1. Introduction
The intraseasonal oscillation (ISO) with a 30–90-day time scale is an important mode in the tropical atmosphere, which is dominated by the boreal winter Madden–Julian oscillation (MJO) and boreal summer ISO (BSISO). The ISO has large impacts on a wide variety of climate phenomena across different spatial and temporal scales, for example, the onset of some El Niño events (Moore and Kleeman 1999; Zhang 2005), the Pacific–North American (PNA) pattern (Mori and Watanabe 2008), and the Arctic Oscillation (Zhou and Miller 2005). The northward propagation of the BSISO is closely related to the active and break periods of the Indian summer monsoon (Annamalai and Slingo 2001).
The ISO always appears as a planetary-scale oscillation, and many studies have been carried out to explain this scale selection. For example, the longwave approximation in the planetary boundary layer (PBL) moisture convergence (Wang 1988; Wang and Rui 1990), the role of nonlinear heating in the Rossby–Kelvin wave system coupled by the PBL moisture convergence (Wang and Xue 1992; Li and Zhou 2009), the large surface friction and strong momentum diffusion of the short waves in the troposphere (Kang et al. 2013), the thermodynamics in the moisture mode (Sobel and Maloney 2012, 2013), and the scale interaction between the ISO and synoptic-scale waves (Wang and Liu 2011; Liu and Wang 2012b) all can select the planetary-scale oscillation of ISO.
The ISO also shows prominent seasonal variation (Madden 1986; Wang and Rui 1990; Zhang and Dong 2004; Kikuchi et al. 2012). In the boreal winter, the planetary-scale, eastward-propagating MJO dominates over the tropical Indo-western Pacific (Madden and Julian 1971, 1972). In the boreal summer, the northward-/northeastward-propagating BSISO dominates over the Indian Ocean (Yasunari 1979; Annamalai and Sperber 2005; Wang et al. 2005), and the northward-/northwestward-propagating BSISO dominates over the western North Pacific (Murakami 1984; Kemball-Cook and Wang 2001). The land surface heat fluxes (Webster and Holton 1982), the interaction between convection and moist stability (Gyoswami and Shukla 1984), the air–sea interaction (Kemball-Cook and Wang 2001), the interaction between baroclinic and barotropic vorticity forced by vertical wind shear (VWS) (Wang and Xie 1997; Jiang et al. 2004; Drbohlav and Wang 2005), and the beta shift (Boos and Kuang 2010) mechanisms have been presented to explain this northward propagation of the BSISO. In the ERA-Interim reanalysis and the super parameterized Community Climate System Model (SP-CCSM), the boundary layer moisture advection and the barotropic vorticity effect are found to be the dominant mechanisms for the northward propagation (DeMott et al. 2013).
Convective momentum transport (CMT) is also found to be important for the ISO. Convection not only provides heating to the atmosphere but also redistributes the atmospheric momentum vertically through relatively fast convective mixing processes. The CMT by cumulus convection may be a positive feedback in the MJO, and the kinetic energy is transferred upscale from subgrid systems to the large-scale zonal flow during the westerly onset phase of the MJO (Wu and Yanai 1994; Tung and Yanai 2002a,b). Recently, some general circulation model (GCM) experiments showed that the CMT by cumulus convection is important for the simulation of northward propagation of the BSISO (Kang et al. 2010, hereafter K10), and their results showed that, among the models that Sperber and Annamali once used (Sperber and Annamalai 2008), most models with the CMT are able to simulate the northward propagation, but all of the models without the CMT fail to reproduce the northward-propagating signal. This conclusion, however, needs to be validated by more modeling studying in the future, since many GCMs considering cumulus momentum transport still cannot simulate the northward propagation of the BSISO. In K10, two experiments with and without the CMT were also compared by using the ocean–atmosphere coupled GCM of Seoul National University; the results showed that the experiment with the CMT can simulate the northward propagation of the BSISO while the experiment without the CMT cannot.
To explain why the model with (without) the CMT can (cannot) simulate the northward propagation, K10 presented one mechanism: the lower-level convergence to the north of convection, which is induced by the secondary meridional circulation associated with the baroclinic CMT under the easterly VWS, contributes to this northward propagation. However, the assumption of zonal symmetry in this mechanism cannot represent the role of waves, and the baroclinic CMT-induced northward propagation of the zonally symmetric rainband is too weak; it has a phase speed of 0.2° day−1 only. In observations, the northward-propagation speed of the ISO is about 0.75° day−1 over the Indian Ocean (K10). The work of K10 was based on the 2-layer baroclinic framework and the instability in their model was generated by slightly negative moisture static stability, so the role of barotropic mode could not be studied. The CMT by cumulus convection actually has a strong barotropic component (Schneider and Lindzen 1976). While its barotropic structure has also been found in the observation (Oh et al. 2015) and in the model simulation (Miyakawa et al. 2012), the role of barotropic CMT has not been studied before. Thus, we try to understand the role of barotropic CMT by cumulus convection using a theoretical ISO model in this study.
