13
Tropical Cyclogenesis Associated with Rossby Wave Energy Dispersion of a Preexisting Typhoon. Part I: Satellite Data Analyses* TIM LI AND BING FU Department of Meteorology, and International Pacific Research Center, University of Hawaii at Manoa, Honolulu, Hawaii (Manuscript submitted 20 September 2004, in final form 7 June 2005) ABSTRACT The structure and evolution characteristics of Rossby wave trains induced by tropical cyclone (TC) energy dispersion are revealed based on the Quick Scatterometer (QuikSCAT) and Tropical Rainfall Measuring Mission (TRMM) Microwave Imager (TMI) data. Among 34 cyclogenesis cases analyzed in the western North Pacific during 2000–01 typhoon seasons, six cases are associated with the Rossby wave energy dispersion of a preexisting TC. The wave trains are oriented in a northwest–southeast direction, with alternating cyclonic and anticyclonic vorticity circulation. A typical wavelength of the wave train is about 2500 km. The TC genesis is observed in the cyclonic circulation region of the wave train, possibly through a scale contraction process. The satellite data analyses reveal that not all TCs have a Rossby wave train in their wakes. The occur- rence of the Rossby wave train depends to a certain extent on the TC intensity and the background flow. Whether or not a Rossby wave train can finally lead to cyclogenesis depends on large-scale dynamic and thermodynamic conditions related to both the change of the seasonal mean state and the phase of the tropical intraseasonal oscillation. Stronger low-level convergence and cyclonic vorticity, weaker vertical shear, and greater midtropospheric moisture are among the favorable large-scale conditions. The rebuilding process of a conditional unstable stratification is important in regulating the frequency of TC genesis. 1. Introduction Tropical cyclone (TC) genesis is the transformation process of a tropical disturbance to a warm-core cy- clonic system. Except for the southeastern Pacific and the southeastern Atlantic where cold SST tongues (and excessively large vertical shears) are pronounced, TC genesis is observed over all other tropical ocean basins. A statistical TC distribution map (Gray 1968) showed that the western North Pacific (WNP) experiences the most frequent TC activity among eight TC genesis re- gions. The average annual number of TCs in the WNP is 22, which contributes 36% of the global total TC numbers. Both the western North Atlantic and South Pacific have an average of seven TCs, while the Bay of Bengal, South Indian Ocean and eastern North Pacific have six TCs. The Arabian Sea and the sea off the northwest Australian coast have an average of three cyclogenesis events each year. To explain the geo- graphic distribution of TC genesis, Gray (1975, 1977) investigated the environmental conditions around the cyclogenesis regions and found that six physical param- eters can be used to evaluate the potential tendency for TC genesis. Based on these parameters, the favorable conditions for TC genesis include positive relative vor- ticity in lower troposphere, a location away from the equator, a warm SST above 26.1°C, a small vertical shear, a large equivalent potential temperature ( e ) ra- dient between 500 hPa and the surface, and a relatively large value of relative humidity in the middle tropo- sphere. Ritchie and Holland (1999) documented environ- mental flow regimes that favor TC genesis in the WNP. They are the monsoon shear line, the monsoon conflu- ence region, and the monsoon gyre. While these favor- able large-scale flow patterns are important in provid- ing background stages for tropical storm development, they are not sufficient to lead to cyclogenesis. Although these patterns are often seen in daily weather maps, cyclogenesis does not occur every day. This implies that * SOEST Contribution Number 6740 and IPRC Contribution Number 354. Corresponding author address: Prof. Tim Li, IPRC, and Dept. of Meteorology, University of Hawaii at Manoa, 2525 Correa Rd., Honolulu, HI 96822. E-mail: [email protected] VOLUME 63 JOURNAL OF THE ATMOSPHERIC SCIENCES MAY 2006 © 2006 American Meteorological Society 1377 JAS3692

Tropical Cyclogenesis Associated with Rossby Wave Energy

  • Upload
    others

  • View
    19

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Tropical Cyclogenesis Associated with Rossby Wave Energy

Tropical Cyclogenesis Associated with Rossby Wave Energy Dispersion of aPreexisting Typhoon. Part I: Satellite Data Analyses*

TIM LI AND BING FU

Department of Meteorology, and International Pacific Research Center, University of Hawaii at Manoa, Honolulu, Hawaii

(Manuscript submitted 20 September 2004, in final form 7 June 2005)

ABSTRACT

The structure and evolution characteristics of Rossby wave trains induced by tropical cyclone (TC)energy dispersion are revealed based on the Quick Scatterometer (QuikSCAT) and Tropical RainfallMeasuring Mission (TRMM) Microwave Imager (TMI) data. Among 34 cyclogenesis cases analyzed in thewestern North Pacific during 2000–01 typhoon seasons, six cases are associated with the Rossby wave energydispersion of a preexisting TC. The wave trains are oriented in a northwest–southeast direction, withalternating cyclonic and anticyclonic vorticity circulation. A typical wavelength of the wave train is about2500 km. The TC genesis is observed in the cyclonic circulation region of the wave train, possibly througha scale contraction process.

