Tropical–Extratropical Interaction Associated with the 10...

21
Journal of the Meteorological Society of Japan, Vol. 80, No. 2, pp. 311--331, 2002 311 Tropical–Extratropical Interaction Associated with the 10–25-day Oscillation over the Western Pacific during the Northern Summer Yoshiki FUKUTOMI 1 and Tetsuzo YASUNARI 2 Institute of Geoscience, University of Tsukuba, Tsukuba, Japan (Manuscript received August 2000, in revised form 11 January 2002) Abstract Ten years (1986–1995) of ECMWF analyses and NOAA outgoing longwave radiation (OLR) data were used to investigate intraseasonal tropical–extratropical interactions on timescales of 10–25 days in the western Pacific region during Northern Hemisphere (NH) summer. Composite analyses were performed separately for the early summer (June–July) and mid-summer (August). The analyses were based on the OLR index in the base region (the South China Sea), and help describe the seasonal characteristics of circulation associated with tropical convection. Localized Eliassen–Palm (E–P) fluxes were used to diagnose eddy–mean flow interactions. Composite circulation patterns associated with enhanced convection on a 10–25-day timescale for June–July were significantly different from those for August. June–July composites show a significant Rossby wave train in both the lower and upper troposphere, extending from the convective region into the North Pacific along the mean westerly flow. In the lower troposphere, subtropical circulation anoma- lies as part of this wavetrain propagate southwestward along the subtropical duct. Tropical–extratropical linkage is less pronounced during August. Circulation structures in August in both the lower and upper troposphere are different from those of June–July; Rossby waves do not emanate from the South China Sea and low-level subtropical circulation anomalies show no southwestward phase propagation. Energy conversion between eddy and seasonal mean flow for each season was computed to investigate the role of the time-mean flow in maintaining the 10–25-day transient eddies. In the lower troposphere, barotropic processes dominate the conversion from the time-mean flow to eddies in the exit region of tropical monsoon westerlies for both June–July and August. In the subtropics, baroclinic processes sus- tain the eddies. Moist baroclinic processes associated with eddy horizontal latent heat transport are im- portant in increasing the potential energy available to eddies, especially during June–July. In the upper troposphere, barotropic conversion of kinetic energy from eddies into the time-mean flow is most notable in the East Asian subtropics to the north of the South China Sea. This suggests that increased eddy activity due to tropical convection is an important source of energy for the mean flow. 1. Introduction The characteristics of higher-frequency in- traseasonal waves linked to tropical convection, and their associated tropical–extratropical in- teractions, have been the subject of consider- able interest in the past decade. Intraseasonal oscillations on so-called submonthly (about 6– 30-day) timescales (e.g., Meehl et al. 1996) are recognized as an important component of the subseasonal atmospheric variability as same as the Madden–Julian Oscillation (MJO: about 1 Present affiliation: Frontier Research System for Global Change, Institute for Global Change Re- search, JAMSTEC Yokohama Institute for Earth Sciences, Yokohama 236-0001, Japan. 2 Corresponding author: Tetsuzo Yasunari, Insti- tute of Geoscience, University of Tshukuba, Tsu- kuba, 305-8571, Japan. E-mail: [email protected] Additional affiliation: Frontier Research System for Global Change. ( 2002, Meteorological Society of Japan

Transcript of Tropical–Extratropical Interaction Associated with the 10...

Page 1: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

Journal of the Meteorological Society of Japan, Vol. 80, No. 2, pp. 311--331, 2002 311

Tropical–Extratropical Interaction Associated with the 10–25-day

Oscillation over the Western Pacific during the Northern Summer

Yoshiki FUKUTOMI1 and Tetsuzo YASUNARI2

Institute of Geoscience, University of Tsukuba, Tsukuba, Japan

(Manuscript received August 2000, in revised form 11 January 2002)

Abstract

Ten years (1986–1995) of ECMWF analyses and NOAA outgoing longwave radiation (OLR) data wereused to investigate intraseasonal tropical–extratropical interactions on timescales of 10–25 days in thewestern Pacific region during Northern Hemisphere (NH) summer. Composite analyses were performedseparately for the early summer (June–July) and mid-summer (August). The analyses were based on theOLR index in the base region (the South China Sea), and help describe the seasonal characteristics ofcirculation associated with tropical convection. Localized Eliassen–Palm (E–P) fluxes were used todiagnose eddy–mean flow interactions.

Composite circulation patterns associated with enhanced convection on a 10–25-day timescale forJune–July were significantly different from those for August. June–July composites show a significantRossby wave train in both the lower and upper troposphere, extending from the convective region intothe North Pacific along the mean westerly flow. In the lower troposphere, subtropical circulation anoma-lies as part of this wavetrain propagate southwestward along the subtropical duct. Tropical–extratropicallinkage is less pronounced during August. Circulation structures in August in both the lower and uppertroposphere are different from those of June–July; Rossby waves do not emanate from the South ChinaSea and low-level subtropical circulation anomalies show no southwestward phase propagation.

Energy conversion between eddy and seasonal mean flow for each season was computed to investigatethe role of the time-mean flow in maintaining the 10–25-day transient eddies. In the lower troposphere,barotropic processes dominate the conversion from the time-mean flow to eddies in the exit region oftropical monsoon westerlies for both June–July and August. In the subtropics, baroclinic processes sus-tain the eddies. Moist baroclinic processes associated with eddy horizontal latent heat transport are im-portant in increasing the potential energy available to eddies, especially during June–July. In the uppertroposphere, barotropic conversion of kinetic energy from eddies into the time-mean flow is most notablein the East Asian subtropics to the north of the South China Sea. This suggests that increased eddyactivity due to tropical convection is an important source of energy for the mean flow.

1. Introduction

The characteristics of higher-frequency in-traseasonal waves linked to tropical convection,and their associated tropical–extratropical in-teractions, have been the subject of consider-able interest in the past decade. Intraseasonaloscillations on so-called submonthly (about 6–30-day) timescales (e.g., Meehl et al. 1996) arerecognized as an important component of thesubseasonal atmospheric variability as sameas the Madden–Julian Oscillation (MJO: about

1 Present affiliation: Frontier Research System forGlobal Change, Institute for Global Change Re-search, JAMSTEC Yokohama Institute for EarthSciences, Yokohama 236-0001, Japan.

2 Corresponding author: Tetsuzo Yasunari, Insti-tute of Geoscience, University of Tshukuba, Tsu-kuba, 305-8571, Japan.E-mail: [email protected] affiliation: Frontier Research Systemfor Global Change.( 2002, Meteorological Society of Japan

Page 2: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

30–60 days) (e.g., Madden and Julian 1972,1994). Vincent et al. (1998) presented a com-prehensive review of observational work onsubmonthly scale variability. After Kiladisand Weickmann (1992a, b) identified significantlinkages between tropical deep convection andlarge-scale circulation, a series of studies clari-fied the nature of tropical convective distur-bances and the associated tropical–extratropicalinteractions on submonthly timescales. For ex-ample, Kiladis and Weickmann (1992b), Kiladis(1998), and Matthews and Kiladis (1999a) in-vestigated midlatitude Rossby-wave forcingof tropical convection over the eastern Pacific.Meehl et al. (1996) showed that this type ofRossby-wave propagation into the tropics couldexcite intertropical convergence zone (ITCZ)convection and generate equatorial Rossbywaves (e.g., Kiladis and Wheeler 1995; Numa-guti 1995; Wheeler and Kiladis 1999). Kiladiset al. (1994) and Meehl et al. (1996) showedthat these equatorial Rossby waves propagatewestward along the equator into the westernPacific, and contribute to tropical westerly windbursts accompanied by flare-ups of deep con-vection. Kiladis and his collaborators used theregression composite method to elucidate therole of submonthly timescale waves during NHwinter. In addition, Mo (1999, 2000) reportedthat midlatitude waves over the Pacific, whichare responsible for North American precipita-tion variability on a similar timescale, link towestward-propagating convection in the tropi-cal Pacific. In particular, Mo (2000) examinedthe teleconnection between the Pacific intra-seasonal mode and North American rainfallduring the NH summer. Relationships betweenextratropical circulations on submonthly time-scales and tropical and subtropical convectionin the South Pacific convergence zone (SPCZ)(e.g., Kiladis and Weickmann 1992b; Schrageand Vincent 1996; Vincent et al. 1997) and theSouth Atlantic convergence zone (SACZ) (e.g.,Kiladis and Weickmann 1992a, 1997; Lieb-mann et al. 1999) have also been studied.

