Northward propagation of the subseasonal variability over the eastern Pacific warm pool

Similar documents
Two dominant subseasonal variability modes of the eastern Pacific ITCZ

Investigation of Common Mode of Variability in Boreal Summer Intraseasonal Oscillation and Tropospheric Biennial Oscillation

Impacts of intraseasonal oscillation on the onset and interannual variation of the Indian summer monsoon

3. Climatic Variability. El Niño and the Southern Oscillation Madden-Julian Oscillation Equatorial waves

Causes of the Intraseasonal SST Variability in the Tropical Indian Ocean

Biennial Oscillation of Tropical Ocean-Atmosphere System Associated with Indian Summer Monsoon

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

Comparison of the Structure and Evolution of Intraseasonal Oscillations Before and After Onset of the Asian Summer Monsoon

Influence of enhanced convection over Southeast Asia on blocking ridge and associated surface high over Siberia in winter

Hui Wang, Mike Young, and Liming Zhou School of Earth and Atmospheric Sciences Georgia Institute of Technology Atlanta, Georgia

Effect of late 1970 s Climate Shift on Interannual Variability of Indian Summer Monsoon Associated with TBO

An ITCZ-like convergence zone over the Indian Ocean in boreal late autumn


Goal: Develop quantitative understanding of ENSO genesis, evolution, and impacts

Intra-seasonal variations in the tropical atmospheric circulation. Climate Prediction Division Yayoi Harada

Satellite observations of intense intraseasonal cooling events in the tropical south Indian Ocean

Lecture 14. Heat lows and the TCZ

NOTES AND CORRESPONDENCE. On Wind, Convection, and SST Variations in the Northeastern Tropical Pacific Associated with the Madden Julian Oscillation*

Influence of El Nino Southern Oscillation and Indian Ocean Dipole in biennial oscillation of Indian summer monsoon

The MJO-Kelvin wave transition

Lecture 18: El Niño. Atmosphere, Ocean, Climate Dynamics EESS 146B/246B

ENSO Cycle: Recent Evolution, Current Status and Predictions. Update prepared by Climate Prediction Center / NCEP 8 March 2010

Joint Institute for Regional Earth System Science and Engineering, University of California, Los Angeles, Los Angeles, California DUANE E.

Lecture 24. El Nino Southern Oscillation (ENSO) Part 1

APPENDIX B NOAA DROUGHT ANALYSIS 29 OCTOBER 2007

Analysis of 2012 Indian Ocean Dipole Behavior

Satellite and Buoy Observations of Boreal Summer Intraseasonal Variability in the Tropical Northeast Pacific

ENSO Cycle: Recent Evolution, Current Status and Predictions. Update prepared by Climate Prediction Center / NCEP 4 September 2012

Lecture 33. Indian Ocean Dipole: part 2

Subsurface Ocean Indices for Central-Pacific and Eastern-Pacific Types of ENSO

Changes of The Hadley Circulation Since 1950

Local vs. Remote SST Forcing in Shaping the Asian-Australian Monsoon Variability

ATMS 310 Tropical Dynamics

Understanding El Nino-Monsoon teleconnections

Intra-seasonal Vagaries of the Indian Summer Monsoon Rainfall

The Amplitude-Duration Relation of Observed El Niño Events

Mesoscale air-sea interaction and feedback in the western Arabian Sea

The Role of the Wind-Evaporation-Sea Surface Temperature (WES) Feedback in Tropical Climate Variability

Subseasonal SST Variability in the Tropical Eastern North Pacific During Boreal Summer. Eric D. Maloney*, Dudley B. Chelton, and Steven K.

Haibo Hu Jie He Qigang Wu Yuan Zhang

Variability in the tropical oceans - Monitoring and prediction of El Niño and La Niña -

Increasing intensity of El Niño in the central equatorial Pacific

Traditional El Niño and El Niño Modoki Revisited: Is El Niño Modoki Linearly Independent of Traditional El Niño?

Propagation of planetary-scale zonal mean wind anomalies and polar oscillations

Intraseasonal Variability in Sea Level Height in the Bay of Bengal: Remote vs. local wind forcing & Comparison with the NE Pacific Warm Pool

Mechanistic links between the tropical Atlantic and the Indian monsoon in the absence of El Nino Southern Oscillation events

The Air-Sea Interaction. Masanori Konda Kyoto University

How fast will be the phase-transition of 15/16 El Nino?

