1 Intraseasonal Variations of the Tropical Total Ozone and 2 Their Connection to the Madden-Julian Oscillation 3 4 Baijun Tian 5 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 6 Yuk L. Yung 7 Div of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA. 8 Duane E. Waliser 9 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 10 Tomasz Tyranowski 11 Jagiellonian University, Krakow, Poland 12 Le Kuai 13 Div of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA. 14 Fredrick W. Irion 15 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 16 17 J. Geophys. Res. 18 Drafted – 10/08/2006 19 20 ------------------------------------------- 21 Corresponding author address: Dr. Baijun Tian, Jet Propulsion Laboratory, California Institute of 22 Technology, M/S 183-501, 4800 Oak Grove Dr., Pasadena CA 91109. Email: Baijun.Tian@jpl.nasa.gov. 23 1 24 Abstract 25 The total ozone from the Total Ozone Mapping Spectrometer (TOMS) instrument on the 26 Nimbus 7 spacecraft and the Atmospheric Infrared Sounder (AIRS) instrument on the Aqua 27 spacecraft, the rainfall from the Global Precipitation Climatology Project (GPCP), and the 28 dynamic fields (geopotential height and stream function) from the NCEP/NCAR reanalysis are 29 employed to study the intraseasonal (30–90 day) variations of the tropical total ozone and their 30 connection to the Madden-Julian Oscillation (MJO). We found that the intraseasonal variations 31 of the tropical total ozone are large (around 5 DU) in both AIRS and TOMS and comparable to 32 those associated with annual cycle, quasi-biennial oscillation, El Nino and Southern Oscillation, 33 and solar cycle. In particular, the intraseasonal total ozone anomalies are mainly evident over the 34 subtropics, with a systematic relationship with the equatorial MJO convection anomaly 35 throughout the MJO cycle. Positive subtropical total ozone anomalies are typically collocated 36 with or lie to the west of the equatorial suppressed MJO convection and negative subtropical 37 total ozone anomalies are typically collocated with or lie to the west of the equatorial enhanced 38 MJO convection. The subtropical total ozone anomalies propagate eastward at a phase speed of 39 ~5 m s-1 similar to the equatorial MJO convection anomaly in the Eastern Hemisphere and they 40 propagate much faster (~15 m s-1) in the Western Hemisphere when the equatorial MJO 41 convection anomaly disappears. Comparison of the horizontal maps of the TOMS total ozone 42 anomalies and the NCEP 200-hPa geopotential height and stream function anomalies show that 43 negative (positive) subtropical total ozone anomalies are collocated with the subtropical upper- 44 tropospheric anticyclones (cyclones) and positive (negative) geopotential height anomalies with 45 a high negative correlation (~–0.75) between the total ozone and geopotential height anomalies. 46 This indicates that the subtropical total ozone anomalies are mainly dynamically driven as a 2 47 result of the subtropical upper-tropospheric cyclones or anticyclones generated by the equatorial 48 MJO convection. AIRS also indicates that negative (positive) equatorial total ozone anomalies 49 are collocated with the enhanced (suppressed) equatorial MJO convection, while TOMS does 50 not. 51 3 52 1. Introduction 53 The total column abundance of the ozone in the tropical atmosphere represents an 54 intricate interaction between chemical and dynamical processes [Brasseur et al., 1999]. Thus, the 55 tropical total ozone and its variability have been extensively studied especially after the launch of 56 the Total Ozone Mapping Spectrometer (TOMS) instrument on the Nimbus 7 spacecraft 57 [Krueger 1989]. Previous studies have investigated the trend [Stolarski, 1986; 1992; Herman et 58 al., 1991], annual cycle [Shiotani, 1992], quasi-biennial oscillation (QBO) [Oltmans and London, 59 1982; Bowman, 1989; Zawodny and McCormick, 1991; Shiotani, 1992; Tung and Yang, 1994a; 60 1994b], El Niño-Southern Oscillation (ENSO) or interannual variation [Shiotani, 1992; Randel 61 and Cobb, 1994; Kayano, 1997; Camp et al., 2003], and solar cycle [Hood, 1997] of the tropical 62 total ozone. It was found that the tropical total ozone varies on the order of ±10 Dobson Unit 63 (DU) (~3% of the mean) for annual cycle and ENSO, about ±15 DU (~5% of the mean) for 64 QBO, and about ±5 DU (~2% of the mean) for solar cycle. However, very few studies have 65 investigated the intraseasonal (30–90 day) variations of the tropical total ozone. The study by 66 Gao and Stanford [1990] indicated that there is a possible intraseasonal variability in the tropical 67 total ozone. However, the spatial and temporal patterns of the intraseasonal variations of the 68 tropical total ozone have not been well documented. In particular, the physical and dynamical 69 processes that are responsible for the intraseasonal variations of the tropical total ozone have 70 never been investigated. 