The remainder of the paper is organized as follows. A theoretical 2.5-layer model of the ISO is presented in section 2. This section also introduces the barotropic CMT processes. In section 3, analytical solutions under different easterly VWSs are presented to illustrate the role of CMT. In section 4 the initial value problem has been calculated to study the role of CMT in the northward propagation of the ISO. Some discussion and concluding remarks are given in section 5.
2. The moist wave dynamical model with barotropic CMT
a. Physical consideration
The essential dynamics of ISO involves coupling both wave dynamics and moisture processes (Wang and Rui 1990; Wang and Xie 1997; Sobel and Maloney 2012, 2013). The theoretical model used in this study is based on the 2.5-layer model (Wang and Rui 1990) and the ISO skeleton model of Majda and Stechmann (Majda and Stechmann 2009), in which the PBL moisture convergence is a moisture source for the free troposphere, and the steady PBL model is used. To represent the moisture processes, we added the tendency of moisture perturbation into the moisture equation. Since the moisture processes are included now, we have to parameterize the precipitation
Cumulus convection tends to mix the winds in the vertical direction quickly (Wu and Yanai 1994). In the NCEP/Department of Energy (DOE) Reanalysis II (Kanamitsu et al. 2002), the climatological mean summer (June–September) zonal wind over India and the Indian Ocean region (5°–25°N) has strong upper-level easterly winds and lower-level westerly winds, which gives a strong easterly VWS (Figs. 1a–c). This easterly VWS also exists over the Indo-western Pacific in the boreal winter (Figs. 1d–f). The convection transports momentum to the upper level that has a strong barotropic component (Kang and Held 1986). In this work we mainly focus on the barotropic mode of CMT. Through the parameterization of vertical transport of momentum by cumulus clouds (Schneider and Lindzen 1976), the linear barotropic CMT is proportional to the upward wind anomaly and the seasonal-mean VWS. Under the easterly VWS, the upward (downward) motion of wet (dry) phase of ISO will mix the westerly (easterly) wind of VWS and produces additional barotropic westerly (easterly) wind forcing. Following the parameterization of Kang and Held (Kang and Held 1986), we parameterize the barotropic CMT by assuming that the CMT is proportional to precipitation and VWS (i.e.,
b. Model formulation
The seasonal-mean flows are found to be important for the ISO (Wang and Xie 1997; Jiang et al. 2004; Drbohlav and Wang 2005). In a perturbation model based on balanced mean flow, the seasonal-mean flows should be included in the momentum equations (Wang and Xie 1997). To focus on the role of the barotropic CMT, the advection term that plays a critical role in the northward propagation of ISO is neglected.
In observations, the ISO shows a dominant baroclinic mode with maximum diabatic heating occurring at the midtroposphere and strong circulation at the upper and lower troposphere (Madden and Julian 1972, 1994; Zhang 2005). Without considering the complicated multicloud vertical structure of the ISO, the ISO can be driven by the midtropospheric diabatic heating, which excites a circulation on the first baroclinic mode (Khouider and Majda 2006, 2007). In this 2-layer troposphere model, we assume that diabatic heating only occurs at the midtroposphere. Since the moisture decreases upward exponentially in the tropical atmosphere and is controlled mainly by lower-tropospheric circulation (Wang 1988), its process can be represented by the first baroclinic mode on the lowest-order assumption (Majda and Stechmann 2009). The frictional moisture convergence in the PBL also pumps moisture and moistens the lower troposphere; thus, the Ekman pumping term can be added to the moisture equation (Wang and Rui 1990; Liu and Wang 2012a).
c. Mathematical methods
These linear equations (1)–(4) can be solved as an eigenvalue problem, or they can be integrated from an initial disturbance. After projecting them onto the zonal wavenumber–frequency space, we can obtain the eigenvalues and eigenvectors by matrix inversion. The frequency and growth rate in the eigenvalue problem are defined by the real and imaginary parts of eigenvalues, respectively. Detail of this calculation can be found in Liu and Wang (2012a).
This projection on the zonal wavenumber–frequency space cannot represent the meridional-propagating waves, so we have to solve the initial value problem. To integrate this model, an initial wavenumber-1 Kelvin wave–like perturbation is used. The finite-difference method is adopted in both time and space. The time integration scheme is centrally differenced with a 2.5-min time step and a time-average coefficient of 0.125. For instance, on step
For the initial value problem, the initial disturbance is set to have the equatorially trapped structure, and the zonal structure is wavenumber 1. The model reaches a steady state from day 30 onward. By this time, initial transient features have decayed, leaving nearly periodic disturbances that are close to steady in their respective commoving reference frames. Thus, we begin our discussion using model results from day 31 of the integration. The phase speed of northward propagation in the initial problem can be calculated by the slope of precipitation at the date line in the Northern Hemisphere.