The relationship between tropical deep con-vection and extratropical Rossby waves at sub-monthly timescales is two-way. Tropical con-vection forces downstream extratropical waves,and the propagation of waves from midlatitudestoward the tropics induces tropical or subtropi-cal convection. Tropical–extratropical interac-

tions on submonthly timescales and the rolesof the interactions in tropical and subtropicalvariability during the NH summer over thePacific basin have not been studied in depth. Asummertime extreme of convective activity inthis frequency range occurs over the westerntropical Pacific region including the SouthChina Sea (e.g., Chen and Chen 1995; Vincentet al. 1998; Kiladis 1998). Our study examineshow this convection is triggered and the im-pact that it has on large-scale circulation. Itseems reasonable to expect that a tropical–extratropical interaction similar to those men-tioned above plays an important role in tropicaland subtropical monsoon activities.

In an earlier study, Fukutomi and Yasunari(1999, hereafter FY99) suggested that a two-way tropical–extratropical interaction on a 10–25-day timescale exists over the western Pacificduring the early summer (June–July). They ex-amined the relationships between tropical con-vection over the South China Sea and large-scale circulation. They showed a lower- andupper-tropospheric wavetrain with zonal wave-number 5@6 scale arcing into North Pacificmidlatitudes and suggested that the wavetrainis amplified as a Rossby-wave response follow-ing enhanced (suppressed) convection over theSouth China Sea. A subtropical ridge (trough)that is part of this wavetrain modulates thesubtropical cloud activities over the Baiu fron-tal zone (BFZ). In the lower troposphere, asubtropical anticyclonic (cyclonic) circulationanomaly propagates southwestward into theSouth China Sea region where the anomaly cansuppress (enhance) subsequent convection.They proposed that this two-way interactionbetween tropics and extratropics is the domi-nant feature in the western Pacific during thisseason. Furthermore, seasonal (June–July)mean westerlies continuously extend from thetropical monsoon region to the North Pacific,acting as a duct that allows southwestwardphase propagation of lower tropospheric wavesfrom the subtropics to the tropics.

FY99, however, considered data only for fourearly summer seasons, from 1991 through 1994,and ignored the behavior of 10–25-day distur-bances during the mid-summer (August). Mostprevious studies of summertime intraseasonaldisturbances considered a standard season(JJA). As will be shown in later sections, how-

Journal of the Meteorological Society of Japan312 Vol. 80, No. 2

Page 3: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

ever, the background basic state on which theintraseasonal disturbances are superimposeddiffers considerably between early and mid-summer over the western Pacific. The changesin the mean background flow that occur as thesummer evolves may affect the development oftropical convective disturbance and their asso-ciated large-scale circulation over the westernPacific. Therefore, the structures of tropical 10–25-day disturbances and associated large-scalecirculation and their differences should be in-vestigated for each phase of the summer seasonto understand the character of subseasonal cir-culation systems in this region.

FY99 also speculated that dynamical inter-actions with the time-mean flow helped amplifyand maintain 10–25-day waves. Few studieshave assessed this interaction between sub-monthly scale eddies and the time-mean flowin the western Pacific during NH summer.Kawamura and Murakami (1995) examined theinteraction between MJO-timescale eddies andthe mean monsoon flow in the lower tropo-sphere. They suggested that eddies are main-tained mostly by barotropic instabilities in thetropics and by baroclinic instabilities in mid-latitudes. It is natural to question whether sim-ilar mechanisms exist to maintain 10–25-dayeddies.

This study extends FY99. Space-time struc-tures of convection and large-scale circulationanomalies at the 10–25-day time scale over thewestern Pacific are detailed to identify featuresin two seasons. Composite relationships be-tween tropical convection and large-scale circu-lations on the 10–25-day timescale during earlysummer (June–July) and mid-summer (August)are used to isolate the dominant circulationpatterns associated with tropical convection forearly and mid-summer. Composites are con-structed using data from 1986 through 1995.The structure and behavior of 10–25-day con-vective disturbances and associated waves inthe subtropical and extratropical latitudes canbe discerned. The target convection amplifiesover the South China Sea. We will identify theseasonal dependence of the large-scale circu-lations associated with tropical convection overthe South China Sea. Localized Eliassen–Palm(E–P) flux diagnostics (Trenberth 1986) andtheir application (Kawamura and Murakami1995) are then used to assess eddy–mean flow

interactions and to quantify the role of the meanflow in maintaining 10–25-day eddies. Eddyforcing of the mean flow is also investigated.

This paper is organized as follows. Section 2describes data sources and processing. In Sec-tion 3, characteristics of disturbances and of themean flow in the two seasons (June–July andAugust) are given. In Sections 4 and 5, compos-ite relationships between the 10–25-day con-vective events and large-scale circulation over aselected base region are presented. The struc-ture and characteristics of disturbances are alsooutlined in these sections. Section 6 includesdiagnostics of the eddy–mean flow interactionsand an evaluation of the dynamical interactionbetween the eddy and time mean flow. Thepaper concludes with a summary and discus-sion in Section 7.

2. Data sources and processing

This study uses the same datasets as FY99.Meteorological data (zonal and meridional wind,temperature, geopotential height, and relativehumidity) at ten standard pressure levels (1000,850, 700, 500, 400, 300, 250, 200, 150, 100 hPa)were extracted from the European Center forMedium-Range Weather Forecasts (ECMWF)/World Climate Research Programme (WCRP)/Tropical Ocean-Global Atmosphere (TOGA)BASIC LEVEL III operational analyses from1986 through 1995. We use uninitialized datadefined on a 2.5� lat-long grid. The original tem-poral resolution was twice daily (0000 and 1200UTC). Humidity data were only available up to300 hPa. Further details on the ECMWF dataare provided in Trenberth and Olson (1988).Specific humidity was computed from relativehumidity and temperature. A daily mean thateliminated the diurnal cycle was computedusing an average of the twice-daily data.Streamfunctions were computed from the hori-zontal winds using R25 truncation.

This study also uses outgoing longwave radi-ation (OLR) data as a proxy for convection. Theinterpolated OLR dataset for times overlappingthe ECMWF data was obtained from the Na-tional Oceanic and Atmospheric Administra-tion (NOAA)/Climate Diagnostics Center (CDC)and has global coverage on a 2.5� lat-long grid.

The first step in the data processing was toremove the seasonal trend from time series of allelements for the five-month period May to Sep-

Y. FUKUTOMI and T. YASUNARI 313April 2002

Page 4: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

tember. The seasonal trend was computed bythe least squares method as a quadratic curvefor the five months. A Butterworth band-passfilter (Kaylor 1977) with six poles was then ap-plied to the time series of each perturbationquantity to filter temporally into the 10–25-dayband. The 10–25-day band was chosen becauseFY99 and Fukutomi (2000) showed that thisfrequency was prominent. In addition, the 10-day low-pass cut off and 25-day high-pass cutoff eliminate synoptic- and MJO-timescale fluc-tuations, respectively. The 10–25-day timescaleis similar to those used by Kiladis (6–30-day)and by Vincent et al. (1998) (6–25-day).

3. Differences in transient disturbanceactivity between June–July andAugust

The 10–25-day disturbance activity is com-puted from the OLR variance and the per-turbation kinetic energy (PKE). The PKE isderived from band-pass filtered meteorologi-cal data and is expressed as in Matthews andKiladis (1999a), i.e.,

PKE ¼ 1

2ðu 02 þ v 02Þ; ð3:1Þ

where u and v are the zonal and meridionalwind components, ð�Þ denotes the time mean,and ð�Þ0 is the time deviation with 10–25-dayband-pass filtering.

PKE and OLR variance distributions andbackground mean fields are displayed in Fig-ures 1 and 2, respectively. From these figures,we can determine where transient disturbancesare most common in June–July and August inthe western Pacific. The PKE fields at 850 hPa(Fig. 1a) show a maximum oriented southwest–northeast from the South China Sea to theNorth Pacific during June–July. This maxi-mum PKE band parallels the mean low-levelwesterlies (as seen in Fig. 1c). In the tropics,the maximum PKE is over the northern SouthChina Sea. The OLR variance field for June–July (Fig. 1b) shows maximum convective activ-ity on the 10–25-day timescale over the SouthChina Sea and Bay of Bengal, in agreementwith FY99. Past studies have shown that thepeak OLR variance for the 10–25-day timescaleoccurs over these regions during the summer(JJA) (e.g., Kiladis and Weickmann 1997; Vin-cent et al. 1998). Here, mean fields for June–

July (Fig. 1c) are consistent with the maturephase of the south and southeast Asian mon-soons and the Baiu front, which is marked byquasi-stationary subtropical cloud activity inEast Asia. The monsoon westerlies and thetropical easterlies on average meet over thePhilippines, where the highest disturbanceactivity in the tropics is found.