71 The Madden-Julian Oscillation (MJO; aka Intraseasonal Oscillation) [Madden and Julian, 72 1972] is the dominant component of the intraseasonal variability in the tropical atmosphere. 73 Since its discovery, the MJO has continued to be a topic of significant interest due to its 74 extensive interactions with other components of the climate system and the fact that it represents 4 75 a connection between the better understood weather and seasonal-to-interannual climate 76 variations. The MJO is characterized by eastward-propagating, large-scale baroclinic oscillations 77 in the tropical wind field [Madden and Julian, 1972; Madden and Julian, 1994]. Over the 78 warmest tropical waters, in the equatorial Indian and western Pacific Oceans, the disturbance 79 manifests itself in the form of large-scale convection anomalies which interact with the upper- 80 level atmospheric flow, surface winds and heat fluxes [Knutson and Weickmann, 1987; Rui and 81 Wang, 1990; Hendon and Salby, 1994; Matthews et al., 2004]. In these regions, where the 82 convective activity is strong, the oscillation propagates slowly (~5 m s-1) and the interaction with 83 the extra-tropics is greatest. Once the disturbance reaches the date line, and thus cooler equatorial 84 waters, convection subsides and the remaining dynamical disturbance behaves much like a 85 damped Kelvin wave with a faster propagation speed [~15 m s-1, Hendon and Salby, 1994]. The 86 above characteristics tend to be most strongly exhibited during the boreal winter and spring 87 (November-April) when the Indo-Pacific warm pool is centered near the equator. During the 88 boreal summer and fall (May-October), the change in the large-scale circulation associated with 89 the Asian summer monsoon results in the largest-scale aspects of the disturbances propagating 90 more northeastward, from the equatorial Indian Ocean into Southeast Asia [e.g., Wang and Rui, 91 1990; Waliser, 2006]. For more comprehensive reviews of the MJO and related issues, the reader 92 is referred to Madden and Julian [1994], Lau and Waliser [2005], and Zhang [2005]. From the 93 discussion above, we can see that the large-scale MJO convection and circulation anomalies have 94 been well documented and understood; however, their connection to the intraseasonal variations 95 of the tropical total ozone has never been investigated. 96 The purpose of this study is to investigate the intraseasonal variations of the tropical total 97 ozone and its possible connection to the large-scale MJO convection and circulation anomalies. 5 98 We are attempting to address the following three questions. 1) Are there any intraseasonal 99 variations of the tropical total ozone? 2) If yes, how large are these variations and what the 100 spatial and temporal patterns of these variations are? 3) What physical or dynamical processes 101 are responsible for these variations? Is there any connection between the large-scale MJO 102 convection and circulation anomalies and the intraseasonal variations of the total ozone? 103 The outline for the rest of this paper is as follows: The data and methodology are 104 introduced in section 2. Section 3 presents the main results of this study and its interpretation. 105 Section 5 provides a summary of the major findings of this study. 106 107 2. Data and Methodology 108 For this study, we have used two data sets for the tropical total ozone: The Atmospheric 109 Infrared Sounder (AIRS) on the Aqua satellite and the TOMS on the Nimbus 7 satellite. The 110 AIRS, together with Advanced Microwave Sounding Unit, form the integrated hyperspectral 111 infrared and microwave atmospheric sounding system which has been operational since 112 September 2002 [Chahine et al., 2006]. The primary AIRS products are twice daily global fields 113 of atmospheric temperature and humidity profiles, ozone profile, surface skin temperature, and 114 cloud related parameters [Susskind et al., 2006]. In this study, we use AIRS Level 3 Version 