
Acknowledgments
Much of the work reported here was completed while WJR was on sabbatical at the Cooperative Research Center for Southern Hemisphere Meteorology at Monash University, Australia. We thank Tim Hall and Darryn Waugh for many useful discussions throughout the course of this work, particularly regarding tropical transport. We thank K.-K. Tung for providing the radiative heating code, and Barbara Naujokat for updated QBO winds. Dan Packman and Charles Cavanaugh managed the UARS data archive atNCAR. We thank Rolando Garcia, Boris Khattatov, and Mark Schoeberl for discussions and comments on the manuscript and Lesley Gray and Phil Mote for constructive and helpful reviews. Marilena Stone expertly prepared the manuscript. This work was partially supported under NASA Grants W-18181 and W-16215. The National Center for Atmospheric Research is sponsored by the National Science Foundation.
REFERENCES
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489, pp.
Bacmeister, J. T., M. R. Schoeberl, M. E. Summers, J. R. Rosenfield, and X. Zhu, 1995: Descent of long-lived trace gasses into the winter polar vortex. J. Geophys. Res.,100, 11669–11684.
Baldwin, M. P., and T. J. Dunkerton, 1991: Quasi-biennial oscillation above 10 mb. Geophys. Res. Lett.,18, 1205–1208.
Bithell, M., L. J. Gray, J. E. Harries, J. M. Russell, and A. F. Tuck, 1994: On the synoptic interpretation of HALOE measurements using PV analysis. J. Atmos. Sci.,51, 2942–2956.
Bowman, K. P., 1989: Global patterns of the quasi-biennial oscillation in total ozone. J. Atmos. Sci.,46, 3328–3343.
——, 1993: Large-scale isentropic mixing properties of the Antarctic polar vortex from analyzed winds. J. Geophys. Res.,98, 23013–23027.
Brewer, A. W., 1949: Evidence for a world circulation provided by measurements of helium and water vapor distribution in the stratosphere. Quart. J. Roy. Meteor. Soc.,75, 351–363.
Buchart, N., and E. E. Remsberg, 1986: The area of the stratospheric polar vortex as a diagnostic for tracer transport on isentropic surfaces. J. Atmos. Sci.,43, 1379–1405.
Chen, P., J. R. Holton, A. O’Neill, and R. Swinbank, 1994: Quasi- horizontal transport and mixing in the Antarctic stratosphere. J. Geophys. Res.,99, 16851–16855.
Chipperfield, M. P., and L. J. Gray, 1992: Two dimensional model studies of the interannual variability of trace gasses in the middle atmosphere. J. Geophys. Res.,97, 5963–5980.
Choi, W.-K., and J. R. Holton, 1991: Transport of N2O in the stratosphere related to the equatorial semiannual oscillation. J. Geophys. Res.,96, 22543–22557.
Dessler, A. E., E. M. Weinstock, E. J. Hintsa, J. G. Anderson, C. R. Webster, R. D. May, J. W. Elkins, and G. S. Dutton, 1994: An examination of the total hydrogen budget of the lower stratosphere. Geophys. Res. Lett.,21, 2563–2566.
Dunkerton, T. J., and D. P. Delisi, 1988: Seasonal variation of the semiannual oscillation. J. Atmos. Sci.,45, 2772–2787.
——, and M. P. Baldwin, 1991: Quasi-biennial modulation of planetary wave fluxes in the Northern Hemisphere winter. J. Atmos. Sci.,48, 1043–1061.
——, and D. O’Sullivan, 1996: Mixing zone in the tropical stratosphere above 10 mb. Geophys. Res. Lett.,23, 2497–2500.
Eluszkiewicz, J., and Coauthors, 1996: Residual circulation in the stratosphere and lower mesosphere as diagnosed from Microwave Limb Sounder data. J. Atmos. Sci.,53, 217–240.