Fig. 1. Distributions of mean fields inJune–July 1986–1995. (a) The 10–25-day perturbation kinetic energy (PKE)at 850 hPa. The contour interval is1 m2 s�2. Shading indicates PKE lessthan 3 m2 s�2. (b) Variance and per-centage variance of 10–25-day filteredOLR. The contour interval is 100 W2

m�4. (c) Mean OLR and 850-hPa windvectors and isotachs of zonal wind. Neg-ative contours are dashed. The regionsof OLR values less than 240 W m�2 areshaded.

Journal of the Meteorological Society of Japan314 Vol. 80, No. 2

Page 5: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

A striking feature during August is the east-ward extension of the PKE band from theSouth China Sea to the western Pacific in thetropics and subtropics (Fig. 2a). The tropicalportion of this PKE maximum exists in con-junction with a large OLR variance (Fig. 2b) at130�–150�E. Active regions of convection ex-pand eastward from the South China Sea to thetropical western Pacific as the mean tropicalmonsoon westerlies penetrate into the tropicalwestern Pacific near 150�E (Fig. 2c). This east-ward shift from June–July through August issimilar to the shift for MJO-scale eddies dem-onstrated by Kawamura and Murakami (1995).

In summary, enhanced disturbance activityin the lower troposphere on the 10–25-daytimescale is concentrated in the exit region ofmonsoon westerlies, i.e., where the zonal gra-dient of the mean westerlies is negative. The

mean fields (Fig. 2c) change dramatically inAugust, as noted by many previous studies(e.g., Matsumoto 1992; Murakami and Matsu-moto 1994). The distribution of the transienteddies also changes in association with thisstructural change in the mean flow. Distur-bance activity in the tropics is closely related tothe spatial distribution of the mean flow, asshown by Webster and Yang (1989).

4. Composite structures duringJune–July

Compositing was used to determine thespace–time structures of disturbances asso-ciated with tropical convection on 10–25-daytimescales. The composite procedure used wasnearly identical to that of FY99. The referenceOLR time series in the base region 10�–20�N,110�–120�E (the South China Sea) was adoptedas a convective index. As noted in Section 3,this is the region of highest convective activityat the 10–25-day timescale as shown in theOLR variance (Figs. 1 and 2). Composite mapsinclude eight states (Categories 1 through 8)that correspond to individual phases of the con-vection lifecycle over the base region. When thenegative (positive) extreme of the OLR indexaveraged over the base region exceeded �1.0(1.0) standard deviations, that extreme was as-signed to Category 3 (Category 7). This proce-dure holds in all composites in this analysis.The mean time span between two subsequentphases is approximately 1.5@2 days. Compositefields were prepared for both June–July andAugust to reveal the structures and charac-teristics of disturbances during these times.The composite patterns are significant at the95% confidence level as revealed by a standardt-test.

4.1 Horizontal structuresTwo-way interactions between tropics and

extratropics over the western Pacific in June–July were outlined in Section 1. The interac-tions are present in the space-time structuresof the 10–25-day wave disturbances and therelated tropical–extratropical interactions dur-ing June–July as discussed in FY99. Compo-sites in FY99, however, were based on a rela-tively small number of June–July cases during1991–94. In the present analysis, we producecomposite fields from a total of 32 cases over 10

Fig. 2. As in Fig. 1, but for August 1986–1995.

Y. FUKUTOMI and T. YASUNARI 315April 2002

Page 6: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

years (1986–1995). June–July composites arebased on the OLR index for the South ChinaSea because the OLR variance is largest in thisregion during June–July. The composites arecomparable to those for August, as discussed inSection 5.

This section focuses only on phases duringenhanced convection (Categories 2–4), as char-acteristics of the entire composite lifecycle (fromCategory 1 through 8) are described in detail byFY99. Figure 3 shows horizontal composites of

OLR, 850-hPa wind vector, and streamfunctionanomalies for Categories 2, 3, and 4, bracketingthe peak in convection over the base region.The relationship between the convection rep-resented by negative OLR anomalies in theSouth China Sea and the large-scale circula-tion anomalies is present, as in FY99. Follow-ing the convective peak (Category 3), a wave-train clearly emanates from the convectiveregion into the North Pacific. This wave struc-ture shows good agreement with FY99, but the

Fig. 3. The composite OLR, 850-hPa streamfunction and vector wind anomalies for Categories 2, 3, and4 for June–July based on the OLR index in region 10�–20�N, 110�–120�E. The contour interval is3.0 105 m2 s�1. OLR anomalies less than �4 W m�2 are shaded. Only locally statistically significantwind vectors are shown.

Journal of the Meteorological Society of Japan316 Vol. 80, No. 2

Page 7: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

anomalies described here are smoother than inFY99, reflecting the larger number of samplesused to construct the composites. The cyclonicor anticyclonic circulation anomalies that arepart of this wavetrain are amplified along thelow-level mean westerlies that extend from thetropics to midlatitudes. Patterns of circulationand OLR anomalies for Categories 6, 7, and 8(not shown) are reversed from Fig. 3. These twocategories represent the phases during whichconvection is suppressed over the South ChinaSea. During the composite cycle (Category 1through 8) provided by FY99, rotational windanomalies propagate southwestward from the

subtropics to the tropics, and precede the sub-sequent suppression or enhancement of convec-tion over the base region. These features arealso clear in the present composite life cycle.We made composites based on the region 10�–20�N, 130�–140�E (not shown); a downstreamwave structure similar to that in Fig. 3 resulted.

Figure 4 is similar to Fig. 3, but shows the200-hPa wind fields. The composite map showsa wavetrain developing away from the north ofthe convection center, crossing the midlatitudePacific. These composite wave anomalies re-semble those in FY99 and Mo (2000). The cross-equatorial outflow anomalies into the Southern

Fig. 4. As in Fig. 3, but for 200-hPa circulation fields. The contour interval is 5.0 105 m2 s�1.

Y. FUKUTOMI and T. YASUNARI 317April 2002

Page 8: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

Hemisphere and the southwesterly outflowanomalies into midlatitudes are strongest overthe convection in Categories 3 and 4. Theseoutflows are the divergent response to the con-vection. Fukutomi (2000) produced compositesof velocity potential and the divergent compo-nent of wind anomalies at 200 hPa. These com-posites show that the upper-level divergent out-flow is displaced from the convection center,which confirms the presence of an immediateresponse to the convectively produced diabaticheating over the South China Sea in the form ofan atmospheric circulation.

4.2 Vertical structuresWe examined the vertical structure of the

10–25-day disturbance using cross sections(category–pressure) of temperature, specific hu-

midity, geopotential height, and zonal windperturbations averaged along 10�–20�N, 110�–120�E, at pressure levels from 1000 to 100 hPain the vertical. The vertical section of geo-potential height (Fig. 5a) shows an equivalentbarotropic structure with negative anomaliesextending to about 300 hPa around Category 3.This suggests that a deep cyclonic vortex in thelow and middle troposphere dominates over thebase region. A sign change occurs between 300and 200 hPa, and positive anomalies (an anti-cyclone) are pronounced in the layers above250 hPa. A similar pattern of anomalies, butof opposite sign, occurs around Category 7. Thezonal wind anomalies (Fig. 5b) show deep west-erlies with a maximum at 700 hPa at a positioncorresponding to the southern part of the low-level monsoon trough in Fig. 3. These deep

Fig. 5. Category–pressure diagrams in the region 10�–20�N, 110�–120�E of the composite values ofthe 10–25-day filtered (a) geopotential height (m), (b) zonal wind (m s�1), (c) temperature (K), and(d) specific humidity (g kg�1). Dark (light) shading represents positive (negative) significant values.

Journal of the Meteorological Society of Japan318 Vol. 80, No. 2

Page 9: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

westerlies are similar to those accompanyingwintertime submonthly-scale waves coupledwith convection in the tropical central Pacific,as described by Kiladis et al. (1994) and Meehlet al. (1996). The upper-level easterlies thatstrengthen in the 200–100-hPa layer, cappingthe deep westerly flow, show the increase of thevertical easterly shear in the middle to uppertroposphere.