The satellite data analyses reveal that not all TCs have a Rossby wave train in their wakes. The occur-rence of the Rossby wave train depends to a certain extent on the TC intensity and the background flow.Whether or not a Rossby wave train can finally lead to cyclogenesis depends on large-scale dynamic andthermodynamic conditions related to both the change of the seasonal mean state and the phase of thetropical intraseasonal oscillation. Stronger low-level convergence and cyclonic vorticity, weaker verticalshear, and greater midtropospheric moisture are among the favorable large-scale conditions. The rebuildingprocess of a conditional unstable stratification is important in regulating the frequency of TC genesis.

1. Introduction

Tropical cyclone (TC) genesis is the transformationprocess of a tropical disturbance to a warm-core cy-clonic system. Except for the southeastern Pacific andthe southeastern Atlantic where cold SST tongues (andexcessively large vertical shears) are pronounced, TCgenesis is observed over all other tropical ocean basins.A statistical TC distribution map (Gray 1968) showedthat the western North Pacific (WNP) experiences themost frequent TC activity among eight TC genesis re-gions. The average annual number of TCs in the WNPis 22, which contributes 36% of the global total TCnumbers. Both the western North Atlantic and SouthPacific have an average of seven TCs, while the Bay ofBengal, South Indian Ocean and eastern North Pacific

have six TCs. The Arabian Sea and the sea off thenorthwest Australian coast have an average of threecyclogenesis events each year. To explain the geo-graphic distribution of TC genesis, Gray (1975, 1977)investigated the environmental conditions around thecyclogenesis regions and found that six physical param-eters can be used to evaluate the potential tendency forTC genesis. Based on these parameters, the favorableconditions for TC genesis include positive relative vor-ticity in lower troposphere, a location away from theequator, a warm SST above 26.1°C, a small verticalshear, a large equivalent potential temperature (�e) ra-dient between 500 hPa and the surface, and a relativelylarge value of relative humidity in the middle tropo-sphere.

Ritchie and Holland (1999) documented environ-mental flow regimes that favor TC genesis in the WNP.They are the monsoon shear line, the monsoon conflu-ence region, and the monsoon gyre. While these favor-able large-scale flow patterns are important in provid-ing background stages for tropical storm development,they are not sufficient to lead to cyclogenesis. Althoughthese patterns are often seen in daily weather maps,cyclogenesis does not occur every day. This implies that

* SOEST Contribution Number 6740 and IPRC ContributionNumber 354.

Corresponding author address: Prof. Tim Li, IPRC, and Dept.of Meteorology, University of Hawaii at Manoa, 2525 Correa Rd.,Honolulu, HI 96822.E-mail: [email protected]

VOLUME 63 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S MAY 2006

© 2006 American Meteorological Society 1377

JAS3692

Page 2: Tropical Cyclogenesis Associated with Rossby Wave Energy

there must be some triggering mechanisms for cyclo-genesis. Previous studies pointed out various synopticperturbation triggering scenarios such as the Rossbywave energy dispersion of a preexisting TC (Frank1982; Davidson and Hendon 1989; Ritchie and Holland1997; Briegel and Frank 1997), energy accumulation ofeasterly waves in a confluent mean zonal flow (Kuo etal. 2001), and a mixed Rossby–gravity wave packet(Dickinson and Molinari 2002). Tropical cyclone gen-esis may also involve wave–mean flow interactions(Ferreira and Schubert 1997; Zehnder et al. 1999; Mo-linari et al. 2000) or development of synoptic-scalewave trains (Lau and Lau 1990; Chang et al. 1996).Sobel and Bretherton (1999) analyzed the energy fluxconvergence in the WNP and found that wave accumu-lation may operate either on waves coming from out-side of the region or in situ.

In this study we particularly focus on the scenariorelated to TC energy dispersion (TCED). As we know,a mature TC is subject to Rossby wave energy disper-sion because of the change of the Coriolis force withlatitude (e.g., Anthes 1982; Flierl 1984; Luo 1994; Mc-Donald 1998). While a TC moves northwestward be-cause of mean flow steering and the beta drift, it emitsRossby wave energy southeastward, forming a synop-tic-scale wave train with alternating anticyclonic andcyclonic vorticity perturbations (Carr and Elsberry1994, 1995). Using barotropic vorticity equation mod-els, Chan and Williams (1987) and Fiorino and Elsberry(1989) showed that a monopole, cyclonic vortex on abeta plane experiences Rossby wave energy dispersionthat generates asymmetric structure with a cyclonicgyre to the west and an anticyclonic gyre to the east ofthe vortex center. While the motion of the vortex as-sociated with the beta drift is toward the northwest, thevortex is subject to energy dispersion due to Rossbywave radiation (Flierl 1984). Luo (1994) focused on theeffect of Rossby energy dispersion on the motion andstructure of a TC-scale vortex on a beta plane in anondivergent barotropic model. A wavelike wake isformed in his model as a result of Rossby energy dis-persion from the vortex. Because of the leaking of en-ergy, the vortex weakens during its northwestwardmovement. Carr and Elsberry (1995) found the exis-tence of TCED-induced Rossby wave trains in both abarotropic model and Naval Operational Global Atmo-spheric Prediction System (NOGAPS) forecast fields.Holland (1995) discussed how sequential TCs might begenerated because of TCED in the WNP. Once a TCdevelops in the vicinity of the monsoon trough/ITCZ, itmoves poleward and westward because of the planetaryvorticity gradient and mean flow steering; group energyassociated with the TC propagates eastward and equa-

