Interhemispheric Propagation of Stationary Rossby Waves in a Horizontally Nonuniform Background Flow

Yanjie Li State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing, China

Search for other papers by Yanjie Li in
Current site
Google Scholar
PubMed
Close
,
Jianping Li College of Global Change and Earth System Science, Beijing Normal University, and Joint Center for Global Change Studies, Beijing, China

Search for other papers by Jianping Li in
Current site
Google Scholar
PubMed
Close
,
Fei Fei Jin School of Ocean and Earth Science and Technology, University of Hawai‘i at Mānoa, Honolulu, Hawaii

Search for other papers by Fei Fei Jin in
Current site
Google Scholar
PubMed
Close
, and
Sen Zhao State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, and University of Chinese Academy of Sciences, Beijing, China

Search for other papers by Sen Zhao in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Significant interhemispheric teleconnections are identified that span the tropical easterlies in the boreal summer 300-hPa streamfunction, such as the North Africa–Antarctic (NAA) and the North Pacific–South America (NPSA) patterns. These patterns are not supported by traditional wave theory, since stationary waves in a basic state without meridional wind would be trapped in the easterlies. To describe the interhemispheric responses more realistically, two-dimensional spherical Rossby wave theory in a horizontally nonuniform basic state is considered. Conditions sufficient for the existence of one propagating wave are obtained, and the meridional group velocity of the wave is shown to have the same direction as the meridional basic wind at the traditional critical latitude. It is concluded that stationary waves with a specific wavelength can propagate across the easterlies from south (north) to north (south) via southerly (northerly) flows. Hence, energy transport by stationary waves on a horizontally nonuniform basic state may produce interhemispheric responses that could pass through the tropical easterly belt.

The wave theory and a barotropic model are then applied to idealized and climatological flows. Model results agree well with the theory. In boreal winter and summer, cross-equatorial flows steer stationary waves propagating from one hemisphere to the other across the tropical easterlies, especially over the Australian–Asian monsoon region. It seems that the large-scale monsoonal background flows play a critical role in the interhemispheric teleconnection. Additionally, the wave ray trajectory and model results suggest that the NAA pattern may result from Rossby wave energy dispersion.

Denotes Open Access content.

Corresponding author address: Prof. Jianping Li, College of Global Change and Earth System Science, Beijing Normal University, Beijing 100875, China. E-mail: ljp@bnu.edu.cn

Abstract

Significant interhemispheric teleconnections are identified that span the tropical easterlies in the boreal summer 300-hPa streamfunction, such as the North Africa–Antarctic (NAA) and the North Pacific–South America (NPSA) patterns. These patterns are not supported by traditional wave theory, since stationary waves in a basic state without meridional wind would be trapped in the easterlies. To describe the interhemispheric responses more realistically, two-dimensional spherical Rossby wave theory in a horizontally nonuniform basic state is considered. Conditions sufficient for the existence of one propagating wave are obtained, and the meridional group velocity of the wave is shown to have the same direction as the meridional basic wind at the traditional critical latitude. It is concluded that stationary waves with a specific wavelength can propagate across the easterlies from south (north) to north (south) via southerly (northerly) flows. Hence, energy transport by stationary waves on a horizontally nonuniform basic state may produce interhemispheric responses that could pass through the tropical easterly belt.

The wave theory and a barotropic model are then applied to idealized and climatological flows. Model results agree well with the theory. In boreal winter and summer, cross-equatorial flows steer stationary waves propagating from one hemisphere to the other across the tropical easterlies, especially over the Australian–Asian monsoon region. It seems that the large-scale monsoonal background flows play a critical role in the interhemispheric teleconnection. Additionally, the wave ray trajectory and model results suggest that the NAA pattern may result from Rossby wave energy dispersion.

Denotes Open Access content.