Fisher, M., and A. O’Neill, 1993: Rapid descent of mesospheric air into the stratospheric polar vortex. Geophys. Res. Lett.,20, 1267–1270.
Garcia, R. R., and S. Solomon, 1983: A numerical model of the zonally averaged dynamical and chemical structure of the middle atmosphere. J. Geophys. Res.,88, 1379–1400.
——, and ——, 1994: A new numerical model of the middle atmosphere 2. Ozone and related species. J. Geophys. Res.,99, 12937–12951.
Gille, J. C., L. V. Lyjak, and A. K. Smith, 1987: The global residual mean circulation in the middle atmosphere for the northern winter period. J. Atmos. Sci.,44, 1437–1452.
Grant, W. B., and Coauthors, 1994: Aerosol-associated changes intropical stratospheric ozone following the eruption of Mount Pinatubo. J. Geophys. Res.,99, 8197–8211.
——, E. V. Browell, C. S. Long, L. L. Stowe, R. G. Grainger, and A. Lambert, 1996: Use of volcanic aerosols to study the tropical stratospheric reservoir. J. Geophys. Res.,101, 3973–3988.
Gray, L. J., and J. A. Pyle, 1986: The semi-annual oscillation and equatorial tracer distributions. Quart. J. Roy. Meteor. Soc.,112, 387–407.
——, and ——, 1987: Two dimensional model studies of equatorial dynamics and tracer distributions. Quart. J. Roy. Meteor. Soc.,113, 635–651.
——, and T. J. Dunkerton, 1990: The role of the seasonal cycle in the quasi-biennial oscillation of ozone. J. Atmos. Sci.,47, 2429–2451.
Hamilton, K., and J. D. Mahlman, 1988: General circulation model simulation of the semiannual oscillation of the tropical middle atmosphere. J. Atmos. Sci.,45, 3212–3235.
Harries, J. E., S. Ruth, and J. M. Russell III, 1996a: On the distribution of mesospheric molecular hydrogen inferred from HALOE measurements of H2O and CH4. Geophys. Res. Lett.,23, 297–300.
——, and Coauthors, 1996b: Validation of measurements of water vapor from the Halogen Occultation Experiment. J. Geophys. Res.,101, 10205–10216.
Hitchman, M. H., M. McKay, and C. R. Trepte, 1994: A climatology of stratospheric aerosol. J. Geophys. Res.,99, 20689–20700.
Holton, J. R., and W.-K. Choi, 1988: Transport circulation deduced from SAMS trace species data. J. Atmos. Sci.,45, 1929–1939.
Jones, R. L., and J. A. Pyle, 1984: Observations of CH4 and N2O by the NIMBUS 7 SAMS: A comparison with in-situ data and two- dimensional numerical model calculations. J. Geophys. Res.,89, 5263–5279.
——, ——, J. E. Harries, A. M. Zavody, J. M. Russell III, and J. C. Gille, 1986: The water vapor budget of the stratosphere studied using LIMS and SAMS satellite data. Quart. J. Roy. Meteor. Soc.,112, 1127–1143.
Kelly, K. K., A. F. Tuck, L. E. Heidt, M. Loewenstein, J. F. Podolske, and S. E. Strahan, 1990: A comparison of ER-2 measurements of stratospheric water vapor between the 1987 Antarctic and 1989 Arctic airborn missions. Geophys. Res. Lett.,17, 465–468.
Kennaugh, R., S. Ruth, and L. J. Gray, 1997: Modeling quasi-biennial variability in the semi-annual double peak. J. Geophys. Res.,102, 16 169–16 188.
Kumer, J. B., J. L. Mergenthaler, and A. E. Roche, 1993: CLAES CH4, N2O and CCl2F2 (F12) global data. Geophys. Res. Lett.,20, 1239–1242.
Lahoz, W. A., and Coauthors, 1996a: Vortex dynamics and the evolution of water vapor in the stratosphere of the Southern Hemisphere. Quart. J. Roy. Meteor. Soc.,122, 423–450.