3. Role of CMT under different VWSs: Analytical solution
Figure 2 shows the eigenvalues for different VWS structures. Without the VWS and its associated CMT, this model gives a peculiar dispersion relation and selects the longest eastward-propagating mode as the most unstable mode. This is inconsistent with the results of Liu and Wang where only the baroclinic mode and the PBL process are coupled (Liu and Wang 2012a). When the easterly VWS is included, the CMT accelerates the eastward propagation of eastward-propagating modes and reduces the instability, especially for the short waves. The CMT under the equatorially trapped easterly VWS seems to act more efficiently to accelerate the eastward propagation than the subtropical easterly VWS does. These results are reversed under the westerly VWS (not shown). This finding is based on the theoretical model, and the mechanism of CMT reducing the frictional convergence instability should be further studied by GCM experiments in the future.
Without the VWS, the strong positive equatorial temperature anomalies are sandwiched by negative subtropical temperature anomalies in front of positive convective center, and the strong upward equatorial Ekman pumping is excited there (Fig. 3a). Thus, the PBL will pump additional moisture into the free troposphere to sustain the growth of eastward-propagating modes. In this model the PBL moisture convergence is maximum at the equator and decays poleward, which favors the growth of the Kelvin waves rather than the Rossby waves, thus the Rossby component and its cyclonic winds are relatively weak.
Without the PBL, the baroclinic eddy momentum transport from synoptic-scale motions can provide an instability source for the MJO (Liu and Wang 2013). Here, the barotropic CMT by cumulus convection will reduce the instability from the PBL moisture convergence, and this negative role of the barotropic CMT is stronger for shorter waves (Fig. 2b), which can be explained by analyzing the horizontal structures. Because of the coupling of equatorial Rossby waves and Kelvin waves (Wang and Rui 1990), the equatorial upward Ekman pumping is sandwiched by subtropical downward Ekman pumping in front of the positive convective center (Fig. 3a). The inclusion of CMT under the equatorially trapped easterly VWS will accelerate the zonal wind and enhance the subtropical gyre through the geostrophic balance, which enhances the subtropical downward Ekman pumping and suppresses the equatorial upward Ekman pumping (Fig. 3b). The subtropical gyre affected by the CMT is stronger for shorter waves. This is because the temperature-anomaly-induced Ekman pumping is much stronger in the subtropics than that at the equator for short waves (Liu and Wang 2012a). Thus, in front of the convective center, the equatorial upward Ekman pumping is greatly reduced for short waves (Fig. 4), and the PBL-pumped moisture for the growth of short waves is reduced by the inclusion of CMT under easterly VWS.
It is interesting that the CMT under the subtropical VWS excites asymmetric structure and the precipitation is enhanced in the Northern Hemisphere (Fig. 3c). Under the easterly VWS of Asian summer monsoon, the CMT is much stronger in the Northern Hemisphere than in the Southern Hemisphere. The simulated temperature anomalies in the Northern Hemisphere are strong, as well as the Ekman-pumping-induced precipitation. The strong positive Ekman pumping to the northwest of positive precipitation region contributes to this asymmetric structure. Since this zonal wavenumber–frequency projection cannot represent the northward propagation of perturbation, it is necessary to integrate this model from an initial perturbation and see how this perturbation evolves.
4. Role of CMT under different VWSs: Initial value problem
Figure 5a shows the model evolution from an initial wavenumber-1 Kelvin wave–like disturbance maximum at the equator, which has a horizontal structure of
As noted by many previous works, the barotropic vorticity effect is the dominant mechanism for the northward propagation of the BSISO (Jiang et al. 2004; DeMott et al. 2013). The CMT under the easterly VWS of Asian monsoon also tends to accelerate the barotropic vorticity to the north of the BSISO convective center, which favors the northward propagation of the BSISO through exciting the upward Ekman pumping.
In Fig. 5a, the strong precipitation centered in the monsoon region will also initiate precipitation with a negative phase near the equatorial region through local self-initiated mechanism (Jiang and Li 2005; Wang et al. 2005; Liu and Wang 2012c). Thus, the oscillation period of the BSISO is determined by the phase speed of northward propagation. In this simple theoretical model, the northward propagation speed is parameter dependent. In Table 1, we show that strong easterly VWS will increase the CMT and accelerate the northward propagation as well. The northward propagation caused by the CMT is steady with respect to the PBL friction
Simulated phase speed of northward propagation as a function of easterly VWS magnitude
The northward propagation also depends on the location of the easterly VWS (Fig. 5). When the easterly VWS is symmetric about the equator (Fig. 5c), it excites symmetric positive Ekman pumping to both sides of the equatorial precipitation, and the model still simulates the Gill-like pattern (Gill 1980), although the subtropical signal is enhanced. When the mean easterly VWS moves to the Northern Hemisphere and y0 = 5°N (Fig. 5b), this symmetry is destroyed and a southeast–northwest-tilted rainband is simulated. Because the CMT and excited Ekman pumping in the Northern Hemisphere are stronger than those in the Southern Hemisphere, the enhanced subtropical Rossby waves emanating from the eastward-propagation forms this tilted rainband (Wang and Xie 1997; Lawrence and Webster 2002; Hsu et al. 2004). The northward propagation of the rain belt is evident in both cases (Figs. 5b,c), and the subtropical VWS also accelerates the eastward propagation by decreasing the equatorial upward Ekman pumping and growth rate. The eastward-propagation speeds of simulated modes under the equatorial and subtropical VWS are 9.7 and 12.0 m s−1, respectively, and modes with subtropical VWS have fast phase speed. This result is different from that of Fig. 2, in which no meridional propagation exists and modes with equatorial VWS have fast phase speed. Associated with the strong northward propagation, the eastward propagation is very fast (Fig. 5a). When the VWS is maximum at 5°N, the weak northward propagation also accelerates the eastward propagation (Fig. 5b). When the easterly VWS moves farther north of y0 = 10°N (Fig. 5a), this model is dominated by the northward propagation of perturbation. Because the CMT and its Ekman pumping reach their maxima at 10°N, they produce strong precipitation to the north of the equatorial precipitation and the dominant northward propagation appears.