Corresponding author address: Prof. Jianping Li, College of Global Change and Earth System Science, Beijing Normal University, Beijing 100875, China. E-mail: ljp@bnu.edu.cn

1. Introduction

Teleconnection is an important and common phenomenon in large-scale atmospheric circulation motion. Atmospheric teleconnection patterns have prominent influences on global climate variability (Lau and Nath 1996; Alexander et al. 2002; Liu and Alexander 2007; Wu et al. 2009). Since the 1980s, many studies have employed various statistical methods to determine the dominant modes of atmospheric variability. Wallace and Gutzler (1981) identified five preferred patterns in the Northern Hemisphere (NH) winter 500-hPa height. Blackmon et al. (1984a,b) examined different time scale fluctuations. The long time scale patterns are similar to results in Wallace and Gutzler’s (1981), while the intermediate-scale 10–30-day fluctuations involve zonal wave trains along the westerly jet. The structure and energetics of these patterns were further investigated by Nakamura et al. (1987). Variations of 700-hPa height have also attracted attention (e.g., Chen 1982; Esbensen 1984; Van den Dool and Livezy 1984; Barnston and Livezey 1987). The Pacific–North America (PNA) pattern and zonally oriented waves along the westerly jet are prominent features. In addition, many researchers worked on circulation variability in the Southern Hemisphere (SH) (e.g., Mo and White 1985; Karoly 1989, 1990; Berbery et al. 1992; Mo and Higgins 1998; Kidson 1999). The Antarctic Oscillation (Rogers and van Loon 1982; Gong and Wang 1999), the zonal wavenumber-3–5 pattern at middle latitudes (Mo and White 1985; Karoly 1990), and the Pacific–South America pattern (PSA; Karoly 1989; Mo and Higgins 1998) were identified.

Boreal winter patterns are generally ENSO-related (Trenberth et al. 1998). Several teleconnection patterns in the NH summer associated with the summer monsoon have also been found, including the Pacific–Japan (PJ) and East Asian–Pacific (Nitta 1987, 1989; Huang and Lu 1989; Huang and Sun 1992) and Indo-Asian–Pacific (J. Li et al. 2013) patterns in convective activity, the “silk road” pattern in the 200-hPa meridional velocity (Lu et al. 2002; Enomoto et al. 2003), and the circumglobal teleconnection (CGT) pattern (Ding and Wang 2005; Ding et al. 2011). Teleconnection patterns are thought of as stationary waves emanating from a source region. These patterns, in general, are either zonally or meridionally oriented. The zonally oriented patterns, such as the Eurasian, the silk road, and the CGT patterns, develop along the westerly jet, which acts as a waveguide (Hoskins and Ambrizzi 1993, hereafter HA93), while the meridionally oriented patterns (e.g., PNA, PJ, and PSA) result from the arcing path of the meridional propagation of two-dimensional spherical Rossby waves (Hoskins and Karoly 1981). Hence, the study of Rossby wave propagation is vital to understand the variations of atmospheric circulation.

Rossby wave theory is based on the concepts of dispersion and group velocity (Rossby 1945; Yeh 1949). Disturbances are usually defined theoretically in terms of the limiting case of a linear perturbation on a basic state. Therefore, the features of the waves obtained depend on the fixed basic state. Rossby (1945) and Yeh (1949) set the basic state to be uniform and used the ß-plane approximation to study the one-dimensional motion. Longuet-Higgins (1964, 1965), Hoskins et al. (1977), and Hoskins and Karoly (1981) extended the theory to two-dimensional wave propagation on a sphere in a latitudinally varying medium. The “great circle” ray path is suggested as the preferred trajectory of stationary wave energy dispersion from the tropics to the extratropics. However, this meridional propagation is limited to westerly flow. Stationary wave energy is absorbed at the critical latitudes where the wavelength becomes very large and amplitude tends to zero as a result of the zonal wind approaching zero.