SERIES ARTICLE The Indian Monsoon

Lecture 20. Active-weak spells and breaks in the monsoon: Part 1

Asymmetry in zonal phase propagation of ENSO sea surface temperature anomalies

Subseasonal SST Variability in the Tropical Eastern North Pacific during Boreal Summer

AILAN LIN TIM LI. IPRC, and Department of Meteorology, University of Hawaii at Manoa, Honolulu, Hawaii

The Effects of Gap Wind Induced Vorticity, the ITCZ, and Monsoon Trough on Tropical Cyclogenesis

Summer monsoon onset in the subtropical western North Pacific

The Influence of Amazon Rainfall on the Atlantic ITCZ through Convectively Coupled Kelvin Waves

Abrupt seasonal variation of the ITCZ and the Hadley circulation

Wintertime intraseasonal SST variability in the tropical South Indian Ocean and Role of Ocean Dynamics in the MJO Initiation

UNIFIED MECHANISM OF ENSO CONTROL ON INDIAN MONSOON RAINFALL SUNEET DWIVEDI

Atlantic warm pool, Caribbean low-level jet, and their potential impact on Atlantic hurricanes

Imprints of Coastal Mountains on Ocean Circulation and Variability

Goal: Describe the principal features and characteristics of monsoons

Indian Ocean dynamics and interannual variability associated with the tropospheric biennial oscillation (TBO)

Cyclogenesis. Text Definition and theory Necessary conditions How does a circulation spin up?

Lecture 13 El Niño/La Niña Ocean-Atmosphere Interaction. Idealized 3-Cell Model of Wind Patterns on a Rotating Earth. Previous Lecture!

Western Pacific Interannual Variability Associated with the El Nino Southern Oscillation

The South American monsoon system and the 1970s climate transition L. M. V. Carvalho 1, C. Jones 1, B. Liebmann 2, A. Silva 3, P. L.

Subsurface Ocean Temperature Indices for Central-Pacific and Eastern-Pacific Types of El Niño and La Niña Events

Use of Interactions between NAO and MJO for the Prediction of Dry and Wet Spell in Monsoon Season

Are Hurricanes Becoming More Furious Under Global Warming?

Physical mechanisms of the Australian summer monsoon: 2. Variability of strength and onset and termination times

The slab ocean El Niño

Trade winds How do they affect the tropical oceans? 10/9/13. Take away concepts and ideas. El Niño - Southern Oscillation (ENSO)

Supplementary Material for Satellite Measurements Reveal Persistent Small-Scale Features in Ocean Winds Fig. S1.

SUPPLEMENTARY INFORMATION

El Niño and La Niña sea surface temperature anomalies: Asymmetry characteristics associated with their wind stress anomalies

Influence of sea surface temperature on the intraseasonal variability of the South China Sea summer monsoon

A Proposed Mechanism for the Asymmetric Duration of El Niño and La Niña

Atmospheric Rossby Waves in Fall 2011: Analysis of Zonal Wind Speed and 500hPa Heights in the Northern and Southern Hemispheres

C.-P. Chang and Tim Li 1 Department of Meteorology, Naval Postgraduate School, Monterey, CA Abstract

The Influence of Indian Ocean Warming and Soil Moisture Change on the Asian Summer Monsoon

Three different downstream fates of the boreal-summer MJOs on their passages over the Maritime Continent

Interannual variation of northeast monsoon rainfall over southern peninsular India

Surface chlorophyll bloom in the Southeastern Tropical Indian Ocean during boreal summer-fall as reveal in the MODIS dataset

Global observations of stratospheric gravity. comparisons with an atmospheric general circulation model

Monsoon Systems Valsavarenche Val d Aosta, Italy

TROPICAL METEOROLOGY. Intertropical Convergence Zone. Introduction. Mean Structure

Effect of Orography on Land and Ocean Surface Temperature

Influence of mechanical mixing on a low summertime SST in the western North Pacific ITCZ region