——, and Coauthors, 1996b: Validation of UARS microwave limb sounder 183 GHz H2O measurements. J. Geophys. Res.,101, 10129–10149.
Leovy, C. B., C.-R. Sun, M. H. Hitchman, E. E. Remsberg, J. M. Russell III, L. L. Gordley, J. C. Gille, and L. V. Lyjak, 1985: Transport of ozone in the middle stratosphere: Evidence for planetary wave breaking. J. Atmos. Sci.,42, 230–244.
LeTexier, H., S. Solomon, and R. R. Garcia, 1988: The role of molecular hydrogen and methane oxidation in the water vapor budget of the stratosphere. Quart. J. Roy. Meteor. Soc.,114, 281–295.
Luo, M., R. J. Cicerone, and J. M. Russell III, 1995: Analysis of Halogen Occultation Experiment HF versus CH4 correlation plots: Chemistry and transport implications. J. Geophys. Res.,100, 13927–13437.
——, J. M. Russell III, and T. Y. W. Huang, 1997: HALOE observations of the quasi-biennial oscillation and the effects of Pinatubo aerosols in the tropical stratosphere. J. Geophys. Res.,102, 19 187–19 198.
Manney, G. L., R. W. Zurek, A.O’Neill, and R. Swinbank, 1994: On the motion of air through the stratospheric polar vortex. J. Atmos. Sci.,51, 2973–2994.
——, L. Froidevaux, J. W. Waters, and R. W. Zurek, 1995: Evolution of microwave limb sounder ozone and the polar vortex during winter. J. Geophys. Res.,100, 2953–2972.
McCormick, M. P., E. W. Chiou, L. R. McMaster, W. P. Chu, J. C. Larsen, D. Rind, and S. Oltmans, 1993: Annual variations of water vapor in the stratosphere and upper troposphere observed by the Stratospheric Aerosol and Gas Experiment II. J. Geophys. Res.,98, 4867–4874.
McIntyre, M. E., and T. N. Palmer, 1983: Breaking planetary waves in the stratosphere. Nature,305, 593–600.
——, and ——, 1984: The “surf zone” in the stratosphere. J. Atmos. Terr. Phys.,46, 825–849.
Mote, P. W., K. H. Rosenlof, J. R. Holton, R. S. Harwood, and J. W. Waters, 1995: Seasonal variations of water vapor in the tropical lower stratosphere. Geophys. Res. Lett.,22, 1093–1096.
——, and Coauthors, 1996: An atmospheric tape recorder: The imprint of tropical tropopause temperatures on stratospheric water vapor. J. Geophys. Res.,101, 3989–4006.
Newell, R. E., and S. Gould-Stewart, 1981: A stratospheric fountain? J. Atmos. Sci.,38, 2789–2796.
Norton, W. A., 1994: Breaking Rossby waves in a model stratosphere diagnosed by a vortex following coordinate system and a technique for advecting material contours. J. Atmos. Sci.,51, 654–673.
Olaguer, E. P., H. Yang, and K. K. Tung, 1992: A re-examination of the radiative balance of the stratosphere. J. Atmos. Sci.,49, 1242–1263.
O’Sullivan, D., 1997: Cross equatorially radiating stratospheric Rossby waves. Geophys. Res. Lett.,24, 1483–1486.
——, and P. Chen, 1996: Modeling the quasi-biennial oscillation’s influence on isentropic tracer transport in the subtropics. J. Geophys. Res.,101, 6811–6821.
——, and T. J. Dunkerton, 1997: The influence of the quasi-biennial oscillation on global constituent distributions. J. Geophys. Res., in press.
Park, J. H., and Coauthors, 1996: Validation of HALOE CH4 measurements from UARS. J. Geophys. Res.,101, 10183–10204.