The structures of composite variables over thebase region have also been examined for Cate-gory 3 using longitude–pressure and latitude–pressure diagrams (Fukutomi 2000). The phasestructures of the anomalous geopotential heightand zonal winds in the zonal direction showlittle tilt in the vertical. Pronounced polewardphase tilts with height are observed in latitude–pressure diagrams in the 300–150-hPa layer.Phase nodes in the upper troposphere near thetropopause imply the predominance of firstbaroclinic modes associated with the strongconvective disturbance. Although Lau and Lau(1990) showed a northwestward phase tilt withheight over the western Pacific of the geo-potential height and vorticity anomalies asso-ciated with synoptic-scale disturbances withperiod of less than 10 days, our composites donot have such a structure. This is one of thedifferences between 10–25-day intraseasonaldisturbances and shorter-period disturbances.

Temperature anomalies (Fig. 5c) have a largemaximum between 500 and 200 hPa, above thelow-level cyclonic circulation associated withthe peak convection (Category 3) in Figure 3.This upper-level warming is probably caused bycondensation. Negative temperature anomaliesin the lower troposphere under the peak in theupper-layer warming may arise from adiabaticcooling associated with ascent. A deep layer ofmoist air embedded in the low-level cycloniccirculation during the former half cycle appearsin the specific humidity field (Fig. 5d). Thesemoist conditions are probably organized by cy-clonic and convergent inflow in the lower tro-posphere (see Fig. 3). An increase in moisturein the lowest (1000–850-hPa) layer slightly pre-cedes the peak in convection. Temperature andspecific humidity perturbations create upperwarm- (cool-) and lower moist- (dry-) core struc-tures within cyclonic (anticyclonic) regionsduring times of enhanced (suppressed) convec-tion.

5. Composite structure during August

Composites are prepared for OLR and windfields in August for comparison to those ofJune–July. August composites use the samebase region as June–July composites do. Thisregion occurs where the relative maxima ofOLR variance (Fig. 2a) and PKE (Fig. 2b) arelocated. Figure 6a shows horizontal compositesfor Category 3 OLR and winds at 850 hPausing the South China Sea base region. Tencases are averaged to produce this composite.The disturbance does have a structure similarto that of June–July. However, a wavetrainemanating from the convection downstream intomidlatitudes, as shown in Fig. 3, is not clearlyseen associated with this peak in convection.Furthermore, southwestward phase propaga-tion of circulation anomalies from the subtropicsinto the base region does not dominate fromCategory 1 through 8 (not shown). In the uppertroposphere (Fig. 6b), weak northerly outflowaway from the base region appears, but themidlatitude trough /ridge signals are weakerthan in June–July. Statistically significant cir-culations over the Pacific are confined to thetropics and subtropics near the convection. Ef-fects on large-scale circulations are weaker thanduring June–July.

6. Dynamical interaction betweentransient eddies and the mean flow

This study will show that the time-mean flowover the western Pacific creates a favorable en-vironment for generating and maintaining high-frequency intraseasonal wave activity. Partic-ular emphasis is placed on processes causinginteractions between eddy activities with time-scales of 10–25 days and the background time-mean flow in June–July and August.

As shown above and in FY99, tropical con-vection centered over the South China Sea islikely an important cause of the amplificationof subtropical and midlatitude waves on 10–25-day timescales. The tropical–extratropical wavestructure in the composites is representative ofstructures during the early summer. Wave am-plification can be triggered by Rossby wave en-ergy dispersion from the localized and periodicheat sources in the region. Li and Nathan(1994, 1997) and Hsu (1996) show that radia-tion of wave energy into the extratropics by en-

Y. FUKUTOMI and T. YASUNARI 319April 2002

Page 10: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

ergy advection and extraction of energy fromthe mean flow must both be considered whenexplaining the rapid growth and subsequentmaintenance of tropically forced wave anoma-lies in the subtropics and midlatitudes. Li andNathan used a spherical nondivergent baro-tropic model, and showed that the extratropicalresponse to intraseasonal tropical forcing isaffected by the background flow. If the extra-tropical response is manifest as disturbanceson the locally unstable basic flow, the distur-bances will amplify as they extract energy fromthe unstable base state. It is interesting that thedisturbances in Li and Nathan have a timescalesimilar to those in this study, although theirnumerical experiment is for a different season.

Wave anomalies that are part of the down-stream wave train retrograde (travel south-westward) from the subtropics equatorward tothe tropics over the western Pacific, as shownin the low-level composite for June–July (Fig.3). The anomalies then enhance or suppressconvection over the South China Sea. These cy-clonic or anticyclonic anomalies maintain theirstrength in the tropical and subtropical west-

erly duct region after the extreme in convectiveactivity, as described by FY99. Maintenance ofthe 10–25-day waves could be linked to inter-actions between the mean flow and the 10–25-day transients within this westerly duct. Thisequatorward-propagating wave signature is oneof the conditions necessary for exciting tropicalconvective disturbances.

The impact of the mean flow on the amplifi-cation of lower-tropospheric tropical distur-bances over the South China Sea must also beconsidered. Kawamura and Murakami (1995)investigated the role of the deformed meantropical and extratropical westerlies on eddygrowth for MJO-scale tropical disturbances.They showed that stretching and shearing de-formation in the exit region of the monsoonwesterlies (in the tropics) supplies barotropicenergy to the MJO-scale eddies. They furtheremphasized that baroclinic instability plays animportant role in eddy maintenance in extra-tropical regions. In the next section, diagnostictools based on dynamical theory are used toquantify the substantial contribution of themean flow to the 10–25-day eddy activity. Tran-

Fig. 6. (a) As in Fig. 3, but for Category 3 in August. (b) As in Fig. 4, but for Category 3 in August.

Journal of the Meteorological Society of Japan320 Vol. 80, No. 2

Page 11: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

sient eddy activity and its relationship to themean flow in both the lower and upper tropo-sphere for individual seasons are given.

6.1 Transient wave activities: E–P fluxdiagnostics

The net effect of the time-mean flow on tran-sient eddies can be analyzed in terms of theEliassen–Palm (E–P) flux, which was originallyformulated by Hoskins et al. (1983) and sub-sequently modified by Hendon and Hartmann(1985), Plumb (1986), and Trenberth (1986,1991). E–P flux vectors are a useful measure ofintraseasonal or high-frequency (synoptic scale)transient eddies. The vectors describe the shapeand propagation characteristics of eddies andtheir feedback onto the mean flow (Matthewsand Kiladis 1999a). Recently, Kiladis (1998) andMatthews and Kiladis (1999a) used Hoskins’sE vector to investigate submonthly scale (6–30-day) wave activities and tropical–extratropicalinteractions over the Pacific in NH winter. Theyshowed extratropical forcing into the tropicsover the eastern Pacific.

Hurrell and Vincent (1991) and Magana andYanai (1991) use Eu represented by Trenberth(1991) to examine the eddy–mean flow interac-tion. This formulation is defined by transformedequations derived from the primitive momen-tum, thermodynamic, and continuity equations.Hurrell and Vincent (1991) used the three-dimensional vectors to investigate the impactof upper-level subtropical westerly maxima ontransient eddies over the South Pacific. Maganaand Yanai (1991) used two-dimensional vectorsto examine the tropical–midlatitude interactionon 30–60-day (MJO) time scales in the upper-troposphere around the mid-Pacific trough.Kawamura and Murakami (1995) extendedTrenberth’s version and derived a modified Eu

for which the vertical (baroclinic) component in-cludes a horizontal eddy latent heat flux effect.They used this modified E–P flux to assessthe net effects of eddy moisture transport andorganized latent heat release by eddies, andshowed that the MJO-scale eddies are main-tained through interactions with the low-levelsummer monsoon. Sobel and Bretherton (1999)applied the barotropic component of wave activ-ity flux (Plumb 1986) to an investigation of syn-optic-scale Rossby waves in the low-level flowover the western Pacific. The radiative part of

the Plumb flux is the same as that developed byTrenberth (1986), except for a pressure weight-ing. E–P flux diagnostics are a convenient wayto examine eddy–mean flow interactions.

In this analysis, we use the two-dimensionalE vector as in Trenberth (1991). In sphericalgeometry,

Eu ¼ 1

2ðv 02 � u 02Þ;�u 0v 0

� �cos f; ð6:1Þ

where f denotes latitude. The horizontal (baro-tropic) components of Eu consist of eddy mo-mentum flux, and the horizontal vectors areparallel to the eddy group velocity relative tothe mean flow (Trenberth 1986; Magana andYanai 1991). Horizontal vectors point in the di-rection of eddy energy propagation relative tothe background flow. The vertical (baroclinic)components of Eu include the meridional eddysensible heat flux and vertical momentum flux(not shown).