torward, forming a synoptic-scale wave train of alter-nating regions of anticyclonic and cyclonic vorticity per-turbations. The perturbation energy, being accumu-lated in the confluence zone, helps release thebarotropic/baroclinic instability in the ITCZ. Cumulusconvection is organized in the region of cyclonic circu-lation, amplifying the disturbances through the convec-tion–frictional convergence feedback (Wang and Li1994). As the process proceeds, a TC-strength vortexmay be generated. Once the newly generated vortexmoves away from the region, it may trigger the newdevelopment of disturbances through the same energydispersion/accumulation processes.

The role of TCED on cyclogenesis was previouslysuggested based on either simple numerical models orindirect observational data (e.g., global model analysisproducts) over open oceans. So far, no direct observa-tional evidence shows the detailed structure and evolu-tion characteristics of the Rossby wave train induced byTCED, nor its relationship with TC genesis. Recentsatellite products [such as the Quick Scatterometer(QuikSCAT)] provide a great opportunity to reveal thedetailed pattern and evolution of the Rossby wave trainand associated cyclogenesis processes (e.g., Li et al.2003). The objective of this paper is to document fromdirect satellite measurements the detailed structure andevolution characteristics of the Rossby wave train inassociation with TCED and its role in cyclogenesis.While Part I is focused on the observed aspects, Li et al.(2006, hereafter Part II) will simulate the TCED-induced cyclogenesis in a 3D model.

Part I is organized as follows. In section 2 we brieflydescribe the data and methodology. In section 3 wepresent the observed structure and evolution character-istics of the Rossby wave train. Then in section 4 weexamine possible factors that control the generation ofthe Rossby wave train. The large-scale dynamic andthermodynamic control on TCED-induced cyclogenesisis discussed in section 5. Finally, conclusions are givenin section 6.

2. Data and methodology

The primary data used in this study are the NationalAeronautics and Space Administration (NASA)QuikSCAT level 3 daily surface wind data and theTropical Rainfall Measurement Mission (TRMM)Microwave Imager (TMI) cloud liquid water data. TheQuikSCAT data are retrieved by a microwave scatter-ometer named SeaWinds that was launched on theQuikBird satellite in 1999. The primary mission ofQuikSCAT is to measure near-surface winds. By de-tecting the sea surface roughness, it can use some em-pirical relationship to calculate the 10-m surface wind

1378 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 3: Tropical Cyclogenesis Associated with Rossby Wave Energy

speed and direction. The products from this scatterom-eter include 10-m surface zonal and meridional windspeed daily data. It covers global oceans with a resolu-tion of 0.25° � 0.25°. The TMI data are retrieved by aradiometer on board the TRMM satellite. It is a jointmission supported by NASA and the National SpaceDevelopment Agency of Japan. The TMI data coverthe tropical region between 40°S–40°N and provide10-m surface wind speed (11 and 37 GHz), sea surfacetemperature, atmospheric water vapor, cloud liquid wa-ter, and precipitation rates with a horizontal resolutionof 0.25° � 0.25°.

In addition, National Centers for Environmental Pre-diction–National Center of Atmosphere Research(NCEP–NCAR) reanalysis data are used. Having ahorizontal resolution of 2.5° � 2.5°, the reanalysis dataare used to fill in gaps in satellite observations with alinear optimal interpolation scheme. The reanalysisdata are also used to construct the vertical structure ofthe wave train in section 3.

We focus on the cyclogenesis events associated withTCED in the WNP during the 2000–01 typhoon seasons(from 1 June to 30 September). The TC best-track dataissued by Joint Typhoon Warning Center (JTWC) areused to determine the cyclogenesis location. To clearlyexamine synoptic-scale wind structures prior to TC for-mation, a time filtering method is used to retain 3–8-daysignals for both QuikSCAT and TMI data. The selec-tion of the 3–8-day bandwidth is based on Lau and Lau(1990, 1992), who noted a significant synoptic-scalewave spectrum appearing in this frequency band in theWNP. The filtering scheme is that of Murakami (1979).The response function for this filter is a Gaussian func-tion. In the 3–8-day band, the maximum response is ata period of about 5 days. A low-pass filter (with a pe-riod greater than 20 days) was used to compose large-scale background conditions associated with cyclogen-esis and noncyclogenesis cases in section 5.

The TCED-induced cyclogenesis events are identi-fied based on the following analysis approach. Thedaily synoptic maps of the filtered surface wind fieldsfrom the QuikSCAT data are first examined. If a newTC formed in the cyclonic circulation embedded in awave train produced by a preexisting TC (based onJTWC best-track data), then this new TC belongs to theTCED-induced cyclogenesis scenario. Note that thisanalysis approach neglects possible triggering processesfrom the upper troposphere.