INTRASEASONAL VARIABILITY IN THE FAR-EAST PACIFIC: INVESTIGATION OF THE ROLE OF AIR-SEA COUPLING IN A REGIONAL COUPLED MODEL

Data Analysis of the Seasonal Variation of the Java Upwelling System and Its Representation in CMIP5 Models

Climatology of the 10-m wind along the west coast of South American from 30 years of high-resolution reanalysis

Walker Circulation and El Niño / La Niña Sea Surface Temperature, Rainfall, and Zonal Wind

Roles of day oscillation in sustained rainstorms of October 2010 over Hainan, China

Mechanism of the Asymmetric Monsoon Transition as Simulated in an AGCM

The role of equatorial waves in the onset of the South China Sea summer monsoon and the demise of El Niño during 1998

Remote influence of Interdecadal Pacific Oscillation on the South Atlantic Meridional Overturning Circulation variability

Climate model errors over the South Indian Ocean thermocline dome and their. effect on the basin mode of interannual variability.

Lightning distribution with respect to the monsoon trough position during the Indian summer monsoon season

Monitoring and prediction of El Niño and La Niña

Transcription:

Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L09814, doi:10.1029/2008gl033723, 2008 Northward propagation of the subseasonal variability over the eastern Pacific warm pool Xianan Jiang 1 and Duane E. Waliser 1 Received 22 February 2008; accepted 9 April 2008; published 15 May 2008. [1] Previous studies suggest that easterly vertical shear of summer mean flow could be a key factor in the northward propagation of the subseasonal variability (SSV) over the Asian monsoon region (e.g., Jiang et al., 2004). Analysis indicates that this same mechanism may also be dictating propagation features over the eastern Pacific (EPAC). Consistent with the presence of local easterly vertical wind shear, northward propagation of the SSV is also evident over the EPAC with a phase speed of about 0.6 deg day 1. Moreover, similar to its counterpart in the Indian Ocean, the northward propagating SSV over this region is also characterized by positive anomalies of equivalent barotropic vorticity and lower-tropospheric specific humidity, and planetary boundary layer (PBL) convergence, to the north of the convection center. Thus, the present study provides another independent example of the essential role that easterly vertical wind shear may be playing in regulating the meridional migration of the SSV. Citation: Jiang, X., and D. E. Waliser (2008), Northward propagation of the subseasonal variability over the eastern Pacific warm pool, Geophys. Res. Lett., 35, L09814, doi:10.1029/2008gl033723. 1. Introduction [2] The significant role of the subseasonal variability (SSV) for tropical climate has been widely acknowledged [e.g., Lau and Waliser, 2005; Zhang, 2005]. In addition to an eastward propagating component (i.e., Madden-Julian Oscillation (MJO)) [Madden and Julian, 1994], which is strongest in boreal winter, SSV with a period of 30 50 days is characterized by prominent northward/north westward movement over the Indian/western Pacific Oceans during boreal summer [e.g., Yasunari, 1979; Hsu and Weng, 2001]. This meridional propagation of the SSV is found to be closely associated with active and break phases of Asian monsoon rainfall [e.g., Sikka and Gadgil, 1980; Lawrence and Webster, 2002], and thus has received significant attention over recent decades. However, current weather/ climate simulation and forecast models still struggle to properly represent this form of variability [e.g., Sperber et al., 2000; Waliser et al., 2003]. Having a better understanding of this northward propagating subseasonal mode would greatly benefit a wide-range of forecasting applications [Goswami, 2005]. [3] A number of theories have been advanced in interpreting this northward propagating SSV. Webster [1983] 1 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, California, USA. Copyright 2008 by the American Geophysical Union. 0094-8276/08/2008GL033723$05.00 emphasized the important role of land surface heat flux into the planetary boundary layer (PBL) to destabilize the atmosphere north of the convection. Wang and Xie [1997] and Lawrence and Webster [2002] suggested that the northward propagation of the SSV is due to Rossby wave emanation from the eastward propagating equatorial Kelvin- Rossby wave packet. Kemball-Cook and Wang [2001] and Fu et al. [2003], on the other hand, proposed that air-sea interaction could be another factor for this northward propagation mode. [4] Recently, based on analyses of both an atmospheric general circulation model simulation and observations, Jiang et al. [2004] identified marked meridional asymmetries in the vorticity and specific humidity fields associated with the northward propagating SSV over the Indian Ocean. A positive vorticity perturbation with an equivalent barotropic structure and maximum specific humidity are found to be located to the north of the convection center by a few degrees. By these findings, they proposed an easterly vertical wind shear mechanism in which the easterly vertical shear of the zonal mean flow over the Asia monsoon region (Figure 1a) could be fundamental for the northward propagation. This mechanism has also been affirmed by numerical simulations based on an intermediate model [Drbohlav and Wang, 2005, 2007]. [5] The key process associated with this mechanism is illustrated by the following equation of the barotropic component of vorticity (z + ), which is based on a simplified 2-D version of an intermediate model on a b-plane (the zonal variations of variables are neglected; for details please refer to Wang [2005]), @z þ @t @w ¼ bv þ U T @y where b is the meridional variation of Coriolis parameter; v + is the barotropic component of meridional wind, and w is the p-level vertical velocity. U T denotes the constant vertical shear of the basic zonal flow, which is defined as the difference in zonal mean wind between upper- and lowerlevels. This equation indicates that in the presence of vertical easterly shear (U T < 0), positive (negative) barotropic vorticity anomalies are generated to the north (south) of the convection center due to the northward decrease (increase) of upward vertical motion associated with the subseasonal convection. The induced positive barotropic vorticity in the free atmosphere further causes a moisture convergence in the PBL due to Ekman pumping, which creates conditions that favor northward movement of enhanced rainfall. Since the Coriolis effect is involved with the Ekman pumping process, this mechanism will only be valid away from the equator and ð1þ L09814 1of6