4 115 total ozone product on a 1º × 1º latitude-longitude grid from September 2002 to July 2006. The 116 AIRS ozone retrieval is a research product that is currently undergoing validation. Gettelman et 117 al. [2004] have compared AIRS ozone retrieval with research aircraft measurements in the upper 118 troposphere; their limited comparisons suggest that the AIRS ozone has a 30% positive bias in 119 the upper troposphere but can reasonably track variability over a range of mixing ratios from 120 approximately 50–500 ppbv. 6 121 The Nimbus 7 TOMS provided daily global coverage of the Earth’s total ozone from 24 122 October 1978 to 6 May 1993 by measuring backscattered ultraviolet (UV) sunlight [Krueger 123 1989; McPeters et al., 1996]. The TOMS ozone data have been well validated and extensively 124 used for climate studies (see introduction). The TOMS will also provide more confidence on 125 MJO results because its longer record (13 years) allows more samples of MJO events than AIRS 126 (4 years). The TOMS data are on a 3º × 5º grid averaged from their original 1º × 1.25º grid. Only 127 data from 1 January 1980 to 31 December 1992 were used in this study. 128 To identify MJO events, the global pentad rainfall data from the Global Precipitation 129 Climatology Project (GPCP, Xie et al. 2003) covering from 1 January 1979 to June 30 2006 130 (should update to June 2006) and on a 2.5º by 2.5º grid were used. In the following discussion, 131 rainfall is used as a proxy for MJO convection. To identify the MJO dynamical fields, daily 132 geopotential height and stream function from the National Center for Environmental 133 Prediction/National Center for Atmospheric Research reanalysis [NCEP, Kalnay et al. 1996] 134 covering from 1979 to 2002 (up to July 2006??) were used. The spatial resolution of the NCEP 135 reanalysis is 2.5° x 2.5°. 136 The MJO analysis and composite procedure of Tian et al. [2006] were followed for this 137 study. All the data were first binned into 5-day average (i.e. pentad) values. Intraseasonal 138 anomalies were obtained by removing the annual cycle and then band-pass filtering (30–90 day) 139 the data. To isolate the dominant structure of the MJO, an extended empirical orthogonal 140 function (EEOF) was applied using time lags of ±5 pentads on boreal winter rainfall for the 141 region 30ºS-30ºN and 30ºE-150ºW. Next, MJO events were chosen based on the amplitude 142 pentad time series of the first EEOF mode of rainfall anomaly. Figure 1 shows the dates of the 143 selected MJO events for TOMS (a) and AIRS (b) data based on the rainfall anomaly. 24 MJO 7 144 events were selected for TOMS and 8 MJO events for AIRS. For each selected MJO event, the 145 corresponding 11-pentad rainfall, total ozone, geopotential height, and stream function anomalies 146 were extracted for each data set (GPCP, AIRS, TOMS, and NCEP). A composite MJO cycle (11 147 pentads) of the anomalies was then obtained by averaging the selected MJO events. 148 149 3. Results and Discussions 150 3.1 Intraseasonal variations of the tropical total ozone 151 Figure 2 shows the horizontal maps of the TOMS total ozone MJO anomalies (DU, color 152 shading) for the composite MJO cycle. For simplicity, only lags ±4, ±2, and 0 pentads of the 153 MJO cycle are shown. In addition, the contour plots overlaid on the color shadings are the 154 corresponding MJO-related GPCP rainfall anomalies (mm day-1). Note that the magnitude of the 155 composite ozone anomalies ranges up to about ±2.5 DU. However, inspection of the individual 156 events shows that they range up to about ±5 DU (~2% of the mean) but the compositing 157 procedure reduces the signal amplitude. This indicates that the intraseasonal variations of the 158 tropical total ozone are significant and comparable to those associated with other times scales, 159 such as the annual cycle, QBO, ENSO, and solar cycle. (Confidence limit) 160 Over the equatorial regions (10ºS–10ºN), the total ozone variations are rather small (<1 161 DU). On the other hand, significant total ozone intraseasonal variations (~2.5 DU) are mainly 162 found over the subtropical regions, with maxima near 25º latitudes, with great symmetry between 163 the Southern and Northern Hemispheres. In particular, systematic relationship is persistent 164 throughout the MJO cycle between the subtropical total ozone anomalies and the equatorial MJO 165 convection anomalies in the Eastern Hemisphere, where MJO