3. Rossby wave train structure

During the summers of 2000–01, we analyzed 34 cy-clogenesis cases in the WNP in which a tropical pertur-

bation finally reached or exceeded tropical storm inten-sity. Among the 34 cases, 6 cyclogenesis cases are iden-tified to be associated with the energy dispersion of apreexisting TC. Figure 1 illustrates such a case. It showsthe evolution of the 3–8-day filtered QuikSCAT windfields from 1 to 9 August 2000. The letter “A” in Fig. 1represents the location of Typhoon Jelawat that formedon 1 August. During the first 2–3 days, because of itsweak intensity, Jelawat did not generate a visibleRossby wave train in its wake. During its northwest-ward journey, the intensity of Jelawat increased. Withits steady intensification, the Rossby wave train becamemore and more evident in its wake. On 6 August, aclear wave train was observed. This Rossby wave trainhas a zonal wavelength of about 2500 km, oriented in anorthwest–southeast direction.

The most remarkable characteristic of this wave trainis that it has a large meridional wavelength. A halfmeridional wavelength is about 4000 km on 6 August.The meridional wavelength of the wave train is greatlyreduced in subsequent days, leading to generation of anew TC named Ewiniar on 9 August. Note that TCEwiniar (represented by the letter “B” in Fig. 1) formedin the cyclonic vorticity center of the Rossby wave train.

The cloud liquid water fields show a similar evolutionfeature (Fig. 2). During the first 2–3 days, convectiveclouds (represented by the cloud liquid water field)were loosely organized. On 3 August, there was anorthwest–southeast-oriented cloud band in the wakeof Jelawat. On 6 August, when the wave train patternwas discerned in the wind field, convective clouds alsostarted to be organized in such a way that the cloudliquid water concentrated in cyclonic circulation re-gions. With the further development of the wave train,one can see clearly the alternating cloudy–clear–cloudypattern, corresponding well to the alternating cyclonic–anticyclonic–cyclonic circulation. Thus, the indepen-dent satellite products provide consistent wind andcloud patterns during the course of the wave train de-velopment.

To clearly demonstrate the role of the Rossby waveenergy dispersion, we calculated an E vector (an energypropagation vector) for TC Jelawat (see Fig. 3), basedon Trenberth (1986). Here, E � ([�u�u� � ����]/ 2,[�u���]), where [ ] represents time averaging, and u�and �� are zonal and meridional components of synop-tic-scale wind perturbations, respectively. The E vectorin Fig. 3 was calculated based on a 15-day period cen-tered on 5 August 2000. It indicates the direction ofenergy dispersion around 5 August, assuming that thewave train is a Rossby wave type. Figure 3 illustratesthat there is indeed southeastward propagation of

MAY 2006 L I A N D F U 1379

Page 4: Tropical Cyclogenesis Associated with Rossby Wave Energy

Rossby wave energy. The E vector is not sensitive tothe averaged period, say, from 9 to 21 days.

The second example of the TCED scenario is thegenesis of Typhoon Prapiroon on 26 August 2000. Fig-ure 4 shows the evolution of Rossby wave trains asso-

ciated with a preexisting Typhoon Bilis, which formedon 18 August. On 22 August, Bilis was located to thenortheast of the Philippines. One day later, it madelandfall on the southeastern coast of China and rapidlydissipated within 48 h. Nevertheless, its wave train still

FIG. 1. Time sequences of synoptic-scale surface wind patterns associated with the Rossby wave energydispersion of Typhoon Jelawat. The letter “A” represents the center location of Jelawat that formed on1 Aug 2000 and “B” represents the center location of a new TC named Ewiniar that formed on 9 Aug2000 in the wake of the Rossby wave train of Jelawat.

1380 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 5: Tropical Cyclogenesis Associated with Rossby Wave Energy

existed. On 26 August, a new TC (Prapiroon) formed inthe cyclonic vorticity region of the wave train.

The third example is the formation of Typhoon Utor.Figure 5 illustrates the time sequence of surface windpatterns associated with a preexisting Typhoon Durianthat formed over the South China Sea on 29 June 2001.

A clear Rossby wave train was discerned in the wake ofDurian on 30 June. This wave train has a zonal wave-length of 3000 km, tilting toward the east-southeast. On1 July, Durian made landfall on China, near HongKong. In the meantime, a new TC (Utor) formed in thecyclonic vorticity center of the wave train.

FIG. 2. Time sequences of the cloud liquid water field (mm) associated with the Rossby wave energydispersion of Typhoon Jelawat. As in Fig. 1, “A” represents center location of Jelawat and “B” repre-sents center location of Ewiniar.

MAY 2006 L I A N D F U 1381

Fig 2 live 4/C

Page 6: Tropical Cyclogenesis Associated with Rossby Wave Energy

To make sure that the wave trains derived above arenot an artifact of the filtering, we subtracted the origi-nal QuikSCAT wind field from a 31-day sliding-windowaverage (from 15 days prior to 15 days after). The re-sults (figures not shown) show that the wave trains areclearly present even with the raw data.

Thus, the satellite (QuikSCAT and TMI) measure-ments provide the direct observational evidence of theTCED-induced Rossby wave train, its structure andevolution characteristics, and its role in TC genesis.This confirms the previous TCED-induced cyclogenesishypothesis based on either limited observational data(e.g., Frank 1982) or numerical model experiments(e.g., Holland 1995).