Figure 1. (a) Vertical shear of climatological summer mean (JJAS) zonal wind (200 hpa minus 850 hpa; unit: m s 1 ) based on reanalysis-2 data set over the 1982 2005 period. (b) Meridional profiles (average over 120W 100W) of the vertical shear of the summer mean zonal wind (black curves with shading donating the easterly vertical shear; lower axis) and summer mean (JJAS) SST based on NOAA OISST v2 (1981 2006; blue curve; upper axis). more effective with the increase of the Coriolis parameters [Jiang et al., 2004]. [6] It is intriguing to note that in Figure 1a, in addition to a vast area of the Asian monsoon region, the easterly vertical shear of the summer mean flow is also evident over a small region of the eastern Pacific (EPAC). Meanwhile, the SSVof convection and winds with a period of about 40 days over the EPAC warm pool during boreal summer has been widely documented [e.g., Knutson and Weickmann, 1987; Maloney and Hartmann, 2000; Maloney and Esbensen, 2007]. These observations raise the following questions, if the easterly vertical wind shear is indeed essential for the northward propagation of the SSV, is northward propagation of the SSV evident over the EPAC? Moreover, is there similarity in the underlying thermodynamic and dynamic structures of the northward propagating SSV between this region and its counterpart in the Asian monsoon sector? The analysis in this paper seeks to address these two questions with the intention of providing an independent test of the easterly vertical wind shear mechanism. 2. Data and Approach [7] A local index of the SSVover the EPAC is obtained by an extended empirical orthogonal function [EEOF] analysis of the TRMM (Tropical Rainfall Measuring Mission, version 3B42) rainfall over the EPAC warm pool (140 W 90 W; EQ-30 N). Before the EEOF analysis, 3-day mean rainfall data from January 1998 to December 2005 with horizontal resolution of 2 2 degrees are calculated based on the raw TRMM data set (0.5 0.5 degrees; daily), and are subject to band-pass time filtering to retain the periods of 10 90 days. The EEOF is conducted with 10 temporal lags of the 3-day mean data. In order to explore the dynamic and thermodynamic structures associated with the meridional propagating SSV over the EPAC, three-dimensional daily u, v, temperature, and relative humidity are employed based on the NCAR-DOE Reanalysis-2 [Kanamitsu et al., 2002] for the same period. Specific humidity is calculated based on the Clausius-Clapeyron equation by using temperature and relative humidity. To further substantiate the evidence obtained based on the reanalysis data set, shorter periods of QuikSCAT surface wind (2000 2005) and specific humidity by the Atmospheric Infrared Sounder (AIRS) [Fetzer et al., 2006; Tian et al., 2006] (2003 2005) are employed. All these variables are also subject to the same time filtering and 3-day mean bin averaging procedures as used for the TRMM rainfall. 3. Results [8] The spatial pattern of EEOF 1 based on the TRMM rainfall and its associated time series are displayed in Figure 2. The EEOF 2 mode is in quadrature with the EEOF 1, and exhibits the same propagation behavior as the EEOF 1 mode. This first pair of EEOFs explains 8.1% of the total anomalous variance of 3-day average data and is well separated from other EEOFs based on North s criterion [North et al., 1982] (figure not shown). Both of them display a dominant period of 40 days based on spectral analysis of the time series. This 40-day mode is more active during boreal summer than winter (Figure 2a), which could 2of6