convection is active. The negative 166 subtropical total ozone anomalies are typically collocated with or lie to the west of the enhanced 8 167 equatorial MJO convection and the positive subtropical total ozone anomalies are generally 168 collocated with or lie to the west of the suppressed equatorial MJO convection. The subtropical 169 total ozone anomalies propagate eastward at a phase speed of ~5 m s-1 similar to the equatorial 170 MJO convection anomaly. In the Western Hemisphere where the equatorial MJO convection 171 anomaly subsides, the remaining subtropical total ozone anomalies are still large and propagate 172 eastward at a much faster phase speed ~15 m s-1. 173 Diagram similar to Figure 2 but for the AIRS total ozone anomalies is presented in Figure 174 3. The total ozone anomalies in AIRS are much larger and noisier than TOMS. This is due 175 probably to the fact that much fewer MJO events considered in AIRS than TOMS. Comparison 176 of Figures 2 and 3 indicates both striking similarity and significant difference exist in the total 177 ozone anomalies between TOMS and AIRS. The similarity is mainly over the subtropics, while 178 the difference is at the equatorial region. Over the subtropics, TOMS and AIRS are consistent 179 with each other because both TOMS and AIRS show large total ozone anomalies (~2 DU) over 180 the subtropics although the total ozone anomalies are much noisier in AIRS than TOMS. Similar 181 to TOMS, AIRS also indicates similar systematic relationship between the subtropical total 182 ozone anomalies and the equatorial MJO convection anomalies comparing to TOMS: negative 183 (positive) subtropical total ozone anomalies are collocated with or lie to the west of the enhanced 184 (suppressed) equatorial MJO convection. The subtropical total ozone anomalies in AIRS also 185 propagate eastward at a phase speed of ~5 m s-1 in the Western Hemisphere and much faster (~15 186 m s-1) in the Western Hemisphere. Given the stark different techniques and time periods in 187 measuring the total ozone from TOMS (backscattered solar UV radiation, 1980-1992) and AIRS 188 (infrared, 2002-2006), the similarity in the subtropical total ozone anomalies and their 189 relationship to the equatorial MJO convection anomaly between TOMS and AIRS is striking. 9 190 This indicates that the intraseasonal variations of the subtropical total ozone and their 191 relationship with the equatorial MJO convection anomaly should be robust rather than an artifact 192 of the instruments. The physical or dynamical processes that are responsible for the intraseasonal 193 variations of the subtropical total ozone will be further discussed in subsection 3.2. 194 Over the equatorial regions, TOMS shows that the equatorial total ozone anomalies are 195 negligible (<1 DU); however, AIRS indicates the equatorial total ozone anomalies are significant 196 (~2 DU) especially in the regions where MJO convection is active, such as the Eastern 197 Hemisphere and South America. In AIRS, negative equatorial total ozone anomalies are typically 198 collocated with the enhanced equatorial MJO convection and positive equatorial total ozone 199 anomalies are generally collocated with the suppressed equatorial MJO convection. The 200 equatorial total ozone anomalies also propagate eastward at a phase speed of ~5 m s-1 similar to 201 the equatorial MJO convection anomaly. The relationship between the equatorial total ozone and 202 convection anomalies in the intraseasonal time scale is consistent with that found in the annual 203 cycle and ENSO time scales [e.g., Shiotani 1992; Hasebe 1993]. 204 3.2 Connection to the MJO dynamics 205 To understand the physical or dynamical processes that are responsible for the 206 intraseasonal total ozone anomalies discussed above, the horizontal maps of the TOMS total 207 ozone (DU, color shading) and the NCEP 200-hPa geopotential height (m, black contours) MJO 208 anomalies for the composite MJO cycle are presented in Figure 4. Clearly, negative (positive) 209 subtropical total ozone anomalies are typically collocated with positive (negative) subtropical 210 200-hPa geopotential height anomalies with a high negative correlation (~–0.75) between them. 211 (Confidence limit?) Similar analysis based on the geopotential height at 100-hPa also indicates a 212 strong negative correlation between the subtropical total ozone and 100-hPa geopotential height 10 213 anomalies. This indicates that the subtropical total ozone anomalies are mainly driven by the 214 vertical movement of the tropopause layer and its height. It is well known that concentrations of 215 ozone are much higher in the stratosphere and much lower in the troposphere with a sharp 216 boundary at tropopause [e.g., Fujiwara et al., 1998; Steinbrecht et al. 1998; Pan et al., 2004; more 217 refs?]