To reveal the vertical structure of the TCED-inducedRossby wave train, we examine the NCEP–NCAR re-analysis fields. The same synoptic-scale filtering tech-nique was applied to the reanalysis data. By comparingthe QuikSCAT-derived wave train patterns with thosefrom the reanalysis data, we noted that the reanalysisdata in general underestimate the strength of the wavetrain; in particular, it completely misses the wave trainpattern induced by Typhoon Saomai (2000). However,it captures the wave train pattern of Typhoon Jelawat(2000) reasonably well. Thus, in the following we ex-

amine the vertical structure of the Rossby wave trainfor Jelawat. Figure 6 shows that the northwest–southeast-oriented wave trains are clearly evident fromthe surface to 500 hPa, and become less organized inthe upper troposphere (200 hPa).

Figure 7 illustrates the vertical cross section of theRossby wave train along a northwest–southeast-oriented axis (i.e., the solid black line at the bottompanel of Fig. 6). From Fig. 7, one can see that theTCED-induced Rossby wave train penetrates throughthe entire tropospheric layer. While the divergence inthe wake region exhibits a dominant first baroclinicmode structure, with maximum divergence or conver-gence centers located at 150 hPa or the surface, thevorticity has an equivalent barotropic structure. In thecyclonic vorticity center of the wave train, the maxi-mum vorticity perturbation appears at 850 hPa. Thezonal and meridional wind perturbations are collocatedand have the same sign, implying a northwest–southeast orientation for the wave train.

4. Factors that influence the wave train generation

The examination of all the 34 TC cases reveals thatnot all TCs have a Rossby wave train in their wake. Thisposes an interesting question: What determines thegeneration of the Rossby wave train? Previous theoret-ical and modeling studies suggested that TCED de-pends greatly on the size of a TC (Carr and Elsberry1995) and its wind profile (e.g., Flierl et al. 1983). A TCwith zero total relative angular momentum tends toconserve its energy (Shapiro and Ooyama 1990).

Because of a lack of reliable observational informa-tion on TC size and wind profiles, we examine relation-ships between the occurrence of the Rossby wave trainand TC intensity, based on the fact that many intenseTCs in the WNP are often large ones. For all the 34TCs, we have a total of 233 (snapshot) samples. Weseparate the 233 samples into three groups based on TCminimum sea level pressure (MSLP). They representstrong (MSLP � 960 hPa), moderate (960 hPa � MSLP� 980 hPa), and weak (MSLP 980 hPa) TC intensi-ties. Figure 8 shows the percentages of occurrence ofthe Rossby wave train at each group. For the strongTCs, the percentage is quite high (about 87%). It dropsquickly to 40% (30%) for moderate (weak) storms. Theresult above shows a great sensitivity of the occurrenceof the Rossby wave train to the TC intensity.

In addition, the occurrence of the Rossby wave trainmay also depend on the mean background flow. Figure 9illustrates the wave train generation and nongenerationcases for the strong TC group (i.e., MSLP � 960 hPa).

FIG. 3. Horizontal map of E vectors calculated based on Tren-berth (1986); E � ([�u�u� � ��� �]/2, [�u�� �]), where [ ] representtime averaging, and u� and � � are synoptic wind perturbationsderived from the QuikSCAT data. The E vectors are summedover a 15-day period centered on 5 Aug 2000. The letter “A”marks position of TC Jelawat on 5 Aug 2000.

1382 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 7: Tropical Cyclogenesis Associated with Rossby Wave Energy

Here closed circles (squares) in the figure representthe cases with (without) the appearance of the wavetrain. For strong TCs, the southeast–northwest-oriented Rossby wave trains were observed west of155°E. They are absent to the east of this longitudewhere the mean easterly trades are large (�4 m s�1 orstronger).

Physical mechanisms that give rise to this mean flowdependence is unclear. It is speculated that it may resultfrom the frequency modulation by the mean flow, asthe strong easterly trades may prevent the southeast-ward propagation of the perturbation kinetic energyemitted from a TC.

5. Large-scale dynamic and thermodynamiccontrol on cyclogenesis

Our synoptic analysis of the QuikSCAT wind fieldshows that for 15 TCs that generated a Rossby wavetrain in their wake, only 6 of them eventually lead tonew TC formation. The fact that some of the Rossbywave trains lead to cyclogenesis but others do not raises

an important question: What determines the cyclogen-esis threshold in the Rossby wave train? In the follow-ing, we intend to address this issue from large-scaledynamics and thermodynamic points of view.

We hypothesize that the background flow and mois-ture conditions are critical for the wave train develop-ment. To identify the difference of the background flowbetween the cyclogenesis and noncyclogenesis cases,we conducted a composite analysis. Six genesis casesand nine nongenesis cases are composed. Figure 10shows the comparisons of the composite large-scalebackground variables for the genesis and nongenesiscases. These variables are averaged in a 5° � 5° boxcentered at the cyclonic vorticity center of the wavetrain, and have been subjected to a low-pass filter toretain signals longer than 20 days so that they reason-ably represent the large-scale environmental conditionsincluding the seasonal cycle and intraseasonal oscilla-tion (ISO) activities. Note that there are clear differ-ences between the genesis and nongenesis cases inlarge-scale background conditions. For instance, thevertical motion profiles have a significant difference in

FIG. 4. Time sequences of synoptic-scale surface wind patterns associated with the Rossby wave energy disper-sion of Typhoon Bilis; “A” represents the center location of Bilis that formed on 18 Aug 2000 and “B” representsthe center location of a new TC named Prapiroon that formed on 26 Aug 2000 in the wake of the Rossby wave trainof Bilis.