Pierce, R. B., W. Grose, J. R. Russell III, A. F. Tuck, R. Swinbank, and A. O’Neill, 1994: Spring dehydration in the Antarctic stratosphere vortex observed by HALOE. J. Atmos. Sci.,51, 2931–2941.
Plumb, R. A., and R. C. Bell, 1982: A model of the quasi-biennial oscillation on an equatorial beta-plane. Quart. J. Roy. Meteor. Soc.,108, 335–352.
Polvani, L. M., D. W. Waugh, and R. A. Plumb, 1995: On the subtropical edge of the stratospheric surf zone. J. Atmos. Sci.,52, 1288–1309.
Prather, M., and E. Remsberg, Eds., 1992: The atmospheric effects of stratospheric aircraft. NASA Ref. Publ. 1292, 268 pp.
Randel, W. J., and F. Wu, 1996: Isolation of the ozone QBO in SAGE II data by singular value decomposition. J. Atmos. Sci.,53, 2546–2559.
——, B. A. Boville, J. C. Gille, P. L. Bailey, S. T. Massie, J. B. Kumer, J. L. Mergenthaler, and A. E. Roche, 1994: Simulation of stratospheric N2O in the NCAR CCM2: Comparison with CLAES data and global budget analysis. J. Atmos. Sci.,51, 2834–2845.
Remsberg, E. E., J. M. Russell III, L. L. Gordley, J. C. Gille, and P. L. Bailey, 1984: Implications of stratospheric water vapor distributions as determined from the Nimbus 7 LIMS Experiment. J. Atmos. Sci.,41, 2934–2945.
——, P. P. Bhatt, and J. M. Russell III, 1996: Estimates of the water vapor budget of the stratosphere from UARS HALOE data. J. Geophys. Res.,101, 6749–6766.
Roche, A. E., and Coauthors, 1996: Validation of CH4 and N2O measurements by the cryogenic limb array etalon spectrometer instrumention on the Upper Atmosphere Research Satellite. J. Geophys. Res.,101, 9679–9710.
Rosenfield, J., P. A. Newman, and M. R. Schoeberl, 1994: Computations of diabatic descent in the stratospheric polar vortex. J. Geophys. Res.,99, 16677–16689.
Rosenlof, K. H., 1995: Seasonal cycle of the residual mean meridional circulation in the stratosphere. J. Geophys. Res.,100, 5173–5191.
——, A. F. Tuck, K. K. Kelly, J. M. Russell III, and M. P. McCormick, 1997: Hemispheric asymmetries in water vapor and inferences about transport in the lower stratosphere. J. Geophys. Res.,102, 13 213–13 234.
Russell J. M., III, A. F. Tuck, L. L. Gordley, J. H. Park, S. R. Drayson, J. E. Harries, R. J. Cicerone, and P. J. Crutzen, 1993a: HALOE Antarctic observations in the spring of 1991. Geophys. Res. Lett.,20, 719–722.
——, and Coauthors, 1993b: The Halogen Occultation Experiment. J. Geophys. Res.,98, 10777–10797.
Ruth, S., R. Kennaugh, and L. J. Gray, 1997: Seasonal, semiannual and interannual variability seen in measurements of methane made by the UARS Halogen Occultation Experiment. J. Geophys. Res.,102, 16 189–16 200.
Sassi, F., R. R. Garcia, and B. A. Boville, 1993: The stratopause semiannual oscillation in the NCAR community climate model. J. Atmos. Sci.,50, 3608–3624.
Schoeberl, M. R., and Coauthors, 1989: Reconstruction of the constituent distribution and trends in the Antarctic polar vortex from ER-2 flight observations. J. Geophys. Res.,94, 16815–16846.
——, L. R. Lait, P. A. Newman, and J. E. Rosenfield, 1992: The structure of the polar vortex. J. Geophys. Res.,97, 7859–7882.
——, M. Luo, and J. E. Rosenfield, 1995: An analysis of the Antarctic Halogen Occultation Experiment trace gas observations. J. Geophys. Res.,100, 5159–5172.