. Thus, any process that depresses the height of the tropopause will tend to replace ozone- 218 poor tropospheric air by ozone-rich stratospheric air, and the total ozone will increase. On the 219 other hand, any process that lifts the tropopause height will tend to replace ozone-rich 220 stratospheric air by ozone-poor tropospheric air, and the total ozone will decrease. This 221 relationship has been noted at many time scales and by many researchers, for example, Dobson 222 et al. [1929; 1946], Reed [1950], Schubert and Munteanu [1988], Krueger [1989], Mote et al. 223 [1991], Salby and Callaghan [1993], and Steinbrecht et al. [1998]. Our results on the relationship 224 between the subtropical total ozone and the subtropical upper-tropospheric geopotential height 225 anomalies in the intraseasonal time scale are consistent with these previous studies. 226 To further elucidate the dynamical processes that cause the vertical movement of the 227 tropopause layer that are responsible for the intraseasonal total ozone anomalies, the horizontal 228 maps of the NCEP 200-hPa stream function (106 m2 s-1, color shading) and geopotential height 229 (m, black contours) MJO anomalies for the composite MJO cycle are shown in Figure 5. In 230 addition, the green contour plots overlaid on the color shadings and black contours are the 231 corresponding MJO-related GPCP rainfall anomalies (mm day-1). Please note negative (positive) 232 stream functions indicate a cyclone (an anticyclone) in the Northern Hemisphere, while negative 233 stream (positive) functions indicate an anticyclone (a cyclone) in the Southern Hemisphere. 234 Figure 5 shows that a subtropical upper-tropospheric anticyclonic couplet is collocated with or 235 lies to the west of the enhanced equatorial MJO convection and a subtropical upper-tropospheric 11 236 cyclonic couplet is collocated with or lies to the west of the reduced equatorial MJO convection. 237 To the east of the enhanced (reduced) equatorial MJO convection are equatorial upper- 238 tropospheric westerly (easterly) anomalies. These upper-tropospheric large-scale circulation 239 features associated with the equatorial MJO convection have been well documented [e.g., 240 Knutson and Weickmann, 1987; Rui and Wang, 1990; Hendon and Salby, 1994; Matthews et al., 241 2004] and can be interpreted as an equatorial Rossby-Kelvin wave response to the equatorial 242 MJO convection [Gill 1980]. 243 From Figure 5, we can clearly see that negative subtropical upper-tropospheric 244 geopotential height anomalies are typically collocated with the subtropical upper-tropospheric 245 cyclones and positive subtropical upper-tropospheric geopotential height anomalies are typically 246 collocated with the subtropical upper-tropospheric anticyclones, with a high positive correlation 247 (~0.92) between geopotential height and stream function anomalies. This indicates that the 248 vertical movement of the subtropical tropopause layer is mainly a dynamical result of the upper- 249 tropospheric subtropical cyclones or anticyclones. 250 Diagram similar to Figure 4 but for the TOMS total ozone and the NCEP 200-hPa stream 251 function (not shown) also clearly demonstrate that negative (positive) subtropical total ozone 252 anomalies are typically collocated with the subtropical upper-tropospheric anticyclones 253 (cyclones), with a high positive correlation (~0.75) between the total ozone and stream function 254 anomalies. Combining Figures 2, 4, and 5 together, it is evident that the subtropical total ozone 255 anomalies are mainly dynamically driven as a result of the subtropical upper-tropospheric 256 cyclones or anticyclones generated by the equatorial MJO convection. 