In boreal summer, the maximum SST is located at 20°N over the northern Indian Ocean. This asymmetric SST prefers enhancing the Rossby wave component of the ISO (Kang et al. 2013; Liu et al. 2015), because the time-mean intertropical convergence zone (ITCZ) is set by the SST and is located away from the equator (Lindzen and Nigam 1987; Sobel 2007). To compare the roles of the asymmetric SST and CMT in the BSISO, sensitivity experiments with maximum SST in the subtropics or at the equator are carried out (Fig. 6), and the SST structures and amplitudes are the same as other experiments. When the maximum SST moves northward, the Rossby wave component is enhanced (Figs. 6a,b), which is inconsistent with the theoretical results (Kang et al. 2013; Liu et al. 2015). Over this asymmetric SST, the northward propagation is simulated. The SST-induced northward propagation, however, is mainly caused by the Rossby wave emanation mechanism (Wang and Xie 1997; Lawrence and Webster 2002). This northward-propagation speed, with values below 0.2° day−1, is determined by the eastward propagation of the simulated signal and is slow compared to the observed northward propagation of about 0.75° day−1. When the subtropical CMT is also included, a significant northward propagation with a speed of 0.65° day−1 is simulated (Fig. 6c). These experiments mean that in this simple model, the barotropic CMT is more sufficient in inducing the northward propagation of ISO than the asymmetric SST.
5. Conclusions
Recent multimodel analysis of the MJO revealed that the coupling of PBL moisture convergence and free-tropospheric convection may be important for MJO simulation, which means that the instability from the PBL is important for understanding the ISO (Jiang et al. 2015). Under the easterly VWS, the CMT will suppress the instability induced by PBL moisture convergence and accelerate the eastward propagation, especially for short waves; this is the new mechanism we use to explain why the MJO prefers the planetary scale. We view examination of this scale-selection mechanism in GCMs as a target for future study.
This work also presents that the CMT by cumulus convection can induce the northward propagation of the BSISO through exciting the barotropic vorticity to the north of the BSISO. Figure 7 summarizes this mechanism for the generation of upward Ekman pumping to the north of the BSISO due to barotropic CMT. Under the easterly VWS of Asian summer monsoon, the CMT caused by positive convection of the BSISO tends to accelerate the barotropic westerly wind. The negative convection or downward motion to the north of this BSISO, however, is going to accelerate the barotropic easterly wind. Thus, a positive barotropic vorticity tendency is induced to the north of the BSISO convective center, which should excite upward Ekman pumping and prepare enough moisture for the northward propagation of the BSISO. The mechanism behind possible CMT impact on the northward propagation described in this study is different from that in K10. K10 stressed the importance of the baroclinic secondary circulation induced by cumulus friction, which results in lower-tropospheric convergence to the north of convection; their simulated northward propagation, however, is too weak compared to the observation. This study emphasizes the importance of the barotropic CMT for the northward propagation of the BSISO.
In this paper, we only focused on the role of the CMT and neglected another important mechanism for the northward propagation of the ISO, the vertical shear mechanism proposed by Jiang et al. (Jiang et al. 2004) and Drbohlav and Wang (Drbohlav and Wang 2005). They argued that the generation of barotropic vorticity due to coupling between atmospheric baroclinic and barotropic modes in the presence of vertical shear in the mean flow causes moisture convergence in the PBL, which leads to the northward shift of convection. This barotropic vorticity can also be effectively generated by the upward transport of westerly momentum by cumulus convection over the Indian monsoon region during the boreal summer. In future, these two mechanisms should be compared using the same model by adding both the advection and CMT terms.