MAY 2006 L I A N D F U 1383

Page 8: Tropical Cyclogenesis Associated with Rossby Wave Energy

the midtroposphere. A stronger upward motion ap-pears in the genesis cases. Consistent with the verticalmotion, stronger low-level convergence, upper-level di-vergence, and low-level cyclonic vorticity occur in thecyclogenesis cases. Vertical wind shear is also relativelysmaller and the relative humidity is larger throughoutthe troposphere in the cyclogenesis cases. Thus, all en-vironmental variables exhibit a favorable condition inthe cyclogenesis cases.

Figure 11a shows the horizontal patterns of the com-posite large-scale background wind fields at 500 and850 hPa associated with the genesis and nongenesiscases. For the cyclogenesis cases, a background cyclonicwind shear appears at both 850 and 500 hPa. However,for the nongenesis cases, an anticyclonic shear appearsin lower troposphere (850 hPa), while rather uniformwinds cross the center at 500 hPa. These wind featuresare well reflected in the vorticity composites (see Fig.11b), which exhibit a positive vorticity near the com-

posite center at both 850 and 500 hPa in the cyclogen-esis cases, compared to a near-zero or negative vorticityin the noncyclogenesis cases. The synoptic-scale wavetrain center nearly collocates with the background vor-ticity center in the genesis cases.

In addition to the large-scale flow pattern, back-

FIG. 5. Time sequences of synoptic-scale surface wind patternsassociated with the Rossby wave energy dispersion of TyphoonDurian; “A” represents the center location of Durian that formedon 29 Jun 2001 and “B” represents the center location of a newTC named Utor that formed on 1 Jul 2001 in the wake of theRossby wave train of Durian.

FIG. 6. Rossby wave train patterns at 200, 500, 850, and 1000hPa produced by TCED from Typhoon Jelawat on 8 Aug 2000.The wind fields are 3–8-day filtered. The thick line in the bottompanel represents the axis of the wave train.

1384 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 9: Tropical Cyclogenesis Associated with Rossby Wave Energy

ground thermodynamic conditions such as humiditydistribution and moist static stability in lower tropo-sphere may also play a role. It is noted that the specifichumidity field has a more favorable large-scale back-ground in the genesis cases, that is, large humidity cen-ters appear near the cyclonic vorticity centers of thewave trains at both 850 and 500 hPa.

Figure 12 illustrates the thermodynamic control inthe formation of TC Prapiroon. Both Prapiroon and theprevious Typhoon Bilis formed in the same area (within10° longitude distance). Strong winds and dense cloudsassociated with Bilis cooled the ocean surface throughenhanced upwelling/evaporation and reduction ofshortwave radiation. As a result, the local ocean surface

became relatively cooler compared to its surroundings,even a few days after Bilis had moved away from theregion. This led to relatively stable stratification inthe region, which did not favor cyclogenesis eventhough the wave train had developed. Then, the at-mospheric conditional instability started to build upgradually, which can be clearly seen from the increaseof both surface moisture and air temperature at 2 mfrom day �4 to day 0 in Fig. 12 (here day 0 refers to thegenesis date for Typhoon Prapiroon). During thisbuilding-up period, surface air temperature increasedby 1°C, while the surface specific humidity increased by2 g kg�1. The increase of both the temperature andmoisture led to the increase of the surface saturation

FIG. 9. Geographic location of the strong TCs (MSLP � 960hPa) that do (closed circle) and do not (closed square) have aRossby wave train in the wake. Superposed are background June–September-averaged wind vectors at 1000 hPa. The region wherethe mean easterly wind speed is greater than 4 m s�1 is contouredand shaded.

FIG. 7. Vertical cross section of (a) divergence (10�6 s�1), (b)vorticity (10�5 s�1), (c) zonal wind (m s�1), and (d) meridionalwind (m s�1) along the Rossby wave train axis defined in Fig. 6.The y axis denotes pressure (hPa). The open and closed circlesrepresent the relative position of cyclonic and anticyclonic centersof the wave train.

FIG. 8. Percentage of occurrence of TCED-induced Rossbywave train for strong, moderate, and weak TCs.

MAY 2006 L I A N D F U 1385

Page 10: Tropical Cyclogenesis Associated with Rossby Wave Energy

equilibrium potential temperature. As a result, (�se/z)decreased in the lower troposphere (between 1000 and850 hPa), leading to more unstable atmospheric strati-fication.