257 Over the equatorial regions, the TOMS total ozone and the NCEP 200-hPa geopotential 258 height MJO anomalies are both small and their correlation is rather low (~0.15). However, AIRS 12 259 indicates significant equatorial total ozone anomalies are collocated with the equatorial MJO 260 convection. Show correlation between the equatorial total ozone anomalies in AIRS and the 261 NCEP 200-hPa geopotential height anomalies???? (expect good correlation or not???). This may 262 be due to the possible biases in the TOMS total ozone retrieval because TOMS cannot measure 263 the tropospheric ozone below optically thick clouds in the tropical deep convective regions 264 (Ref???). This indicates that the vertical movement of the tropopause will play a lesser in the 265 equatorial total ozone anomalies than in subtropics. There may be a contribution from the 266 coupled chemistry and dynamics of ozone in the troposphere and need further investigation. 267 268 4. Conclusions 269 This study is the first attempt to document the spatial and temporal patterns of the 270 intraseasonal variations of the tropical total ozone and their dynamical connection to the large- 271 scale MJO convection and circulation anomalies. This is motivated by the fact that the total 272 ozone is an important tracer for atmospheric motions [Dobson et al. 1929; Krueger 1989]. Based 273 on the total ozone data from TOMS and AIRS, we found that the intraseasonal variations of the 274 tropical total ozone are large (around 5 DU) in both TOMS and AIRS and comparable to those 275 associated with annual cycle, QBO, ENSO, and solar cycle. The intraseasonal total ozone 276 anomalies are mainly evident over the subtropics. In particular, there is a systematic relationship 277 between the subtropical total ozone anomalies and the equatorial MJO convection anomaly, that 278 is, positive anomalies are typically collocated with or lie to the west of the equatorial suppressed 279 MJO convection and negative anomalies are typically collocated with or lie to the west of the 280 equatorial enhanced MJO convection. The subtropical total ozone anomalies propagate eastward 281 at a phase speed of ~5 m s-1 similar to the equatorial MJO convection anomaly in the Eastern 13 282 Hemisphere and they propagate much faster (~15 m s-1) in the Western Hemisphere when the 283 equatorial MJO convection anomaly disappears. Comparison of the horizontal maps of the 284 TOMS total ozone anomalies and the NCEP 200-hPa geopotential height and stream function 285 anomalies show that negative (positive) subtropical total ozone anomalies are collocated with the 286 subtropical upper-tropospheric anticyclones (cyclones) and positive (negative) geopotential 287 height anomalies with a high negative correlation (~–0.75) between the total ozone and 288 geopotential height anomalies. This indicates that the subtropical total ozone anomalies are 289 mainly dynamically driven as a result of the subtropical upper-tropospheric cyclones or 290 anticyclones generated by the equatorial MJO convection. AIRS also indicates that negative 291 (positive) equatorial total ozone anomalies are collocated with the enhanced (suppressed) 292 equatorial MJO convection, while TOMS does not. Thus, there may be a contribution to the 293 equatorial total ozone anomalies from the coupled chemistry and dynamics of ozone in the 294 troposphere. 295 The strong connection between the total ozone variation and the large-scale MJO 296 convection and circulation anomalies has important implications for the future MJO studies. The 297 total ozone, unlike winds and geopotential height, has been routinely monitored on a global basis 298 by satellite since the launch of TOMS in 1978. Thus, the abundance of the satellite total ozone 299 data will provide an excellent tool to study the large-scale MJO convection and circulation 300 anomalies and their impact on the global weather and climate. 301 302 Acknowledgements 303 The research described in this paper was carried out at the Jet Propulsion Laboratory 304 (JPL), California Institute of Technology, under a contract with NASA. It was jointly supported 14 305 by the Research and Technology Development program, Human Resources Development fund, 306 and the Atmospheric Infrared Sounder project at JPL. The AIRS and TOMS ozone data and 307 GPCP rainfall data were downloaded from GSFC DAAC, respectively. 308 15 309 References 310 Bowman, K. P., 1989. Global patterns of the quasi-biennial oscillation in total ozone. J. Atmos. 311 312 313 Sci., 46, 3328-3343. Brasseur, G. P., J. J. Orlando, and G. S. Tindall (Eds.), 1999: Atmospheric Chemistry and Global Change, 654 pp., Oxford Univ. Press, New York. 314 Camp, C. D., M. S. Roulston, and Y. L. Yung, 2003: Temporal and spatial patterns of the 315 interannual variability of total ozone in the tropics. J. Geophys. Res., 108, 4643, 316 doi:10.1029/2001JD001504. 