These results are based on zonally symmetric flow, which is far from the real situation. Modeling studies by Simmons (1982), Webster and Holton (1982), and Branstator (1983) emphasize the longitudinal variation of the zonal basic flow and suggest the existence of a “westerly duct” for cross-equatorial waves. Atmospheric circulation in the NH and SH interacts through the Hadley cell or the cross-equatorial flows. Cross-equatorial teleconnection was found over the tropical east Pacific and Atlantic during the NH winter in the upper troposphere (Hsu and Lin 1992; Tomas and Webster 1994). This is termed as Rossby wave propagation from the NH to SH via the westerly window (HA93). However, some wave train patterns associated with boreal summer monsoon have been identified across tropical easterly belts (e.g., Nitta 1987; Krishnamurti et al. 1997; Lin 2009; Y. Li et al. 2013). He (1990) stated the significant interactions between the atmospheric circulations in the NH and SH in boreal summer related to the large-scale monsoon systems. Recent modeling works by Lee et al. (2013) and Liu and Wang (2013) suggest that the NH summer monsoon can influence the SH circulation anomalies through barotropic responses, in which the meridional background wind plays an important role. So far, there are two different ways of illustrating the cross-easterly responses. One is the meridional propagation of stationary Rossby waves directly from the forcing to the other hemisphere, obtained by adding a mean meridional flow in the nondivergent vorticity equation and shallow water equations (Schneider and Watterson 1984; Farrell and Watterson 1985; Watterson and Schneider 1987; Esler et al. 2000; Kraucunas and Hartmann 2007). The other is the advection of the vorticity by perturbation divergent flow, which can transmit the tropical heating in the easterly to subtropical westerly jet and then make Rossby waves effectively. This way is proved by Sardeshmukh and Hoskins (1988) using a linearized barotropic vorticity equation with divergence and decomposing the wind into rotational and divergent components. Even though the two illustrations are still far from the real atmosphere (Held et al. 2002), they are landmarks in the effort made by meteorology researchers to understand the complicated reality.

Results in the barotropic model (Schneider and Watterson 1984; Esler et al. 2000; Kraucunas and Hartmann 2007), baroclinic model (Lee et al. 2013), and even the intermediate model (Liu and Wang 2013; Ji et al. 2014) have suggested the importance of the former way in the interhemispheric teleconnection. However, the fundamental equation in the theory is linearized on a zonal mean state, neglecting the longitudinal variation, which is remarkable in the meridional wind particularly. Unlike the zonal wind, the meridional wind features a notable regional structure. These longitude-dependent structures must have different effects on the stationary wave propagation. Hence, we present the interhemispheric stationary wave propagation theory based on a latitudinally and longitudinally varying flow (herein referred to as a horizontally nonuniform flow) in this study as an extension of previous studies (Karoly 1983; Schneider and Watterson 1984).

The dispersion relation and ray equations for Rossby waves based on a horizontally nonuniform flow were derived by Karoly (1983). Although this more realistic flow complicates the wave theory and may undermine the validity of the Wentzel–Kramers–Brillouin (WKB) approximation, more realistic atmospheric responses could be understood qualitatively. Shaman and Tziperman (2005, 2007) used two-dimensional ray trajectories through low-pass filtered climatological wind fields to illustrate the ENSO-teleconnection influences on the South Asian monsoon. Additionally, the propagating behavior of nonzero frequency waves differs from stationary waves (Karoly 1983; Yang and Hoskins 1996; Li and Nathan 1994, 1997) and is not discussed here.

Li (1991) analyzed the global characteristics of the 30–60-day oscillation and found cross-equatorial correlation centers over the middle Pacific at 500 hPa. Ambrizzi et al. (1995) examined the global teleconnection patterns of the 10–30-day variability in the 200-hPa austral winter streamfunction for the period 1979–89. However, the cross-equatorial pattern in their work is unclear. The different results may indicate that the cross-equatorial teleconnection patterns depend on time scale. We attempt to further examine the interhemispheric teleconnections in the boreal summer mean streamfunction field.