The slow thermodynamic building-up process mightbe responsible for the observed frequency of TC gen-esis in the WNP. In an observational study, Frank(1982) found that the typical separation distance andtime period between the two subsequent TCs in theWNP were around 2000 km and 7–10 days. For theTyphoon Prapiroon case, it is likely that the timing ofcyclogenesis was determined by both dynamic factorssuch as the strength of the wave train and backgroundflow conditions and the thermodynamic accumulationof the surface moist static energy. Prapiroon formed inthe same region as Bilis did 8 days prior. Strong windsand thick cloud cover associated with Bilis reduced theSST and the surface moist static energy. It required a

“recovery” period to rebuild conditional unstablestratification in the region.

6. Conclusions

The structure and evolution characteristics of Rossbywave trains induced by TC energy dispersion are re-vealed with direct satellite (QuikSCAT and TMI) mea-surements. An E vector calculation shows that the wavetrain is indeed associated with the TC Rossby waveenergy dispersion. For 34 TCs analyzed in the WNPduring 2000–01 typhoon seasons, there are 6 cyclogen-esis cases that are associated with the Rossby wave en-ergy dispersion of a preexisting TC. In all these cases, anew TC formed in the cyclonic vorticity region of apreexisting well-defined wave train. It seems that thereis a scale (particularly meridional scale) contractionprocess associated with cyclogenesis. The cause of this

FIG. 10. Composite vertical structure of 20-day low-pass-filtered divergence, p vertical velocity, vorticity, zonalwind, specific humidity, and relative humidity averaged over the cyclonic vorticity center of the Rossby wave trainsfor six genesis (open circle) and nine nongenesis (closed circle) cases.

1386 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 11: Tropical Cyclogenesis Associated with Rossby Wave Energy

scale contraction is unknown at the moment, and it isspeculated that it may result from the wave train–meanflow interaction and energy accumulation in a conver-gent mean background flow.

Our observational analyses indicate that not all TCshave Rossby wave trains in their wakes. The occurrenceof the wave train seems dependent on TC intensity andthe background mean flow. For stronger TCs, there are

more chances to form a Rossby wave train in theirwakes.

We also note that not all TCs that have a Rossbywave train finally lead to new TC formation. Whetheror not the Rossby wave train results in cyclogenesisdepends on background dynamic and thermodynamicconditions that are related to both the seasonal cycleand ISO activities. The stronger low-level convergence

FIG. 11. (a) Composites of 20-day low-pass-filtered wind fields at 500 and 850 hPa forthe cyclogenesis and noncyclogenesis cases. The vertical (horizontal) axis is relativedistance in latitude (longitude) degrees to the center of cyclonic circulation of the wavetrain. (b) Same as (a) except for vorticity (10�5 s�1) fields.

MAY 2006 L I A N D F U 1387

Page 12: Tropical Cyclogenesis Associated with Rossby Wave Energy

and cyclonic vorticity, weaker vertical shear, andgreater midtropospheric moisture are among the favor-able background conditions for cyclogenesis. The re-building process of a conditional unstable stratificationis important in regulating the frequency of TC genesis.

In Part I we focus on the observational aspects of theRossby wave train. In reality, the TC energy dispersionmight not be a sole factor, and other processes such aseasterly waves (e.g., Li et al. 2003) and upper tropo-spheric forcing may also play a role. Thus it is necessaryto conduct a composite analysis to obtain common fea-tures associated with the Rossby wave train scenario. InPart II, we will simulate the process associated withTCED-induced cyclogenesis in a 3D model to revealfundamental dynamics that give rise to the TC genesis.

Acknowledgments. The authors thank Drs. BinWang, Yuqing Wang, and Melinda Peng for valuable

discussions; anonymous reviewers for constructivecomments; and Dr. Francis Tam for the E vectors cal-culation. This work was supported by NSF GrantATM01-19490 and ONR Grants N000140310739 andN000140210532. International Pacific Research Centeris partially sponsored by the Japan Agency for Marine-Earth Science and Technology (JAMSTEC).

REFERENCES

Anthes, R. A., 1982: Tropical Cyclones: Their Evolution, Structureand Effects. Meteor. Monogr., No. 41, Amer. Meteor. Soc.,208 pp.

Briegel, L. M., and W. M. Frank, 1997: Large-scale influences ontropical cyclogenesis in the western North Pacific. Mon. Wea.Rev., 125, 1397–1413.

Carr, L. E., III, and R. L. Elsberry, 1994: Systematic and inte-grated approach to tropical cyclone track forecasting. Part I.Approach overview and description of meteorological basis.NPS Tech. Rep. NPS MR-94-002, 273 pp.

——, and ——, 1995: Monsoonal interactions leading to suddentropical cyclone track changes. Mon. Wea. Rev., 123, 265–289.

Chan, J. C. L., and R. T. Williams, 1987: Analytical and numericalstudies of the beta-effect in tropical cyclone motion. Part I:Zero mean flow. J. Atmos. Sci., 44, 1257–1265.

Chang, C.-P., J.-M. Chen, P. A. Harr, and L. E. Carr, 1996: North-westward-propagating wave patterns over the tropical west-ern North Pacific during summer. Mon. Wea. Rev., 124, 2245–2266.

Davidson, N. E., and H. H. Hendon, 1989: Downstream develop-ment in the Southern Hemisphere monsoon during FGGE/WMONEX. Mon. Wea. Rev., 117, 1458–1470.