Figure 2. (a) Time series of temporal coefficients for the EEOF 1 mode. (b) Lagged regression patterns of TRMM rainfall (shaded) and QuikSCAT surface wind (vectors) versus the standardized time series of temporal coefficients for the EEOF 1 mode during boreal summer (JJAS). The time interval between adjacent panels is 3 days. be ascribed to local cold SST in the winter due to the wind stress curl induced by strong gap winds over this region [Chelton et al., 2000; Xie et al., 2005]. The main features of this 40-day mode in rainfall and surface wind fields (Figure 2b) as derived by lag-regression against time series of the EOF 1 largely resemble those documented by Maloney and Esbensen [2007]. Note that a global MJO index has been employed in their study to obtain the SSV patterns in the EPAC, suggesting that the SSV over the EPAC could be a local expression of the MJO. [9] While the eastward movement associated with this SSV has been discussed and is also clearly evident by Figure 2b, the focus here is to check the meridional migration associated with this mode. Figure 3a illustrates a Hovmöller diagram of rainfall over the longitudes of 130 W 90 W. The northward propagation of convection is readily discerned and has a phase speed of about 0.6 latitude deg day 1, which is slower than that observed over the Indian Ocean (about 1 deg day 1 )[Jiang et al., 2004]. In addition, the northward movement is generally confined between 5 N and 25 N which is largely consistent with the region where easterly vertical shear of the mean wind is present (grey shading in Figure 1b), except for the equatorial region south of 5 N where convection is greatly suppressed due to cold SST (blue curve in Figure 1b). Note that the maximum easterly vertical wind shear is observed around 9 N, while faster northward propagation tends to occur between 10 15 N (Figure 3a). This could be explained by the stronger coupling between the mean vertical wind shear and perturbation vertical motion associated with the SSV over 10 15 N [Equation 1], given much stronger convection over these latitudes supported by local warmest SST (blue curve, Figure 1b). Moreover, as previously mentioned, more effective Ekman effect due to a larger Coriolis parameter over 10 15 N could also support stronger northward migration of the SSV than lower latitudes. In addition, the stronger easterly vertical wind shear over the Asian sector (Figure 1a) compared to the EPAC is generally consistent with the observed faster northward propagation speed over the former region. [10] In order to examine the comprehensive evolution features associated with this subseasonal mode, regression 3of6

Figure 3. (a) Time-latitude distribution of rainfall perturbation (units: mm day 1 ). Blue (red) dots show temporal points during negative (positive) rainfall phase which are employed for composite analysis to obtain meridional structures of various variables with respect to the convection center. Meridional-vertical structures of (b) vorticity (10 6 s 1 ) and (c) divergence (10 6 s 1 ) based on the reanalysis. (d) and (e) Black curves represent vorticity and divergence at the surface based on QuikSCAT winds (10 6 s 1 ; left ordinates), while green curves are rainfall based on TRMM observations (mm day 1 ; right ordinates). (f) and (g) Specific humidity structures based on the reanalysis and AIRS observations (g kg 1 ). X-axis in Figures 3b 3g is the meridional distance (deg) with respect to the convection center. The positive (negative) value means to the north (south). All variables are averaged over longitudes between 130 W 90 W. 4of6