——, A. E. Roche, J. M. Russell III, D. Ortland, P. B. Hays, and J. W. Waters, 1997: An estimation of the dynamical isolation of the tropical lower stratosphere using UARS wind and trace gas observations of the quasi-biennial oscillation. Geophys. Res. Lett.,24, 53–56.
Solomon, S., J. T. Kiehl, R. R. Garcia, and W. Grose, 1986: Tracer transport by the diabatic circulation deduced from satellite observations. J. Atmos. Sci.,43, 1603–1617.
Stanford, J. L., J. R. Ziemke, and S. Y. Gao, 1993: Stratospheric circulation features deduced from SAMS constituent data. J. Atmos. Sci.,50, 226–246.
Strahan, S. E., J. E. Nielsen, and M. C. Cerniglia, 1996: Long-lived tracer transport in the Antarctic stratosphere. J. Geophys. Res.,101, 26615–26629.
Sutton, R., 1994: Lagrangian flow in the middle atmosphere. Quart. J. Roy. Meteor. Soc.,120, 1299–1321.
Swinbank, R., and A. O’Neill, 1994: A stratosphere–troposphere data assimilation system. Mon. Wea. Rev.,122, 686–702.
Trepte, C. R., and M. H. Hitchman, 1992: Tropical stratospheric circulation deduced from satellite aerosol data. Nature,355, 626–628.
——, R. E. Veiga, and M. P. McCormick, 1993: The poleward dispersal of the Mt. Pinatubo volcanic aerosol. J. Geophys. Res.,98, 18562–18573.
Tuck, A. F., J. M. Russell III, and J. E. Harries, 1993: Stratospheric dryness: Antiphased dessication over Micronesia and Antarctica. Geophys. Res. Lett.,20, 1227–1230.
Volk, C. M., and Coauthors, 1996: Quantifying transport between the tropical and midlatitude lower stratosphere. Science,272, 1763–1768.
Wallace, J. M., R. L. Panetta, and J. Estberg, 1993: Representation of the equatorial quasi-biennial oscillation in EOF phase space. J. Atmos. Sci.,50, 1751–1762.
Waugh, D. W., 1996: Seasonal variation of isentropic transport out of the tropical stratosphere. J. Geophys. Res.,101, 4007–4023.
Zawodny, J. M., and M. P. McCormick, 1991: Stratospheric aerosol and gas experiment II measurements of the quasi-biennial oscillation in ozone and nitrogen dioxide. J. Geophys. Res.,96, 9371–9377.
Scatter diagrams of HALOE CH4 observations at 10 mb during September–October 1993 plotted as a function of latitude (top) and potential vorticity (PV) (bottom). Potential vorticity is expressed in terms of equivalent latitude.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Comparison of HALOE CH4 observations during January 1994 calculated using zonal averages (top) and averages along PV contours (equivalent latitude mapping—bottom).