In recent global 7-km cloud-resolving model simulation (Miyakawa et al. 2012) and observation analysis based on the NOAA/Climate Forecast System Reanalysis (CFSR) (Oh et al. 2015), the CMT showed a 3-layer structure: positive momentum tendency anomalies near the surface, negative (positive) in the lower to midtroposphere, and strong positive (negative) in the upper troposphere were found within and to the west (to the east) of the MJO convection. This means CMT not only has the barotropic mode but also has a baroclinic mode. Over the convective center, the baroclinic CMT should accelerate the lower-tropospheric easterly wind and produce anticyclonic wind anomalies to the north of the convective center, which may be a negative feedback in the northward propagation of BSISO. The observed positive PBL CMT and positive barotropic CMT in convective center supports our strong barotropic mode assumption. In this paper, we focused on demonstrating the role of the barotropic CMT. A more realistic 3-layer vertical structure, instead of this simple 2-layer model, should be tested in the future.
These results imply that accurate simulation of the mean state in the GCM is important to capture the realistic CMT by cumulus convection, which is important for ISO simulation. In a recent study (Hung et al. 2013), the improved tropical intraseasonal variability in two Coupled Model Intercomparison Project phase 5 (CMIP5) models with CMT included (i.e., CCSM4 and CNRM-CM5) as compared to their CMIP3 versions suggests the potential importance of CMT (Yukimoto et al. 2012; Zhou et al. 2012). The role of CMT in the ISO should be further studied in more CMIP5 models.
Acknowledgments
We are grateful for helpful and encouraging comments from three anonymous reviewers and Chongyin Li. This work was supported by China National 973 Project 2015CB453200, the NOAA/MAPP project (under Award NA10OAR4310247), the Asian-Pacific Economic Cooperation (APEC) Climate Center, and the National Research Foundation (NRF) of Korea through a Global Research Laboratory (GRL) Grant (MEST, 2011-0021927). In-Sik Kang was supported by the National Research Foundation of Korea Grant (MEST) (NRF-2012M1A2A2671775) and the BK21 program.
REFERENCES
Annamalai, H., and J. Slingo, 2001: Active/break cycles: Diagnosis of the intraseasonal variability of the Asian Summer Monsoon. Climate Dyn., 18, 85–102, doi:10.1007/s003820100161.
Annamalai, H., and K. Sperber, 2005: Regional heat sources and the active and break phases of boreal summer intraseasonal (30–50 day) variability. J. Atmos. Sci., 62, 2726–2748, doi:10.1175/JAS3504.1.
Boos, W. R., and Z. Kuang, 2010: Mechanisms of poleward propagating, intraseasonal convective anomalies in cloud system-resolving models. J. Atmos. Sci., 67, 3673–3691, doi:10.1175/2010JAS3515.1.
DeMott, C. A., C. Stan, and D. A. Randall, 2013: Northward propagation mechanisms of the boreal summer intraseasonal oscillation in the ERA-Interim and SP-CCSM. J. Climate, 26, 1973–1992, doi:10.1175/JCLI-D-12-00191.1.
Drbohlav, H.-K. L., and B. Wang, 2005: Mechanism of the northward-propagating intraseasonal oscillation: Insights from a zonally symmetric model. J. Climate, 18, 952–972, doi:10.1175/JCLI3306.1.
Frierson, D. M., A. J. Majda, and O. M. Pauluis, 2004: Large scale dynamics of precipitation fronts in the tropical atmosphere: A novel relaxation limit. Commun. Math. Sci., 2, 591–626, doi:10.4310/CMS.2004.v2.n4.a3.
Gill, A. E., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106, 447–462, doi:10.1002/qj.49710644905.
Gyoswami, B., and J. Shukla, 1984: Quasi-periodic oscillations in a symmetric general circulation model. J. Atmos. Sci., 41, 20–37, doi:10.1175/1520-0469(1984)041<0020:QPOIAS>2.0.CO;2.
Hsu, H.-H., C.-H. Weng, and C.-H. Wu, 2004: Contrasting characteristics between the northward and eastward propagation of the intraseasonal oscillation during the boreal summer. J. Climate, 17, 727–743, doi:10.1175/1520-0442(2004)017<0727:CCBTNA>2.0.CO;2.
Hsu, P.-c., and T. Li, 2012: Role of the boundary layer moisture asymmetry in causing the eastward propagation of the Madden–Julian oscillation. J. Climate, 25, 4914–4931, doi:10.1175/JCLI-D-11-00310.1.
Hung, M.-P., J.-L. Lin, W. Wang, D. Kim, T. Shinoda, and S. J. Weaver, 2013: MJO and convectively coupled equatorial waves simulated by CMIP5 climate models. J. Climate, 26, 6185–6214, doi:10.1175/JCLI-D-12-00541.1.
Jiang, X., and T. Li, 2005: Reinitiation of the boreal summer intraseasonal oscillation in the tropical Indian Ocean. J. Climate, 18, 3777–3795, doi:10.1175/JCLI3516.1.