The paper is organized as follows. In section 2, the data, methodology, and the model used in this work are introduced. Teleconnection patterns, especially those across the equator, in the June–August (JJA)-mean streamfunction field are analyzed in section 3 using more recent and longer time series than that in Ambrizzi et al. (1995). In section 4, the fundamentals of two-dimensional spherical Rossby wave theory over horizontally nonuniform flow are reviewed, and the possibility of interhemispheric propagation of stationary waves is discussed theoretically. Then, both the wave theory and barotropic model are applied to investigate the interhemispheric stationary wave propagation in idealized and realistic flows. Results from wave theory are shown in section 5 and from the barotropic model in section 6. Conclusions and dynamical implications are discussed in section 7.

2. Data, methodology, and model

a. Data and methodology

The 40-yr European Centre for Medium-Range Weather Forecasts (ECMWF) Re-Analysis (ERA-40) data from 1958 to 2001 (Uppala et al. 2005) are used in this study. The data are obtained from a global spectral model with T159 truncation over a 320 × 160 global grid. The JJA-mean streamfunction is used in the teleconnectivity analysis, and the climatological JJA and December–February (DJF) flows are employed to study the interhemispheric propagation of stationary waves.

For the detection of global teleconnections, streamfunction is a better choice than geopotential height in terms of its suitability to describe the low-latitude circulation (e.g., Hsu and Lin 1992). Teleconnectivity analysis (Wallace and Gutzler 1981) is applied to the 300-hPa streamfunction. This technique presents the geographical distribution of coherent structures of multiple one-point correlation patterns in a single map. The base points are set to have the same resolution as the data grids. Teleconnectivity is defined as
eq1
where Rij is the correlation between the base point i and every grid point j.

The WKB approximation used in Rossby wave theory requires the variation of the perturbation to be much smaller than that of the basic state. For this reason, a two-dimensional Fourier expansion truncated at zonal and meridional wavenumber 5 is used to smooth the climatological flows. The smoothed flows preserve most of the large-scale signals and approximate the desired slowly varying media in which the Rossby waves propagate. Details of this method are given in Sun and Li (2012).

b. Barotropic model

Both barotropic and baroclinic models have been employed in previous studies (e.g., Hoskins and Karoly 1981; Simmons 1982; HA93; Ambrizzi et al. 1995; Jin and Hoskins 1995; Ambrizzi and Hoskins 1997; Wang et al. 2005). The baroclinic model results are qualitatively similar to those of the barotropic model (HA93; Ambrizzi and Hoskins 1997). Therefore, we utilize the linearized barotropic model with forcing, damping, and fourth-order diffusion to explore the interhemispheric responses:
e1
where u and υ are the zonal and meridional wind velocities; ψ is the streamfunction; f is the Coriolis force; λ and φ are longitude and latitude; and a is Earth’s radius. The overbar indicates basic state; the prime indicates perturbation. The τ and μ are damping and diffusion coefficients given the value of 1.16 × 10−6 s−1 and 2.34 × 1016 m4 s−1. The variables and J are spherical Lagrange and Jacobi operators. The forcing S′ takes the same form as in Wang et al. (2005):
e2
where S0 denotes the strength of the forcing, (x0, y0) is the center location, and xd and yd are scales in latitude and longitude. The model has a resolution of T42, and the steady state is derived by a 25-day integration.