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Time series of HALOE H2O and CH4 data at 10 mb, 28°S. Crosses denote monthly mean data, and thin lines denote (repeating) seasonal cycles fit by harmonic regression.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Time series of CH4 data at 10 mb, 72°S. Crosses and solid lines denote HALOE observations and fit seasonal cycle; note data are only available during September–February for each year. Circles and dashed lines during 1992–93 show adjusted CLAES data and seasonal cycle fit.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Components of temporal variance in HALOE CH4 data over 1991–97. Shown are total variance (left), seasonal cycle variance (middle), and residual interannual variance (right). Contours in all panels are 0.002 ppmv2.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Seasonal cycle estimates of CH4 in January, April, July, and October. Contour interval is 0.1 ppmv.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Seasonal cycle estimates of H2O in January, April, July, and October. Contour interval is 0.2 ppmv. Values below 3.6 ppmv are hatched, and values above 6.0 ppmv are stippled.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Meridional cross section of the quantity H2O + 2CH4 in October, derived from the seasonal cycle H2O and CH4 data (Figs. 6 and 7). Contour interval is 0.2 ppmv, and values less than 6.6 ppmv are shaded.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time sections of seasonal cycle variations in CH4 at 2.2 mb (top), 10 mb (middle), and 68 mb (bottom). Contour interval is 0.01 ppmv. The dashed lines in the lower panel are the
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time sections of residual mean vertical velocity (
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time sections of seasonal cycle variations in CH4 at 76°N (top) and 76°S (bottom). Contour interval is 0.1 ppmv. Note the respective time axes are shifted by 6 months between the two panels.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Solid lines show seasonal cycles of CH4 at 76°S, 10 mb (top), and 46 mb (bottom). Dashed lines show CH4 estimated from
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Observed CH4 variations at 76°N and 10 mb (solid line) and that calculated from
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time sections of seasonal cycle variations in CH4 (top) and H2O (bottom) over the equator. Left panels show full fields, while right panels show seasonal anomalies (calculated by subtracting the time means at each pressure level). Values below 3.6 ppmv in the lower-left panel are shaded.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Time series of H2O + 2CH4 over 4°N–4°S during 1991–97, for pressure levels 100 mb to 10 mb. Crosses show monthly means and thin lines show (repeating) seasonal cycle fits at each level.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time plots of seasonality in H2O (left) and H2O + 2CH4 (right) at 100 mb (bottom), 68 mb (middle), and 46 mb (top). Contour interval is 0.2 ppmv in each panel.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Meridional cross sections of individual terms in the seasonal CH4 continuity equation [Eq. (1)], for averages during January–February. Contour interval in all panels is 0.05 ppmv month−1, with zero contours omitted. Vectors in the upper right-hand panel denote components of the residual mean circulation (
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time sections of terms in the seasonal CH4 continuity equation [Eq. (1)] over the equator. Shown are −
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time section of interannual anomalies in CH4 over the equator. Contour interval is 0.05 ppmv, with zero contours omitted. The vertical lines denote January for each year.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time sections of interannual anomalies in (a) zonal-mean zonal wind and (b) QBO-filtered vertical velocity
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Altitude–time sections of interannual anomalies in H2O over the equator. Contour interval is 0.1 ppmv, and zero contours are omitted.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Heavy solid and dashed lines with shading show the spatial structure of QBO SVD1 (left) and SVD2 (right) for CH4. Contours show local mixing ratio variations with arbitrary units. Light lines denote background annual mean CH4 structure.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Phase–space diagram of the projection of the monthly global CH4 anomalies onto spatial structures SVD1 and SVD2 (shown in Fig. 22). Time progression corresponds to counterclockwise orbit transits, with locations during several months noted.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Meridional cross sections of CH4 during April 1993, 1994, 1995, and 1996 (time periods denoted in Fig. 23). Note the double- peaked upper stratosphere patterns in 1993 and 1995.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time sections of CH4 at 3 mb during 1991–97. Shown are the full field (top) and the detrended interannual anomalies (bottom). Dashed lines and arrows denote periods of double-peaked (d) latitudinal structure.
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time sections of CH4 at 10 mb over 1991–97, for the full field (top—contour interval 0.1 ppmv) and interannual anomalies (bottom—contour interval of 0.05 ppmv, with zero contours omitted).
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time sections of H2O at 10 mb over 1991–97, for the full field (top—contour interval 0.2 ppmv) and interannual anomalies (bottom—contour interval of 0.1 ppmv, with zero contours omitted).
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
Latitude–time section of interannual anomalies in TEM vertical velocity (
Citation: Journal of the Atmospheric Sciences 55, 2; 10.1175/1520-0469(1998)055<0163:SCAQVI>2.0.CO;2
These seasonal cycle datasets (and similar ones for other HALOE constituents) are available to potential users via anonymous FTP from NCAR (contact the author at randel@ucar.edu).