Jiang, X., T. Li, and B. Wang, 2004: Structures and mechanisms of the northward propagating boreal summer intraseasonal oscillation. J. Climate, 17, 1022–1039, doi:10.1175/1520-0442(2004)017<1022:SAMOTN>2.0.CO;2.
Jiang, X., and Coauthors, 2015: Vertical structure and diabatic processes of the Madden–Julian Oscillation: Exploring key model physics in cimate simulations. J. Geophys. Res., doi:10.1002/2014JD022375, in press.
Kanamitsu, M., W. Ebisuzaki, J. Woollen, S.-K. Yang, J. Hnilo, M. Fiorino, and G. Potter, 2002: NCEP–DOE AMIP-II Reanalysis (R-2). Bull. Amer. Meteor. Soc., 83, 1631–1643, doi:10.1175/BAMS-83-11-1631.
Kang, I.-S., and I. M. Held, 1986: Linear and nonlinear diagnostic models of stationary eddies in the upper troposphere during northern summer. J. Atmos. Sci., 43, 3045–3057, doi:10.1175/1520-0469(1986)043<3045:LANDMO>2.0.CO;2.
Kang, I.-S., D. Kim, and J.-S. Kug, 2010: Mechanism for northward propagation of boreal summer intraseasonal oscillation: Convective momentum transport. Geophys. Res. Lett., 37, L24804, doi:10.1029/2010GL045072.
Kang, I.-S., F. Liu, M.-S. Ahn, Y.-M. Yang, and B. Wang, 2013: The role of SST structure in convectively coupled Kelvin–Rossby waves and its implications for MJO formation. J. Climate, 26, 5915–5930, doi:10.1175/JCLI-D-12-00303.1.
Kemball-Cook, S., and B. Wang, 2001: Equatorial waves and air–sea interaction in the boreal summer intraseasonal oscillation. J. Climate, 14, 2923–2942, doi:10.1175/1520-0442(2001)014<2923:EWAASI>2.0.CO;2.
Khouider, B., and A. J. Majda, 2006: A simple multicloud parameterization for convectively coupled tropical waves. Part I: Linear analysis. J. Atmos. Sci., 63, 1308–1323, doi:10.1175/JAS3677.1.
Khouider, B., and A. J. Majda, 2007: A simple multicloud parameterization for convectively coupled tropical waves. Part II: Nonlinear simulations. J. Atmos. Sci., 64, 381–400, doi:10.1175/JAS3833.1.
Kikuchi, K., and Y. N. Takayabu, 2004: The development of organized convection associated with the MJO during TOGA COARE IOP: Trimodal characteristics. Geophys. Res. Lett., 31, L10101, doi:10.1029/2004GL019601.
Kikuchi, K., B. Wang, and Y. Kajikawa, 2012: Bimodal representation of the tropical intraseasonal oscillation. Climate Dyn., 38, 1989–2000, doi:10.1007/s00382-011-1159-1.
Kiladis, G. N., K. H. Straub, and P. T. Haertel, 2005: Zonal and vertical structure of the Madden–Julian Oscillation. J. Atmos. Sci., 62, 2790–2809, doi:10.1175/JAS3520.1.
Lawrence, D. M., and P. J. Webster, 2002: The boreal summer intraseasonal oscillation: Relationship between northward and eastward movement of convection. J. Atmos. Sci., 59, 1593–1606, doi:10.1175/1520-0469(2002)059<1593:TBSIOR>2.0.CO;2.
Li, T., and C. Zhou, 2009: Planetary scale selection of the Madden–Julian oscillation. J. Atmos. Sci., 66, 2429–2443, doi:10.1175/2009JAS2968.1.
Lindzen, R. S., and S. Nigam, 1987: On the role of sea surface temperature gradients in forcing low-level winds and convergence in the tropics. J. Atmos. Sci., 44, 2418–2436, doi:10.1175/1520-0469(1987)044<2418:OTROSS>2.0.CO;2.
Liu, F., and B. Wang, 2012a: A frictional skeleton model for the Madden–Julian oscillation. J. Atmos. Sci., 69, 2749–2758, doi:10.1175/JAS-D-12-020.1.
Liu, F., and B. Wang, 2012b: A model for the interaction between 2-day waves and moist Kelvin waves. J. Atmos. Sci., 69, 611–625, doi:10.1175/JAS-D-11-0116.1.
Liu, F., and B. Wang, 2012c: A conceptual model for self-sustained active-break Indian summer monsoon. Geophys. Res. Lett., 39, L20814, doi:10.1029/2012GL053663.
Liu, F., and B. Wang, 2013: Impacts of upscale heat and momentum transfer by moist Kelvin waves on the Madden–Julian oscillation: A theoretical model study. Climate Dyn., 40, 213–224, doi:10.1007/s00382-011-1281-0.
Liu, F., G. Huang, and Y. Mi, 2015: Role of SST meridional structure in coupling the Kelvin and Rossby waves of the intraseasonal oscillation. Theor. Appl. Climatol., doi:10.1007/s00704-014-1266-0, in press.