The stationary Rossby wavenumber Ks isused diagnostically to represent the mean back-ground state on which transients are super-imposed. Following Hoskins and Ambrizzi(1993), this wavenumber is derived from thenondivergent perturbation vorticity equationfor time-mean flow and is defined as

Ks ¼b � Uyy

U

� �1=2

; ð6:2Þ

where U is the temporal mean of the zonalwind, b � Uyy is the temporal mean of themeridional absolute vorticity gradient, and Ks

is the wavenumber at which a barotropic Rossbywave is stationary at a particular location in agiven background zonal flow. The wavenumberacts as a refractive index for stationary Rossbywaves (Kiladis 1998; Matthews and Kiladis1999a). Originally, Ks predicted the location ofwaveguides for stationary Rossby waves (e.g.,Karoly and Hoskins 1982; Hoskins and Am-brizzi 1993). However, Ambrizzi et al. (1995),Kiladis (1998), Matthews and Kiladis (1999a)show how Ks can qualitatively describe theeffects of the base state on the propagationof high-frequency transient waves in the uppertroposphere. In this study, Ks is also used todetect westerly ducts that act as waveguidesfor 10–25-day waves in both the lower andupper troposphere. Note that Ks is a complex

Y. FUKUTOMI and T. YASUNARI 321April 2002

Page 12: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

value; Rossby waves are evanescent in theimaginary Ks region.

Figure 7 shows the 10–25-day horizontal Eu

vectors and the mean zonal winds at 850 hPain June–July (Fig. 7a), August (Fig. 7b) and at200 hPa in June–August (Fig. 7c) for 1986–1995 data over Asia and the Pacific. At 850-hPain June–July (Fig. 7a), Eu vectors point west-ward and southwestward along the subtropical

and midlatitude westerlies over the westernPacific. In addition, increased southward fluxesare present at the low-level jet core over thecentral Pacific (30�–50�N, 170�E–170�W). Eu

vectors point southwestward in the westerlyexit region of the tropical monsoon westerlies,and point westward to the east of SoutheastAsia. These vectors suggest westward andsouthwestward eddy energy propagation rela-

Fig. 7. Mean spatial distribution of Eu (m2 s�2) and zonal wind at (a) 850 hPa during June–July,(b) 850 hPa during August, and (c) 200 hPa during June–August. The contour interval is (a),(b) 2 m s�1, and (c) 5 m s�1. Regions of westerlies are shaded.

Journal of the Meteorological Society of Japan322 Vol. 80, No. 2

Page 13: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

tive to the mean flow, which is consistent witheastward and northeastward transient momen-tum transport within the tropical and subtrop-ical westerly duct. These stronger componentsof Eu, which are directed up the gradient of themean westerly flow, may be associated with theseasonal ensemble disturbance amplificationshown in Fig. 3. These flux vectors highlightthe predominance of barotropic eddy–mean flowinteractions in this area. Divergence of Eu vec-

tors can be interpreted as a forcing on the west-erly mean flow (Trenberth 1986). Such diver-gence occurs locally over the South China Seaand western Pacific north of 10�N. Convergenceof Eu implies mean easterly acceleration andoccurs over southern China, near the Equatorover the South China Sea and western Pacific(south of 10�N), and in the subtropics southeastof Japan. Figure 8a shows the distribution of Ks

at 850 hPa in the same season. Regions of real

Fig. 8. Mean spatial distribution of Ks at (a) 850 hPa during June–July, (b) 850 hPa during August,and (c) 200 hPa during June–August. Contours are drawn at zonal wavenumbers 0, 4, 5, 6, 7, 8, 10,15, 20. Regions of imaginary Ks are shaded.

Y. FUKUTOMI and T. YASUNARI 323April 2002

Page 14: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

Ks values are superimposed over climatologicalvalues for June–July tropical monsoon west-erlies in South and Southeast Asia, and sub-tropical and midlatitude westerlies over the Pa-cific. The westerlies are a preferred region forRossby wave energy propagation. Low-levelcomposite wave anomalies (Fig. 3) are locatedalong the distribution of real Ks values.

The distributions of Eu and the mean zonalwind change significantly in the lower tropo-sphere in August (Fig. 7b). One big changefrom the mean state of June–July is the east-ward penetration of tropical monsoon west-erlies into the western tropical Pacific near150�E. Another change is the disappearance ofa continuous duct (present in June–July) asthe mean westerly component over the sub-tropics to the south of the East China Sea andJapan weakens. The Eu vectors are charac-terized by a westward component directed up-stream within the mean monsoon westerlies tothe west of the South China Sea. The vectorshave the same magnitude as in June–July.Near the exit region of this westerly currentover the western tropical Pacific, where higheramplitude PKE is found (Fig. 2b), the south-westward pointing Eu vectors are of largermagnitude than in June–July. Large westwardand northwestward vectors are observed in thesubtropics near Japan, where weak mean east-erlies dominate. This region of subtropicaleasterlies corresponds to a region of imaginaryKs (Fig. 8b).

A waveguide at 200-hPa accompanying thewesterlies and the real Ks values is presentfrom June through August (Fig. 7c and Fig. 8c).A second duct exists over the tropical centraland eastern Pacific. A region of imaginary Ks

to the north of the midlatitude waveguide is aresult of the negative meridional gradient ofmean absolute vorticity. The Eu vectors pointsouthward, which indicates northward trans-port of zonal momentum due to eddies in themidlatitude waveguide. The largest vectors arenear the core of the mean westerly flow overChina and the East Asian coast (20�–50�N,100�–140�E). These southward and southwest-ward vectors converge over the East China Seaand north of the South China Sea. Eddy energypropagates southwestward from subtropics totropics in the central Pacific westerly duct(10�–25�N, 160�E–140�W). This energy propa-

gation may amplify the mid-Pacific trough asdiscussed by Magana and Yanai (1991). An-other feature to note is the group of southwardand southeastward vectors in the exit regionof the westerly jet (35�–50�N, 180�E–150�W).Enhanced signals in Eu vectors are indicativeof the eddy–mean flow interaction at these lo-cations. A separate examination of the upper-level distribution of Ks and Eu for June–Julyand for August similar to that done for Fig. 8(not shown) reveals little change in the struc-ture of the upper-level waveguide from Junethrough August.

6.2 Interaction between transients and themean flow

The barotropic Eu vector signatures asso-ciated with 10–25-day eddies along the lower-and upper-level waveguide implies transient-mean interactions in the westerly flow. Asimple energetics analysis is performed to de-scribe maintenance processes for the 10–25-daytransient wave activities forced by eddy–meaninteractions in barotropically and baroclini-cally unstable background flows. The formula-tion is obtained from the correlation betweenthe three-dimensional localized E–P flux com-ponents and the horizontal and vertical shearsof the mean zonal and meridional flows (Kawa-mura and Murakami 1995). It includes bothbarotropic and baroclinic eddy–mean interac-tions. An advantage of the formulation is thatthe effects associated with eddy moisture trans-port and the release of latent heat by eddies areincluded (see e.g., Salustri and Stone 1983).Following Kawamura and Murakami (1995),the local barotropic interaction is given by

CK ¼�� 1

2ðv 02 � u 02Þ qu

qxþ u 0v 0 qu cos f

cos fqy

þ u 0v 0 qv

qxþ 1

2ðv 02 �u 02Þ qv cos f

cos fqy

�cos f:

ð6:3Þ

Here qx ¼ a cos fql, qy ¼ aqf, a is radius ofthe earth, and l refers to longitude. Negativevalues of CK occur when kinetic energy is con-verted from the mean flow into eddies asso-ciated with the eddy momentum transportwithin the horizontal shear of the time-meanflow. The first (fourth) term of CK is the inter-action between anisotropic eddies and zonal

Journal of the Meteorological Society of Japan324 Vol. 80, No. 2

Page 15: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

(meridional) shear of the mean westerly (north-erly) flow. For example, if zonally elongatededdies (u 02 > v 02) are embedded in a regionwhere qu=qx < 0, the first term becomes nega-tive, which implies a conversion of perturbationkinetic energy from the mean. The second andthird terms are typical forms of barotropicinstability. If the eddy momentum transport(u 0v 0) is positive in the region of qv=qx < 0 andqu=qy < 0, eddies can gain kinetic energy fromthe mean flow. In this analysis, barotropic in-teractions between the 10–25-day eddies andthe time mean flow at both the 850-hPa and200-hPa levels are estimated.