patterns of various variables versus the time series of the EEOF 1 mode (Figure 2a) during boreal summer (JJAS) are calculated at various lags. Similarly to Jiang et al. [2004], meridional structures of these variables with respect to the convection center during the northward propagation can be obtained based on a composite analysis. The temporal points used for the composite are labeled by the dots at the bottom of Figure 3a, with blue dots for the negative rainfall phase and red ones for the positive phase. At each of these temporal points, the convection center is identified; then meridional structures of vorticity, divergence, and specific humidity with respect to the convection center are averaged between 130 W and 90 W. These structures are averaged over the negative and positive rainfall phases. They are further combined together (positive phase minus negative phase) since they largely mirror each other but with an opposite sign, and are illustrated in Figures 3b 3g. [11] It is noteworthy that the most salient features associated with the northward propagating 30 50 day SSV over the Indian Ocean, as illustrated by Jiang et al. [2004], are also identified with the SSV over the EPAC. In particular, a positive vorticity field with an equivalent barotropic structure is observed 2 3 degrees to the north of convection center (Figure 3b). Meanwhile, convergence in the PBL also tends to lead the convection center to the north (Figure 3c). The northward displacement of PBL convergence is largely consistent with positive equivalent barotropic vorticity perturbation, thus indicating the role of Ekman-pumping as suggested by Jiang et al. [2004]. The positive vorticity and PBL convergence to the north of the convection center, as identified in the reanalysis data set shown in Figures 3b and 3c, are further affirmed by the vorticity (Figure 3d) and divergence (Figure 3e) signals at the surface based on QuikSCAT observations. [12] Moreover, as with its counterpart over Asian monsoon sector, positive specific humidity anomalies associated with northward migrating SSV over the EPAC is also found to lead the convection center, especially in the lower troposphere, based on both the reanalysis (Figure 3f) and AIRS data sets (Figure 3g). Note that differences can be evident in the humidity structures depicted by these two data sets. The northward shift of the maximum specific humidity anomaly relative to the convection center is about three degrees by the reanalysis and eight degrees by the AIRS data. The amplitude of the moisture perturbation is slightly weaker in the AIRS observations than that in the reanalysis. Also, stronger vertical tilting of the moisture field is found in the AIRS observations, with clear PBL drying near the convection center, which is not evident in the reanalysis. This PBL drying near the convection center could be associated with mesoscale downdraft processes [e.g., Kiladis et al., 2005; Fu et al., 2006; Tian et al., 2006]. However, a period of eight years of the reanalysis data set has been employed when calculating the composite patterns, and only three years for the AIRS observations, and thus could partially be responsible for the difference in these two composite moisture fields. 4. Summary and Discussion [13] Previous investigation indicates that easterly vertical shear of the summer mean flow may exert direct influence on the northward propagation of the SSV over the Asian monsoon region through the easterly vertical wind shear mechanism. In addition to the Asian monsoon region, easterly vertical shear of the mean flow is also evident over the EPAC warm pool during boreal summer, although with relatively weaker amplitude and a smaller spatial extent. Thus, inspection of meridional migration of the SSV over the EPAC offers an excellent and independent opportunity to test this mechanism. [14] The analysis illustrates that a clear northward propagation of the SSV is indeed evident over the EPAC warm pool. The northward movement begins around the northern edge of the cold tongue of SST (5 N), and proceeds until 25 N with a phase speed around 0.6 deg day 1 (about 0.8 m s 1 ). The occurrence of the northward propagation over this meridional extent is largely consistent with the presence of easterly vertical wind shear over this region, supporting the important role of the easterly vertical wind shear in the northward propagation of the SSV as was proposed for its Asian Monsoon counterpart. [15] Moreover, it is demonstrated that similar meridional structures of a number of variables associated with the northward propagating SSV in the Indian Ocean are also evident over the EPAC. In particular, a positive vorticity with equivalent barotropic structure is found several degrees north of the convection center. Meanwhile, a northward shift of PBL convergence relative to the convection center is also evident and corresponds well to the positive vorticity, suggesting the influence of the free atmosphere on the PBL through Ekman-pumping. These important features in the perturbation vorticity and divergence fields, derived from the reanalysis, are further affirmed by independent evidence at the surface via Quikscat observations. Moreover, as with its counterpart over the Indian Ocean, the northward propagating SSV over the EPAC is also characterized by the northward shift of positive specific humidity to the convection center, based on both the reanalysis data set and recent AIRS observations, although differences in the moisture fields based on these two data sets are discernible. [16] Since characteristics of the northward movement of the SSV as well as its associated asymmetric structures in the dynamic and thermodynamic fields over the Indian/ Pacific sectors and EPAC exhibit great similarity, the result in this study indicates that internal atmospheric dynamics may play a key role in organizing the meridional migrating SSV. It is also noteworthy that, similar to its counterpart over the Asian sector, coherent SSV in SST is also observed in association with the northward propagating convection signal over the EPAC, with suppressed (enhanced) rainfall perturbation leading positive (negative) SST anomalies by 7 10 days [Maloney et al., 2008]. Further investigation is warranted to quantify the relative roles of internal dynamics versus air-sea interaction in contributing to the northward propagation of subseasonal convection over the EPAC. Since the SSV over the EPAC exerts significant influences on regional climate features, such as tropical cyclone genesis, easterly wave activity, as well as surge events over the Gulf of California associated with the North American monsoon variability [e.g., Lorenz and Hartmann, 2006], improved understanding of the SSV over this region would greatly benefit climate prediction for North and Central America. 5of6