Madden, R. A., 1986: Seasonal variations of the 40–50 day oscillation in the tropics. J. Atmos. Sci., 43, 3138–3158, doi:10.1175/1520-0469(1986)043<3138:SVOTDO>2.0.CO;2.
Madden, R. A., and P. R. Julian, 1971: Detection of a 40–50 day oscillation in the zonal wind in the tropical Pacific. J. Atmos. Sci., 28, 702–708, doi:10.1175/1520-0469(1971)028<0702:DOADOI>2.0.CO;2.
Madden, R. A., and P. R. Julian, 1972: Description of global-scale circulation cells in the tropics with a 40–50 day period. J. Atmos. Sci., 29, 1109–1123, doi:10.1175/1520-0469(1972)029<1109:DOGSCC>2.0.CO;2.
Madden, R. A., and P. R. Julian, 1994: Observations of the 40–50-day tropical oscillation—A review. Mon. Wea. Rev., 122, 814–837, doi:10.1175/1520-0493(1994)122<0814:OOTDTO>2.0.CO;2.
Majda, A. J., and J. A. Biello, 2004: A multiscale model for tropical intraseasonal oscillations. Proc. Natl. Acad. Sci. USA, 101, 4736–4741, doi:10.1073/pnas.0401034101.
Majda, A. J., and S. N. Stechmann, 2009: The skeleton of tropical intraseasonal oscillations. Proc. Natl. Acad. Sci. USA, 106, 8417–8422, doi:10.1073/pnas.0903367106.
Miyakawa, T., Y. N. Takayabu, T. Nasuno, H. Miura, M. Satoh, and M. W. Moncrieff, 2012: Convective momentum transport by rainbands within a Madden–Julian oscillation in a global nonhydrostatic model with explicit deep convective processes. Part I: Methodology and general results. J. Atmos. Sci., 69, 1317–1338, doi:10.1175/JAS-D-11-024.1.
Moore, A. M., and R. Kleeman, 1999: Stochastic forcing of ENSO by the intraseasonal oscillation. J. Climate, 12, 1199–1220, doi:10.1175/1520-0442(1999)012<1199:SFOEBT>2.0.CO;2.
Mori, M., and M. Watanabe, 2008: The growth and triggering mechanisms of the PNA: A MJO-PNA coherence. J. Meteor. Soc. Japan, 86, 213–236, doi:10.2151/jmsj.86.213.
Murakami, M., 1984: Analysis of the deep convective activity over the Western Pacific and Southeast Asia. II: Seasonal and intraseasonal variations during Northern Summer. J. Meteor. Soc. Japan, 62, 88–108.
Myers, D. S., and D. E. Waliser, 2003: Three-dimensional water vapor and cloud variations associated with the Madden–Julian Oscillation during Northern Hemisphere winter. J. Climate, 16, 929–950, doi:10.1175/1520-0442(2003)016<0929:TDWVAC>2.0.CO;2.
Oh, J.-H., X. Jiang, D. Waliser, M. W. Moncrieff, and R. H. Johnson, 2015: Convective momentum transport associated with the Madden–Julian oscillation based on a reanalysis dataset. J. Climate, in press.
Salby, M. L., and H. H. Hendon, 1994: Intraseasonal behavior of clouds, temperature, and motion in the Tropics. J. Atmos. Sci., 51, 2207–2224, doi:10.1175/1520-0469(1994)051<2207:IBOCTA>2.0.CO;2.
Schneider, E. K., and R. S. Lindzen, 1976: A discussion of the parameterization of momentum exchange by cumulus convection. J. Geophys. Res., 81, 3158–3160, doi:10.1029/JC081i018p03158.
Sobel, A. H., 2007: Simple models of ensemble-averaged precipitation and surface wind, given the sea surface temperature. The Global Circulation of the Atmosphere, T. Schneider and A. H. Sobel, Eds., Princeton University Press, 219–251.
Sobel, A. H., and E. Maloney, 2012: An idealized semi-empirical framework for modeling the Madden–Julian oscillation. J. Atmos. Sci., 69, 1691–1705, doi:10.1175/JAS-D-11-0118.1.
Sobel, A. H., and E. Maloney, 2013: Moisture modes and the eastward propagation of the MJO. J. Atmos. Sci., 70, 187–192, doi:10.1175/JAS-D-12-0189.1.
Sperber, K. R., and H. Annamalai, 2008: Coupled model simulations of boreal summer intraseasonal (30–50 day) variability. Part 1: Systematic errors and caution on use of metrics. Climate Dyn., 31, 345–372, doi:10.1007/s00382-008-0367-9.
Tian, B., D. E. Waliser, E. J. Fetzer, B. H. Lambrigtsen, Y. L. Yung, and B. Wang, 2006: Vertical moist thermodynamic structure and spatial temporal evolution of the MJO in AIRS observations. J. Atmos. Sci., 63, 2462–2485, doi:10.1175/JAS3782.1.