The eddy–mean interactions due to dry andmoist baroclinic processes are calculated toexamine the role of local baroclinic processesin maintaining eddies associated with the low-level 10–25-day waves. These interactions aredefined as

CAS ¼ f

sv 0T 0 qu

qp� u 0T 0 qv

qp

� �cos f; ð6:4Þ

CAL ¼ L

cp

f

sv 0q 0 qu

qp� u 0q 0 qv

qp

� �cos f: ð6:5Þ

Where f is the Coriolis parameter, p is thepressure, T is the air temperature, q is thespecific humidity, L is the latent heat of con-densation, and cp is the heat capacity at a con-stant pressure. The static stability parameter(s) is calculated using s ¼ RT=cp p � ðqT=qpÞ,where R is the gas constant. These are scalarproducts between the horizontal eddy sensibleand latent heat fluxes (the horizontal part ofthree-dimensional E vectors) and the verticalshear of the mean flow.

Negative CAS and CAL represent local bar-oclinic conversions from the mean availablepotential energy to eddy available potentialenergy and are proportional to the amount ofhorizontal sensible or latent heat transport andthe vertical shear of the mean flows. The scalarproduct CAS is similar to the form used by Caiand Mak (1990) and Robinson (1993). The effectof latent heating induced by transient eddiescan be isolated in the CAL estimation. In thesubtropical and midlatitude lower tropospheres,northward eddy sensible heat and moisturefluxes (v 0T 0 > 0; v 0q 0 > 0) can exist in regions ofwesterly vertical shear (qu=qp < 0), which isconsistent with dry and moist baroclinic insta-

bility. Calculation of these values for the layerbetween 850 and 700-hPa levels follows meth-ods in Kawamura and Murakami (1995).

Figure 9 shows CK at 850 and 200 hPa. At850 hPa in June–July (Fig. 9a), large negativevalues (shaded region) extend from the SouthChina Sea to the tropical western Pacific, theexit region of the tropical monsoon westerlies.There is a local minimum over southern China.This suggests that barotropic interaction withthe mean flow is important in the growth andmaintenance of 10–25-day waves in these re-gions. Estimation of each CK term indicatesthat the first term is of primary importance inthis interaction (not shown). Cyclonic or anti-cyclonic circulations are zonally elongated inthis tropical jet exit region (Fig. 3); this distri-bution allows mean to eddy energy conversionsby the first term in CK. MJO-scale eddies aremaintained in the same way in the same region,as discussed by Kawamura and Murakami(1995). In the subtropics near Japan, however,a band of positive CK values is located alongthe low-level mean subtropical westerly flow.An energy transfer from eddy to mean flow oc-curs at this low-level subtropical jet entrance.

In August (Fig. 9b), the barotropic energyconversion from the mean flow to the 10–25-dayeddy increases over the western tropical Pacificfrom 130�–160�E, but decreases slightly overthe South China Sea. As identified in the dis-tributions of PKE and Eu, the center of dy-namical interaction between the mean flow andtransient eddies in the tropics shifts from theSouth China Sea into the western Pacific. Aweak mean to eddy energy transfer appears inthe subtropics southwest of Japan.

At 200-hPa (Fig. 9c), large positive CK valuespersist throughout JJA over East Asia and thesubtropical western Pacific. This maximum co-incides with a large area of southward pointingEu vectors (Fig. 7c) located to the north of theregion of convection (as in Fig. 3). This agreeswith eddies supplying kinetic energy to theupper-level mean flow in the region. The secondterm in the CK equation is dominant south ofthe mean westerly core. Conversely, negativevalues lie north of the upper-level jet that ex-tends from central Asia to eastern Pacific duringJJA. The 10–25-day eddies receive kinetic en-ergy from the mean within this region of nega-tive CK values.

Y. FUKUTOMI and T. YASUNARI 325April 2002

Page 16: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

Baroclinic energy processes for June–Julyare shown in Figure 10. These processes domi-nate in the maintenance of 10–25-day pertur-bations in extratropical regions just as for thecase of MJO-scale perturbations (Kawamuraand Murakami 1995). The dry baroclinic pro-cess (CAS) (Fig. 10a) shows large negativevalues over the midlatitude North Pacific andeastern Asia, with comparatively smaller nega-tive values over the subtropics near the East

Asian coast. Energy conversion from the meanavailable potential energy to the 10–25-dayeddy available potential energy is associatedwith transient sensible heat fluxes in these re-gions. However, the small negative values ofCAS over the subtropical western Pacific sug-gest that CAS is not important in maintaining10–25-day waves in this region. In addition, aband of negative CAS values lies to the north ofthe composite wavetrain (Fig. 3). Significant

Fig. 9. Mean spatial distribution of barotropic eddy–mean flow interactions (CK) at (a) 850 hPaduring June–July, (b) 850 hPa during August, and (c) 200 hPa during June–August. The contourinterval is (a), (b) 5.0 10�6 m2 s�3, and (c) 5.0 10�5 m2 s�3. Negative values are shaded.

Journal of the Meteorological Society of Japan326 Vol. 80, No. 2

Page 17: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

values of CAS are not observed in the tropics,suggesting that CAS is unimportant for 10–25-day waves in the tropical lower troposphere.

The field of moist baroclinic processes (CAL)for June–July (Fig. 10b) is characterized by azonally large area of negative values in the ex-tratropical regions of Asia and the Pacific. TheCAL contribution exceeds that of CAS in thesubtropics and midlatitudes over the entirePacific. The band of negative values is morepronounced over the North Pacific, and the arealextent of negative values of CAL is greaterthan that of CAS. This suggests that moistbaroclinic processes are more important thandry baroclinic processes in maintaining 10–25-day eddies in the subtropics and midlatitudesduring June–July. The prominence of baroclinicprocesses in the subtropics and midlatitudes isassociated with the northward eddy sensibleand latent heat transport within the mean

westerly vertical wind shear in the lower tro-posphere. This physical process is the same asfor extratropical 40-day eddies (Kawamura andMurakami 1995). There is a region of negativeCAL over the South China Sea, indicating thatmoist baroclinic eddy–mean flow interactionsoccurs locally within this very narrowly definedregion.

The August distribution of baroclinically-forced eddy–mean flow interactions showsslight differences to that of June–July. The bandof negative CAS values along the midlatitudewesterlies shifts slightly to higher latitudes.This poleward shift is also observed in the neg-ative CAL to the east of Japan. These shiftsmight be caused by a decrease in westerly ver-tical shear of the mean zonal wind in the sub-tropics and lower-midlatitudes, which is attrib-uted to the northward shift of mean westerliesand the disappearance of a mean westerly com-

Fig. 10. Mean spatial distribution of (a) dry baroclinic (CAS) and (b) moist baroclinic (CAL) eddy–mean flow interactions. Contour intervals are 2.0 10�5 m2 s�3. Negative values are shaded.

Y. FUKUTOMI and T. YASUNARI 327April 2002

Page 18: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

ponent in the subtopics. A negative CAL regionoverspreads the subtropical Pacific east of140�E.

7. Summary and discussion

The structure and maintenance of 10–25-daydisturbances over the western Pacific duringsummer were investigated. The disturbanceswere examined for two different seasons,namely June–July and August. The character-istics and behavior of anomalies in compositesare identified for individual seasons and com-pared between seasons. E–P flux diagnosticsand an energetics analysis reveal the roles ofbarotropic and baroclinic interactions betweeneddy and time-mean flow in maintaining tran-sient wave activities in the lower troposphere.In particular, the forcing of transient eddies onthe time-mean flow in the upper troposphere isalso described.

Composites for June–July suggest significantrelationships between tropical convection andextratropical circulations with timescales of 10–25 days. FY99 and this paper both suggest thattropical convection over the South China Seacan force subtropical and midlatitude circula-tions through a downstream Rossby wave re-sponse. Subsequent to the maximum enhance-ment (suppression) of convection, a downstreamwave train is well developed toward the NorthPacific along the subtropical and midlatitudewesterlies in both the lower and upper tropo-sphere. In the lower troposphere, a subtropicaltrough (ridge) is strengthened as part of thiswavetrain. Subtropical anomalies in this wave-train then move southwestward into the tropics,and initiate subsequent suppression (enhance-ment) of convection in the South China Searegion (cf., FY99). This two-way interactionbetween the tropics and the extratropics is adominant feature of the convectively coupled10–25-day waves during June–July.

August composites show the large-scalestructure and behavior of circulation anomaliesin the lower-troposphere to be considerably dif-ferent from June–July. The downstream wave-train emanating from the South China Sea ispoorly defined in the August composites. More-over, there is no evidence of southwestwardanomaly propagation around this base region.The composite circulation anomalies have west-ward and northwestward tracking character-

istics. A two-way interaction mechanism doesnot become established in the lower troposphereduring August.