317 318 Chahine, M. T., and Coauthors. 2006: AIRS: Improving weather forecasting and providing new data on greenhouse gases. Bull. Amer. Meteor. Soc., 87, 911-926. 319 Dobson, G. M. B., D. N. Harrison, and J. Lawrence, 1929: Measurements of the amount of ozone 320 in the Earth's atmosphere and its relation to other geophysical conditions-Part III. Proc. 321 R. Soc. A122, 456-486. 322 323 Dobson, G. M. B., A. W. Brewer, and B. M. Cwilong, 1946: Meteorology of the lower stratosphere. Proc R. Soc London, 185, 144-175. 324 Fujiwara, M., K. Kita, and T. Ogawa, 1998: Stratosphere-troposphere exchange of ozone 325 associated with the equatorial Kelvin wave as observed with ozonesondes and 326 rawinsondes. J. Geophys. Res., 103, 19173-19182. 327 328 Gao, X. H., and J. L. Stanford, 1990: Low-frequency oscillations in total ozone measurements. J. Geophys. Res., 95, 13797-13806, 10.1029/90JD00899. 329 Gettelman, A., and Coauthors, 2004: Validation of satellite data in the upper troposphere and 330 lower stratosphere with in-situ aircraft instruments, Geophys. Res. Lett., 31, L22107, 331 doi:10.1029/2004GL020730. 16 332 333 334 335 336 337 Gill, A. E., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106, 447-462. Hasebe, F., 1993: Dynamical response of the tropical total ozone to sea surface temperature changes. . J. Atmos. Sci., 50, 345–356. Hendon, H. H., and M. L. Salby, 1994: The life cycle of the Madden-Julian oscillation. J. Atmos. Sci., 51, 2225-2237. 338 Herman, J. R., R. McPeters, R. Stolarski, D. Larko, and R. Hudson, 1991: Global average ozone 339 change from November 1978 to May 1990. J. Geophys. Res., 96, 17,297–17,305. 340 Hood, L. L., 1997: The solar cycle variation of total ozone: Dynamical forcing in the lower 341 342 343 344 345 stratosphere. J. Geophys. Res., 102, 1355–1370. Kalnay, E., et al., 1996: The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteor. Soc., 77, 437–471. Kayano, M. T., 1997: Principal modes of the total ozone on the Southern Oscillation timescale and related temperature variations. J. Geophys. Res., 102, 25,797–25,806. 346 Knutson, T. R. and K. N. Weickmann, 1987: 30-60 day atmospheric oscillations: Composite life 347 cycles of convection and circulation anomalies. Mon. Wea. Rev., 115, 1407-1436. 348 Krueger, A. J., 1989: The global distribution of total ozone: TOMS satellite measurements. 349 350 351 352 353 354 Planet. Space Sci., 37, 1555-1565. Lau, K.-M., and D. E. Waliser, 2005: Intraseasonal Variability in the Atmosphere-Ocean Climate System, Springer Praxis Books, 474 pp. Madden, R. A., and P. R. Julian, 1972: Description of global scale circulation cells in the tropics with a 40-50 day period. J. Atmos. Sci., 29, 1109-1123. Madden, R. A., and P. R. Julian, 1994: Observations of the 40-50-day tropical oscillation - a 17 355 review. Mon. Wea. Rev., 122, 814-837. 356 Matthews, A. J., B. J. Hoskins, and M. Masutani, 2004: The global response to tropical heating 357 in the Madden-Julian Oscillation during northern winter. Quart. J. Roy. Meteor. Soc., 358 130, 1991-2011. 359 360 361 362 363 364 McPeters, R., et al., 1996: Nimbus-7 Total Ozone Mapping Spectrometer (TOMS) data products user's guide, NASA Tech. Rep. Mote, P. W., J. R. Holton, and J. M. Wallace, 1991: Variability in total ozone associated with baroclinic waves. J. Atmos. Sci., 48, 1900-1903. Oltmans, S. J., and J. London, 1982: The quasi-biennial oscillation in atmospheric ozone, J. Geophys. Res., 87, 8981–8989. 365 Pan, L. L., W. J. Randel, B. L. Gary, M. J. Mahoney, and E. J. Hintsa, 2004: Definitions and 366 sharpness of the extratropical tropopause: A trace gas perspective, J. Geophys. Res., 109, 367 D23103, doi:10.1029/2004JD004982. 368 369 370 371 372 373 374 375 376 377 Randel, W. J., and J. B. Cobb, 1994: Coherent variations of monthly mean total ozone and lower stratospheric temperature. J. Geophys. Res., 99, 5433-5447. Reed, R. J., 1950: The role of vertical motions in ozone-weather relationships. J. Meteor., 7, 263267. Rui, H., and B. Wang, 1990: Development characteristics and dynamical structure of tropical intraseasonal convection anomalies. J. Atmos. Sci., 47, 357-379. Salby, M. L., and P. F. Callaghan, 1993: Fluctuations of total ozone and their relationship to stratospheric air motions. J. Geophys. Res., 98, 2715–2727. Schubert, S. D., and M.-J. Munteanu, 1988: An analysis of tropopause pressure and total ozone correlations. Mon. Wea. Rev., 116, 569-582. 