Tung, W.-W., and M. Yanai, 2002a: Convective momentum transport observed during the TOGA COARE IOP. Part I: General features. J. Atmos. Sci., 59, 1857–1871, doi:10.1175/1520-0469(2002)059<1857:CMTODT>2.0.CO;2.
Tung, W.-W., and M. Yanai, 2002b: Convective momentum transport observed during the TOGA COARE IOP. Part II: Case studies. J. Atmos. Sci., 59, 2535–2549, doi:10.1175/1520-0469(2002)059<2535:CMTODT>2.0.CO;2.
Wang, B., 1988: Dynamics of tropical low-frequency waves: An analysis of the moist Kelvin wave. J. Atmos. Sci., 45, 2051–2065, doi:10.1175/1520-0469(1988)045<2051:DOTLFW>2.0.CO;2.
Wang, B., and H. Rui, 1990: Dynamics of the coupled moist Kelvin–Rossby wave on an equatorial β-plane. J. Atmos. Sci., 47, 397–413, doi:10.1175/1520-0469(1990)047<0397:DOTCMK>2.0.CO;2.
Wang, B., and Y. Xue, 1992: Behavior of a moist Kelvin wave packet with nonlinear heating. J. Atmos. Sci., 49, 549–559, doi:10.1175/1520-0469(1992)049<0549:BOAMKW>2.0.CO;2.
Wang, B., and T. Li, 1993: A simple tropical atmosphere model of relevance to short-term climate variations. J. Atmos. Sci., 50, 260–284, doi:10.1175/1520-0469(1993)050<0260:ASTAMO>2.0.CO;2.
Wang, B., and T. Li, 1994: Convective interaction with boundary-layer dynamics in the development of a tropical intraseasonal system. J. Atmos. Sci., 51, 1386–1400, doi:10.1175/1520-0469(1994)051<1386:CIWBLD>2.0.CO;2.
Wang, B., and X. Xie, 1997: A model for the boreal summer intraseasonal oscillation. J. Atmos. Sci., 54, 72–86, doi:10.1175/1520-0469(1997)054<0072:AMFTBS>2.0.CO;2.
Wang, B., and F. Liu, 2011: A model for scale interaction in the Madden–Julian oscillation. J. Atmos. Sci., 68, 2524–2536, doi:10.1175/2011JAS3660.1.
Wang, B., P. J. Webster, and H. Teng, 2005: Antecedents and self-induction of active-break south Asian monsoon unraveled by satellites. Geophys. Res. Lett., 32, 4704, doi:10.1029/2004GL020996.
Webster, P. J., and J. R. Holton, 1982: Cross-equatorial response to middle-latitude forcing in a zonally varying basic state. J. Atmos. Sci., 39, 722–733, doi:10.1175/1520-0469(1982)039<0722:CERTML>2.0.CO;2.
Wu, X., and M. Yanai, 1994: Effects of vertical wind shear on the cumulus transport of momentum: Observations and parameterization. J. Atmos. Sci., 51, 1640–1660, doi:10.1175/1520-0469(1994)051<1640:EOVWSO>2.0.CO;2.
Yano, J.-I., and K. Emanuel, 1991: An improved model of the equatorial troposphere and its coupling with the stratosphere. J. Atmos. Sci., 48, 377–389, doi:10.1175/1520-0469(1991)048<0377:AIMOTE>2.0.CO;2.
Yasunari, T., 1979: Cloudiness fluctuations associated with the Northern Hemisphere summer monsoon. J. Meteor. Soc. Japan, 57, 227–242.
Yukimoto, S., and Coauthors, 2012: A New Global Climate Model of the Meteorological Research Institute: MRI-CGCM3—Model description and basic performance (special issue on recent development on climate models and future climate projections). J. Meteor. Soc. Japan, 90A, 23–64, doi:10.2151/jmsj.2012-A02.
Zhang, C., 2005: Madden–Julian Oscillation. Rev. Geophys., 43, RG2003, doi:10.1029/2004RG000158.
Zhang, C., and M. Dong, 2004: Seasonality in the Madden–Julian Oscillation. J. Climate, 17, 3169–3180, doi:10.1175/1520-0442(2004)017<3169:SITMO>2.0.CO;2.
Zhou, L., and I.-S. Kang, 2013: Influence of Convective Momentum Transport on Mixed Rossby–Gravity Waves: A Contribution to Tropical 2-Day Waves. J. Atmos. Sci., 70, 2467–2475, doi:10.1175/JAS-D-12-0300.1.
Zhou, L., R. B. Neale, M. Jochum, and R. Murtugudde, 2012: Improved Madden–Julian oscillations with improved physics: The impact of modified convection parameterizations. J. Climate, 25, 1116–1136, doi:10.1175/2011JCLI4059.1.
Zhou, S., and A. J. Miller, 2005: The interaction of the Madden–Julian Oscillation and the Arctic Oscillation. J. Climate, 18, 143–159, doi:10.1175/JCLI3251.1.