Dickinson, M., and J. Molinari, 2002: Mixed Rossby–gravitywaves and western Pacific tropical cyclogenesis. Part I: Syn-optic evolution. J. Atmos. Sci., 59, 2183–2196.

Ferreira, R. N., and W. H. Schubert, 1997: Barotropic aspects ofITCZ breakdown. J. Atmos. Sci., 54, 261–285.

Fiorino, M., and R. L. Elsberry, 1989: Some aspects of vortexstructure related to tropical cyclone motion. J. Atmos. Sci.,46, 975–990.

Flierl, G. R., 1984: Rossby wave radiation from a strongly nonlin-ear warm eddy. J. Phys. Oceanogr., 14, 47–58.

——, M. E. Stern, and J. A. Whitehead, 1983: The physical sig-nificance of modons: Laboratory experiments and generalintegral constraints. Dyn. Atmos. Oceans, 7, 233–263.

Frank, W. M., 1982: Large-scale characteristics of tropical cy-clones. Mon. Wea. Rev., 110, 572–586.

Gray, W. M., 1968: Global view of the origin of tropical distur-bances and storms. Mon. Wea. Rev., 96, 669–700.

——, 1975: Tropical cyclone genesis. Dept. of Atmospheric Sci-ence Paper 234, Colorado State University, Ft. Collins, CO,121 pp.

——, 1977: Tropical cyclone genesis in the western North Pacific.J. Meteor. Soc. Japan, 55, 465–482.

Holland, G. J., 1995: Scale interaction in the western Pacific mon-soon. Meteor. Atmos. Phys., 56, 57–79.

Kuo, H.-C., J.-H. Chen, R. T. Williams, and C.-P. Chang, 2001:Rossby wave in zonally opposing mean flow: Behavior innorthwest Pacific summer monsoon. J. Atmos. Sci., 58, 1035–1050.

Lau, K.-H., and N.-C. Lau, 1990: Observed structure and propa-

FIG. 12. Evolution of (top) air temperature at 2 m above seasurface, (middle) surface specific humidity, and (bottom) satura-tion potential temperature difference between 1000 and 850 hPa(�se|1000hPa � �se|850hPa) during the formation of TC Prapiroon.The horizontal axis is time (days), and day 0 refers to the genesisdate (26 Aug 2000) for Typhoon Prapiroon.

1388 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 63

Page 13: Tropical Cyclogenesis Associated with Rossby Wave Energy

gation characteristics of tropical summertime synoptic scaledisturbances. Mon. Wea. Rev., 118, 1888–1913.

——, and ——, 1992: The energetics and propagation dynamics oftropical summertime synoptic-scale disturbances. Mon. Wea.Rev., 120, 2523–2539.

Li, T., B. Fu, X. Ge, B. Wang, and M. Peng, 2003: Satellite dataanalysis and numerical simulation of tropical cyclone forma-tion. Geophys. Res. Lett., 30, 2122–2126.

——, X. Ge, B. Wang, and T. Zhu, 2006: Tropical cyclogenesisassociated with Rossby wave energy dispersion of a prexist-ing typhoon. Part II: Numerical simulations. J. Atmos. Sci.,63, 1390–1409.

Luo, Z., 1994: Effect of energy dispersion on the structure andmotion of tropical cyclone. Acta Meteor. Sin., 8, 51–59.

McDonald, N. R., 1998: The decay of cyclonic eddies by Rossbywave radiation. J. Fluid Mech., 361, 237–252.

Molinari, J., D. Vollaro, S. Skubis, and M. Dickinson, 2000: Originand mechanism of eastern Pacific tropical cyclogenesis: Acase study. Mon. Wea. Rev., 128, 125–139.

Murakami, M., 1979: Large-scale aspects of deep convective ac-tivity over the GATE area. Mon. Wea. Rev., 107, 994–1013.

Ritchie, E. A., and G. J. Holland, 1997: Scale interactions during

the formation of Typhoon Irving. Mon. Wea. Rev., 125, 1377–1396.

——, and ——, 1999: Large-scale patterns associated with tropicalcyclogenesis in the western Pacific. Mon. Wea. Rev., 127,2027–2043.

Shapiro, L. J., and K. V. Ooyama, 1990: Vortex evolution on abeta plane. J. Atmos. Sci., 47, 170–187.

Sobel, A. H., and C. S. Bretherton, 1999: Development of synop-tic-scale disturbances over the summertime tropical north-west Pacific. J. Atmos. Sci., 56, 3106–3127.

Trenberth, K. E., 1986: An assessment of the impact of transienteddies on the zonal flow during a blocking episode usinglocalized Eliassen–Palm flux diagnostics. J. Atmos. Sci., 43,2070–2087.

Wang, B., and T. Li, 1994: Convective interaction with boundary-layer dynamics in the development of a tropical intraseasonalsystem. J. Atmos. Sci., 51, 1386–1400.

Zehnder, J. A., D. Powell, and D. Ropp, 1999: The interaction ofeasterly waves, orography, and the ITCZ in the genesis ofeastern Pacific tropical cyclones. Mon. Wea. Rev., 127, 1566–1585.

MAY 2006 L I A N D F U 1389