[17] Acknowledgments. We are indebted to G. Kiladis, E. Maloney, X. Fu, and E. Fetzer for their valuable discussions. The TRMM 3B42 rainfall and the AIRS water vapor data sets were processed and provided by B. Tian. We also thank anonymous reviewers for their insightful comments, which lead to considerable improvements of this manuscript. This research was carried out at the Jet Propulsion Laboratory (JPL), Caltech, under a contract with NASA. X. Jiang and D. Waliser were jointly supported by the Research and Technology Development program and Human Resources Development Fund at JPL, as well as the NASA MAP and NEWS programs. References Chelton, D. B., M. H. Freilich, and S. N. Esbensen (2000), Satellite observations of the wind jets off the Pacific coast of Central America. Part I: Case studies and statistical characteristics, Mon. Weather Rev., 128, 1993 2018. Drbohlav, H.-K. L., and B. Wang (2005), Mechanism of the northwardpropagating intraseasonal oscillation: Insights from a zonally symmetric model, J. Clim., 18, 952 972. Drbohlav, H.-K. L., and B. Wang (2007), Horizontal and vertical structures of the northward-propagating intraseasonal oscillation in the South Asian Monsoon region simulated by an intermediate model, J. Clim., 20, 4278 4286. Fetzer, E. J., B. H. Lambrigtsen, A. Eldering, H. H. Aumann, and M. T. Chahine (2006), Biases in total precipitable water vapor climatologies from Atmospheric Infrared Sounder and Advanced Microwave Scanning Radiometer, J. Geophys. Res., 111, D09S16, doi:10.1029/2005jd006598. Fu, X., B. Wang, T. Li, and J. P. McCreary (2003), Coupling between northward-propagating, intraseasonal oscillations and sea surface temperature in the Indian Ocean, J. Atmos. Sci., 60, 1733 1753. Fu, X., B. Wang, and L. Tao (2006), Satellite data reveal the 3-D moisture structure of tropical intraseasonal oscillation and its coupling with underlying ocean, Geophys. Res. Lett., 33, L03705, doi:10.1029/ 2005GL025074. Goswami, B. N. (2005), South Asian Monsoon, in Intraseasonal Variability in the Atmosphere-Ocean System, edited by K. M. Lau and D. E. Waliser, pp. 19 61, Springer, Heidelberg, Germany. Hsu, H.-H., and C.-H. Weng (2001), Northwestward propagation of the intraseasonal oscillation in the western North Pacific during the boreal summer: Structure and mechanism, J. Clim., 14, 3834 3850. Jiang, X., T. Li, and B. Wang (2004), Structures and mechanisms of the northward propagating boreal summer intraseasonal oscillation, J. Clim., 17, 1022 1039. Kanamitsu, M., et al. (2002), NCEP-DOE AMIP-II Reanalysis (R2), Bull. Am. Meteorol. Soc., 83, 1631 1643. Kemball-Cook, S., and B. Wang (2001), Equatorial waves and air-sea interaction in the boreal summer intraseasonal oscillation, J. Clim., 14, 2923 2942. Kiladis, G. N., K. H. Straub, and P. T. Haertel (2005), Zonal and vertical structure of the Madden-Julian Oscillation, J. Atmos. Sci., 62, 2790 2809. Knutson, T. R., and K. M. Weickmann (1987), 30 60 day atmospheric circulations: Composite life cycles of convection and circulation