18 378 379 Shiotani, M., 1992: Annual, quasi-biennial, and El Niño-Southern Oscillation (ENSO) time-scale variations in equatorial total ozone. J. Geophys. Res., 97, 7625–7633. 380 Steinbrecht, W., H. Claude, U. Köhler, K. P. Hoinka, 1998: Correlations between tropopause 381 height and total ozone: Implications for long-term changes. J. Geophys. Res., 103, 19183- 382 19192, 10.1029/98JD01929. 383 Stolarski, R. S., A. J. Krueger, M. R. Schoeberl, R. D. McPeters, P. A. Newman, and J. C. 384 Alpert, 1986: Nimbus 7 SBUV/TOMS measurements of the springtime Antarctic ozone 385 hole. Nature, 322, 808-811. 386 387 Stolarski, R. S., R. Bojkov, L. Bishop, C. Zerefos, J. Staehelin, and J. Zawodny, 1992: Measured trends in stratospheric ozone. Science, 256, 342 - 349. 388 Susskind, J., and Coauthors, 2006: Accuracy of geophysical parameters derived from 389 Atmospheric Infrared Sounder/Advanced Microwave Sounding Unit as a function of 390 fractional cloud cover, J. Geophys. Res., 111, D09S17, doi:10.1029/2005JD006272. 391 Tian, B. J., D. E. Waliser, E. J. Fetzer, B. H. Lambrigtsen, Y. Yung, and B. Wang, 2006: Vertical 392 moist thermodynamic structure and spatial-temporal evolution of the MJO in AIRS 393 observations. J. Atmos. Sci., 63, 2462-2485. 394 395 396 397 398 399 400 Tung, K. K., and H. Yang, 1994a: Global QBO in circulation and ozone: Part I. Reexamination of observational evidence. J. Atmos. Sci., 51, 2699–2707. Tung, K. K., and H. Yang, 1994b: Global QBO in circulation and ozone. Part II: A simple mechanistic model. J. Atmos. Sc., 51, 2708–2721. Waliser, D. E., 2006: Intraseasonal Variability, in The Asian Monsoon, edited by B. Wang, pp. 844 Springer, Heidelberg, Germany. Wang, B., and H. Rui, 1990: Synoptic climatology of transient tropical intraseasonal convection 19 401 anomalies. Meteor. Atmos. Phys., 44, 43-61. 402 Xie, P., J. E. Janowiak, P.A. Arkin, R. F. Adler, A. Gruber, R. R. Ferraro, G. J. Huffman, S. 403 Curtis, 2003: GPCP pentad precipitation analyses: An experimental dataset based on 404 gauge observations and satellite estimates. J. Climate, 16, 2197-2214. 405 Zawodny, J. M., and M. P. McCormick, 1991: Stratospheric Aerosol and Gas Experiment II 406 measurements of the quasi-biennial oscillations in ozone and nitrogen dioxide. J. 407 Geophys. Res., 96, 9371–9377. 408 409 Zhang, C., 2005: Madden-Julian Oscillation. Rev. Geophys., 43, RG2003, doi:10.1029/2004RG 000158. 410 411 20 Figure Captions 412 413 414 Figure 1: The dates of the selected MJO events for TOMS (a) and AIRS periods based on the 415 amplitude pentad time series for the first EEOF mode of GPCP (a) and TRMM (b) 416 rainfall anomaly from NH wintertime (November–April) and the region 30°N–30°S 417 and 30°E–150°W. 418 Figure 2: Composite TOMS total ozone MJO anomalies (DU, color shading). The superimposed 419 solid (dashed) black contours denote the GPCP rainfall MJO anomalies (mm day-1). 420 For simplicity, only lags -4, -2, 0, +2, and +4 pentads of the MJO cycle are shown. 421 Figure 3: As Figure 2 but for the AIRS total ozone and TRMM rainfall MJO anomalies. 422 Figure 4: Composite MJO cycles of the TOMS total ozone (DU, color shading) and the NCEP 423 200-hPa geopotential height (m, black contours) MJO anomalies. 424 Figure 5: Composite MJO cycles of the NCEP 200-hPa stream function (106 m2 s-1, color 425 shading) and the NCEP 200-hPa geopotential height (m, black contours) MJO 426 anomalies. 427 428 429 21 430 Figure 1: The dates of the selected MJO events for TOMS (a) and AIRS periods based on the amplitude pentad time series for the first EEOF mode of GPCP (a) and TRMM (b) rainfall anomaly from NH wintertime (November–April) and the region 30°N–30°S and 30°E–150°W. 431 432 22 433 Figure 2: Composite TOMS total ozone MJO anomalies (DU, color shading). The superimposed solid (dashed) black contours denote the GPCP rainfall MJO anomalies (mm day-1). For simplicity, only lags -4, -2, 0, +2, and +4 pentads of the MJO cycle are shown. 434 435 23 436 Figure 3: As Figure 2 but for the AIRS total ozone and TRMM rainfall MJO anomalies. 437 438 24 439 Figure 4: Composite MJO cycles of the TOMS total ozone (DU, color shading) and the NCEP 200-hPa geopotential height (m, black contours) MJO anomalies. 440 441 25 442 Figure 5: Composite MJO cycles of the NCEP 200-hPa stream function (106 m2 s-1, color shading) and the NCEP 200-hPa geopotential height (m, black contours) MJO anomalies. 443 26