The difference in the structure of the meanbackground flow between seasons may be re-sponsible for the differences in the compositestructures in the lower troposphere betweenJune–July and August. During June–July, con-tinuous mean westerlies extend from the trop-ics to midlatitudes in the lower troposphere.FY99 and this study suggest that this region ofwesterlies is a waveguide for the Rossby wave-train and its southwestward phase propagationfrom the subtropics into the tropics over thewestern Pacific. During August, however, thestructure of the low-level mean flow changesand the subtropical portion of the duct van-ishes. This inhibits subtropic to tropic phasepropagation. In addition, the absence of thearc-like downstream wavetrain may be relatedto the large-scale basic state, which includes adiscontinuous waveguide between tropics andextratropics. The sensitivity of the extratropicalresponse to transient convection acting on abasic state, and the phase propagation of forcedwaves can be checked using GCMs, as done byJin and Hoskins (1995), Newman and Sardesh-mukh (1998), and Matthews and Kiladis(1999b).

Eddy–mean flow interactions that maintainthe 10–25-day waves are investigated. In thelower troposphere, barotropic processes main-tain the 10–25-day eddies in the exit region ofthe monsoon westerly flow in the tropical west-ern Pacific. Convectively coupled rotational dis-turbances are amplified in this region, not onlyby diabatic effects such as external heating(or cooling) associated with convection, but alsoby barotropic eddy–mean flow interactions. Incontrast, the 10–25-day eddies supply kineticenergy to the mean flow over the subtropicalwestern Pacific in June–July. Dry and moistbaroclinic processes maintain the 10–25-dayeddies in the subtropics and midlatitudes. Themain contributor to the growth of the south-westward propagating waves during June–Julyis condensational heating produced by eddiesembedded within a westerly vertical shear, es-pecially in the subtropics. This subtropical ductwith westerly vertical shear can supportfavorable conditions for wave maintenance.Baroclinic growth of the 10–25-day eddies

Journal of the Meteorological Society of Japan328 Vol. 80, No. 2

Page 19: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

decreases over the subtropics south of Japanduring August reflecting the weaker verticalshear in that season.

An energetics analysis shows that baroclinicprocesses associated with eddy sensible and la-tent heat transport are minor contributors tothe maintenance or amplification of the low-level 10–25-day waves in the tropical domain.Barotropic processes dominate eddy–mean flowinteractions in the tropical lower troposphere.Our results emphasize that barotropic processesand diabatic processes forced by convection areimportant in amplifying low-level waves in thisregion. A complete three-dimensional energeticsanalysis using composite anomalies and theirlife cycles, as done by Lau and Lau (1992), couldclarify the relative roles of barotropic and bar-oclinic processes.

E–P flux diagnostics and energetics for thelower troposphere suggest that eddy energy ac-cumulates in the exit region of tropical mon-soon westerlies. This energy accumulation maybe related to phase propagation of wave anoma-lies from other regions, as shown in the compo-sites, and the energy can be enhanced throughbarotropic interactions with the mean flow.Negative longitudinal stretching shear of thezonal wind (qu=qx < 0) is effective for baro-tropic energy conversions from the mean flow toeddies in the exit region of the westerly mon-soon. In particular, as a Rossby response to theconvective heating, a wavetrain emanates fromthis energy-accumulation region into the extra-tropics during June–July. Webster and Chang(1988) proposed a similar energy accumulation–emanation mechanism.

Eddy–mean flow interactions in the uppertroposphere have been analyzed. The 10–25-day eddies act as an energy source for the meanflow in the subtropics north of the tropical con-vection from 100�E to 150�E in the summer.There are large southward pointing Eu vectorsand barotropic conversion from eddy to meankinetic energy over the subtropics in this longi-tudinal sector, and its location roughly corre-sponds to that of the amplified waves in thecomposites (Fig. 4). Enhanced eddies centeredover the South China Sea produced by convec-tion at this timescale may be an important en-ergy source for the upper-level subtropical meanflow, especially in June–July.

We have focused on the structure and main-

tenance of the 10–25-day disturbances over thewestern pacific domain in Northern Hemispheresummer, leaving other subjects to be examinedlater. As discussed by Hartmann et al. (1992),Schnadt et al. (1998), Heta (1990), and Chenand Weng (1999), the relationship between theoccurrence of tropical cyclones and this higher-frequency intraseasonal oscillation in the trop-ics is an important subject for future research.Furthermore, the interaction between high fre-quency eddies and the MJO will be demon-strated in the future, as was done with theNorthern Hemisphere winter case by Matthewsand Kiladis (1999a). Scale interactions asso-ciated with the low-frequency waves, as dem-onstrated by Higgins and Schubert (1993, 1994)in the summer over Asia and the Pacific, re-quire further study. Since our focus was on thecharacteristics of the 10–25-day disturbance it-self, scale interactions were not considered here.The interannual variability of convection andthe activity of transient eddies on this timescaleand their relation to ENSO and the troposphericbiennial oscillation is also an interesting topicfor future research.

Acknowledgments

We wish to acknowledge the insightful com-ments and suggestions from anonymous re-viewers and Dr. N. Iwasaka. They were ex-tremely valuable in revising an earlier versionof this manuscript. We would also like to thankDr. R. Kawamura for encouragement and use-ful discussion. This paper is part of the Ph.D.thesis of Y. Fukutomi of the Institute of Geo-science, University of Tsukuba. ECMWF anal-yses were kindly provided by Dr. H.L. Tanakaand are the same as those used by Tanaka andKimura (1997). The OLR data were obtainedfrom the NOAA Climate Diagnostics Center(http://www.cdc.noaa.gov). Figures were madewith the Generic Mapping Tools (GMT) graph-ics system developed by Wessel and Smith(1995). This study was supported in part by aGrant-in-Aid for Scientific Research on PriorityAreas (B) from the Japanese Ministry of Educa-tion, Science, Sports, and Culture, No. 11201101(Energy and Water Cycle in Asian MonsoonRegion), and was completed under the auspicesof the Hydrological Cycle Research Program ofthe FRSGC.

Y. FUKUTOMI and T. YASUNARI 329April 2002

Page 20: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

References

Ambrizzi, T., B.J. Hoskins, and H.-H. Hsu, 1995:Rossby wave propagation and teleconnectionpatterns in the austral winter. J. Atmos. Sci.,52, 3661–3672.

Cai, M. and M. Mak, 1990: On the basic dynamics ofregional cyclogenesis. J. Atmos. Sci., 47, 1417–1442.

Chen, T.-C. and J.-M. Chen, 1995: An observationalstudy of the South China Sea monsoon duringthe 1979 summer: Onset and life cycle. Mon.Wea. Rev., 123, 2295–2318.

——— and S.-P. Weng, 1999: Interannual and intra-seasonal variations in monsoon depressionsand their westward-propagating predecessors.Mon. Wea. Rev., 127, 1005–1020.

Fukutomi, Y. and T. Yasunari, 1999: 10–25-day in-traseasonal variations of convection and circu-lation over East Asia and western NorthPacific during early summer. J. Meteor. Soc.Japan, 77, 753–769.

———, 2000: Tropical–extratropical interactions onthe intraseasonal timescales over the westernPacific during early summer. Ph.D. thesis, In-stitute of Geoscience, University of Tsukuba,162 pp.

Hartmann, D.L., M.L. Michelsen, and S.A. Klein,1992: Seasonal variations of tropical intra-seasonal oscillations: A 20–25-day oscillationin the western Pacific. J. Atmos. Sci., 49, 1277–1289.

Hendon, H.H. and D.L. Hartmann, 1985: Variabilityin a nonlinear model of the atmosphere withzonally symmetric forcing. J. Atmos. Sci., 42,2783–2797.

Heta, Y., 1990: An analysis of tropical wind fields inrelation to typhoon formation over the westernPacific. J. Meteor. Soc. Japan, 68, 65–77.

Higgins, R.W. and S.D. Schubert, 1993: Low-fre-quency synoptic-eddy activity in the Pacificstorm track. J. Atmos. Sci., 50, 1672–1690.

——— and ———, 1994: Simulated life cycles ofpersistent anticyclonic anomalies over theNorth Pacific: role of synoptic-scale eddies. J.Atmos. Sci., 51, 3238–3260.

Hoskins, B.J., I.N. James, and G.H. White, 1983: Theshape, propagation and mean-flow interactionof large-scale weather systems. J. Atmos. Sci.,40, 1595–1612.

——— and T. Ambrizzi, 1993: Rossby wave propa-gation on a realistic longitudinally varyingflow. J. Atmos. Sci., 50, 1661–1671.

Hsu, H.-H., 1996: Global view of the intraseasonaloscillation during northern winter. J. Climate,96, 2386–2406.