anomalies, Mon. Weather Rev., 115, 1407 1436. Lau, W. K. M., and D. E. Waliser (Eds.) (2005), Intraseasonal Variability of the Atmosphere-Ocean Climate System, 474 pp., Springer, Heidelberg, Germany. Lawrence, D. M., and P. J. Webster (2002), The boreal summer intraseasonal oscillation: Relationship between northward and eastward movement of convection, J. Atmos. Sci., 59, 1593 1606. Lorenz, D. J., and D. L. Hartmann (2006), The effect of the MJO on the North American monsoon, J. Clim., 19, 333 343. Madden, R. A., and P. R. Julian (1994), Observations of the 40 50 day tropical oscillation A review, Mon. Weather Rev., 122, 814 837. Maloney, E. D., and S. K. Esbensen (2007), Satellite and buoy observations of boreal summer intraseasonal variability in the tropical northeast Pacific, Mon. Weather Rev., 135, 3 19. Maloney, E. D., and D. L. Hartmann (2000), Modulation of eastern North Pacific hurricanes by the Madden-Julian Oscillation, J. Clim., 13, 1451 1460. Maloney, E. D., D. B. Chelton, and S. K. Esbensen (2008), Subseasonal SST variability in the tropical eastern North Pacific during boreal summer, J. Clim., in press. North, G. R., T. L. Bell, R. F. Calahan, and F. J. Moeng (1982), Sampling errors in the estimation of empirical orthogonal functions, Mon. Weather Rev., 110, 699 706. Sikka, D. R., and S. Gadgil (1980), On the maximum cloud zone and the ITCZ over Indian longitudes during the southwest monsoon, Mon. Weather Rev., 108, 1840 1853. Sperber, K. R., J. M. Slingo, and H. Annamalai (2000), Predictability and the relationship between subseasonal and interannual variability during the Asian Summer Monsoon, Q. J. R. Meteorol. Soc., 126, 2545 2574. Tian, B., D. E. Waliser, E. Fetzer, B. Lambrigtsen, Y. Yung, and B. Wang (2006), Vertical moist thermodynamic structure and spatial-temporal evolution of the Madden-Julian Oscillation in Atmospheric Infrared Sounder observations, J. Atmos. Sci., 63, 2462 2485. Waliser, D. E., et al. (2003), AGCM simulations of intraseasonal variability associated with the Asian Summer Monsoon, Clim. Dyn., 21, 423 446. Wang, B. (2005), Theory, in Intraseasonal Variability in the Atmosphere- Ocean System, editedbyk.m.lauandd.e.waliser,pp.19 61, Springer, Heidelberg, Germany. Wang, B., and X. Xie (1997), A model for the boreal summer intraseasonal oscillation, J. Atmos. Sci., 54, 72 86. Webster, P. J. (1983), Mechanisms of low-frequency variability: Surface hydrological effects, J. Atmos. Sci., 40, 2110 2124. Xie, S.-P., H. Xu, W. S. Kessler, and M. Nonaka (2005), Air-sea interaction over the eastern Pacific warm pool: Gap winds, thermocline dome, and atmospheric convection, J. Clim., 19, 5 20. Yasunari, T. (1979), Cloudiness fluctuations associated with the Northern Hemisphere summer monsoon, J. Meteorol. Soc. Jpn., 57, 227 242. Zhang, C. (2005), The Madden-Julian Oscillation, Rev. Geophys., 43, RG2003, doi:10.1029/2004rg000158. X. Jiang and D. E. Waliser, Jet Propulsion Laboratory, California Institute of Technology, MS-183-501, Pasadena, CA 91109, USA. (xianan@caltech.edu) 6of6