Hurrell, J.W. and D.G. Vincent, 1991: On the main-tenance of short-term subtropical wind max-ima in the Southern Hemisphere during SOP-1, FGGE. J. Climate, 4, 1009–1022.

Jin, F. and B.J. Hoskins, 1995: The direct responseto tropical heating in a baroclinic atmosphere.J. Atmos. Sci., 52, 307–319.

Karoly, D.J. and B.J. Hoskins, 1982: Three-dimensional propagation of planetary waves. J.Meteor. Soc. Japan, 60, 109–123.

Kawamura, R. and T. Murakami, 1995: Interactionbetween the mean summer monsoon flow and45-day transient perturbations. J. Meteor. Soc.Japan, 73, 1087–1114.

Kaylor, R.E., 1977: Filtering and decimation of digi-tal time series. Tech. Note BN 850, Instituteof Physical Science Technology, University ofMaryland, College Park, 42 pp.

Kiladis, G.N. and K.M. Weickmann, 1992a: Circula-tion anomalies associated with tropical convec-tion during northern winter. Mon. Wea. Rev.,120, 1900–1923.

——— and ———, 1992b: Extratropical forcing oftropical convection during northern winter.Mon. Wea. Rev., 120, 1924–1938.

———, G.A. Meehl, and K.M. Weickmann, 1994:Large-scale circulation associated with west-erly wind bursts and deep convection over thewestern equatorial Pacific. J. Geophys. Res.,99(D9), 18527–18544.

——— and M. Wheeler, 1995: Horizontal and verti-cal structure of observed tropospheric equato-rial Rossby waves. J. Geophys. Res., 100(D11),22981–22997.

——— and K.M. Weickmann, 1997: Horizontalstructure and seasonality of large-scale circu-lations associated with submonthly tropicalconvection. Mon. Wea. Rev., 125, 1997–2013.

———, 1998: Observations of Rossby waves linked toconvection over the eastern tropical Pacific. J.Atmos. Sci., 55, 321–339.

Lau, K.-H. and N.-C. Lau, 1990: Observed structureand propagation characteristics of tropicalsummertime synoptic scale disturbances. Mon.Wea. Rev., 118, 1888–1913.

——— and ———, 1992: The energetics and propa-gation dynamics of tropical summertime syn-optic-scale disturbances. Mon. Wea. Rev., 120,2523–2539.

Li, L. and T.R. Nathan, 1994: The global atmo-spheric response to low-frequency tropical forc-ing: Zonally averaged basic states. J. Atmos.Sci., 51, 3412–3426.

——— and ———, 1997: Effects of low-frequencyTropical Forcing on intraseasonal tropical–extratropical interactions. J. Atmos. Sci., 54,332–346.

Journal of the Meteorological Society of Japan330 Vol. 80, No. 2

Page 21: Tropical–Extratropical Interaction Associated with the 10 ...mausam.hyarc.nagoya-u.ac.jp/~yasunari/list/pdf/... · In an earlier study, Fukutomi and Yasunari (1999, hereafter FY99)

Liebmann, B., G.N. Kiladis, J.A. Marengo, T. Am-brizzi, and J.D. Glick, 1999: Submonthly con-vective variability over South America and thesouth Atlantic convergence zone. J. Climate,12, 1877–1891.

Madden, R.A. and P.R. Julian, 1972: Description ofglobal scale circulation cells in the tropics with a40–50 day period. J. Atmos. Sci., 29, 1109–1123.

——— and ———, 1994: Observations of the 40–50-day tropical oscillation: A review. Mon. Wea.Rev., 122, 814–837.

Magana, V. and M. Yanai, 1991: Tropical–midlatitude interaction on the time scale of 30to 60 days during the northern summer of1979. J. Climate, 4, 180–201.

Matsumoto, J., 1992: The seasonal changes in Asianand Australian monsoon regions. J. Meteor.Soc. Japan, 70, 257–273.

Matthews, A.J. and G.N. Kiladis, 1999a: Thetropical–extratropical interaction betweenhigh-frequency transients and the Madden–Julian oscillation. Mon. Wea. Rev., 127, 661–677.

——— and ———, 1999b: Interactions betweenENSO, transient circulation, and tropical con-vection over the Pacific. J. Climate, 12, 3062–3086.

Meehl, G.A., G.N. Kiladis, K.M. Weickmann, M.Wheeler, D.S. Gutzler, and G.P. Compo, 1996:Modulation of equatorial subseasonal convec-tive episodes by tropical–extratropical interac-tion in the Indian and Pacific Ocean regions. J.Geophys. Res., 101(D10), 15033–15049.

Mo, K.C., 1999: Alternating wet and dry episodesover California and intraseasonal oscillations.Mon. Wea. Rev., 127, 2759–2776.

———, 2000: Intraseasonal modulations of summerprecipitation over North America. Mon. Wea.Rev., 128, 1490–1505.

Murakami, T. and J. Matsumoto, 1994: Summermonsoon over the Asian continent and West-ern Pacific. J. Meteor. Soc. Japan, 72, 719–745.

Newman, M. and P.D. Sardeshmukh, 1998: The im-pact of the annual cycle on the North Pacific/North American response to remote low-frequency forcing. J. Climate, 11, 2491–2504.

Numaguti, A., 1995: Characteristics of 4-to-20-day-period disturbances observed in the equatorialPacific during the TOGA COARE IOP. J.Meteor. Soc. Japan, 73, 353–377.

Plumb, R.A., 1986: Three-dimension propagation oftransient quasi-geostrophic eddies and its re-lationship with the eddy forcing of the time-mean flow. J. Atmos. Sci., 43, 1657–1678.

Robinson, W.A., 1993: Mechanisms of low-frequency

variability in a simple model with orography.J. Atmos. Sci., 50, 878–888.

Salustri, G. and P.H. Stone, 1983: A diagnostic studyof the forcing of the Ferrel cell by eddies withlatent heat effects included. J. Atmos. Sci., 40,1101–1109.

Schrage, J.M. and D.G. Vincent, 1996: Tropical con-vection on 7–21 day timescales over the West-ern Pacific. J. Climate, 9, 587–607.

Schnadt, C., A. Fink, D.G. Vincent, J.M. Schrage,and P. Speth, 1998: Tropical cyclones, 6–25day oscillations, and tropical–extratropical in-teraction over the northwestern Pacific. Me-teor. Atmos. Phys., 68, 151–169.

Sobel, A.H. and C.S. Bretherton, 1999: Developmentof synoptic-scale disturbances over the sum-mertime tropical northwest Pacific. J. Atmos.Sci., 56, 3106–3127.

Tanaka, H.L. and K. Kimura, 1998: Time seriesanalysis of natural variability in barotropicenergy of the atmosphere with ECMWF globalanalysis. J. Meteor. Soc. Japan, 76, 207–274.

Trenberth, K.E., 1986: An assessment of the impactof transient eddies on zonal flow during ablocking episode using localized Eliassen–Palm flux diagnostics. J. Atmos. Sci., 43, 2070–2087.

———, 1991: Storm tracks in the Southern Hemi-sphere. J. Atmos. Sci., 48, 2159–2178.

——— and J.G. Olson, 1988: An evaluation and in-tercomparison of global analyses from the Na-tional Meteorological Center and the EuropeanCenter for Medium Range Weather Forecasts.Bull. Amer. Meteor. Soc., 69, 1047–1057.

Vincent, D.G., K.-C. Ko, and J.M. Schrage, 1997:Subtropical jet streaks over the South Pacific.Mon. Wea. Rev., 125, 438–447.

———, A. Fink, J.M. Schrage, and P. Speth, 1998:High- and low-frequency intraseasonal vari-ance of OLR on annual and ENSO timescales.J. Climate, 11, 968–986.

Webster, P.J. and H.-R. Chang, 1988: Equatorial en-ergy accumulation and emanation regions: Im-pacts of a zonally varying basic state. J. Atmos.Sci., 45, 803–829.

——— and S. Yang, 1989: The three-dimensionalstructure of perturbation kinetic energy and itsrelationship to the zonal wind field. J. Climate,2, 1210–1222.

Wheeler, M. and G.N. Kiladis, 1999: Convectivelycoupled equatorial waves: Analysis of cloudsand temperature in the wavenumber-frequencydomain. J. Atmos. Sci., 56, 374–399.

Wessel, P. and W.H.F. Smith, 1995: New version ofthe Generic Mapping Tools released. EOSTrans. Amer. Geophys. U., 76, 329.

Y. FUKUTOMI and T. YASUNARI 331April 2002