3. Interhemispheric teleconnections in the JJA 300-hPa streamfunction

Figure 1 shows the teleconnectivity distribution for the 300-hPa streamfunction and eddy streamfunction (with zonal mean removed) fields. The significant cross-equatorial correlations in the tropics are the most prominent feature in the streamfunction field (Fig. 1a). This indicates that the circulations in the two hemispheric tropics are closely related. These patterns are mainly dipoles and are zonally symmetric on the one-point correlation maps of mostly base points (not shown), as seen in previous studies (Hsu and Lin 1992; Lau et al. 1994; Ambrizzi et al. 1995). They seem to be the symmetric responses to tropical heating. However, the correlation pattern over the Pacific sector has a wave train–like structure. This pattern is more clearly defined in the eddy streamfunction field but for two additional centers at high latitudes in the SH. The one-point correlation maps of the base points at 22.991°S, 204.75°E (Fig. 2c) and 12.897°N, 200.25°E (Fig. 2d) show the positive and negative centers distributed at intervals over the Pacific and Atlantic. The structure is PSA-like, but previous PSA studies have not mentioned the north tropical Pacific center. We refer to it as the North Pacific–South America (NPSA) pattern here to distinguish it from the traditional PSA. Figures 2a and 2b show another wave train pattern in the eddy streamfunction field with significant correlation centers over North Africa, South Africa, the south Indian Ocean, the Southern Ocean, and the Antarctic. This pattern, called the North Africa–Antarctic (NAA) pattern, has a wavenumber-2 structure in the SH. It indicates that climate variability over Africa and Antarctica may be related, a hypothesis that will require specific statistical and dynamical studies. In addition, the cross-equatorial correlations over the tropical Indian Ocean and equatorial South America are also distinct. Thus, it can be inferred that interhemispheric teleconnections are clearly evident in JJA in the upper troposphere.

Fig. 1.
Fig. 1.

Teleconnectivity maps for the JJA fields of (a) streamfunction and (b) eddy streamfunction (with zonal mean removed) at 300 hPa. Negative signs have been omitted and correlation coefficients multiplied by 100. The gray shading indicates regions where the value is >55. The easterly regions are filled in with red dots. Arrows connect centers of strongest teleconnectivity with the grid points. The calculation is based on the JJA-mean data from ECMWF for the period 1958–2001.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

Fig. 2.
Fig. 2.

One-point correlation maps of eddy streamfunction for the base points (a) 25.234°S, 15.75°E, (b) 29.719°N, 16.875°E, (c) 22.991°S, 204.75°E, and (d) 12.897°N, 200.25°E. The gray shading indicates areas significant at the 95% confidence level. The easterly regions are filled in with red dots. The red symbols “+” and “Δ” denote the location of the base point and the minimum negative correlation, respectively.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

In DJF, the interhemispheric teleconnection is found over the tropical east Pacific (Hsu and Lin 1992), where the westerly window is embedded in the tropical easterly belt (Fig. 3b). The spatial distribution of the stationary wavenumber (Fig. 3d; defined as in HA93) suggests that the westerly wind over this region acts as a stationary wave duct for interhemispheric propagation. However, the climatological JJA zonal wind is a strong circumglobal easterly (Fig. 3a). The stationary wavenumber approaches infinity at critical latitudes where and is imaginary when . These limitations suggest the impossibility of interhemispheric stationary wave propagation in the basic state that excludes the meridional wind, whereas observations show that there are indeed intensive interhemispheric responses across the critical latitudes. Kraucunas and Hartmann (2007) implied that the cross-equator mean meridional winds can promote wave activity propagation and contribute to the symmetric responses to hemispheric asymmetric forcing (Hartmann 2007). In the next section, we examine theoretically the possibility of interhemispheric propagation of stationary waves in a horizontally nonuniform basic flow.

Fig. 3.
Fig. 3.

(a),(b) The climatological 300-hPa wind vectors (m s−1) and the meridional gradient of absolute vorticity Qy (color shading, 10−11 m−1 s−1); (c),(d) the total wavenumber of the stationary waves defined as (2.13) in HA93. (left) JJA and (right) DJF. The climatological flows are smoothed using a two-dimensional Fourier series truncated at zonal and meridional wavenumber 5.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

4. Interhemispheric propagation theory in a horizontally nonuniform flow

a. Fundamentals

The barotropic nondivergent vorticity equation on a sphere is usually employed to describe the remote atmospheric response to a local forcing:
e3
and the Mercator projection (e.g., Hoskins and Karoly 1981) can simplify the equation and retain the spherical effect of Earth:
e4
Linearized about the horizontally nonuniform basic state , the projected equation for the perturbation streamfunction takes the form
e5
where , is the basic-state absolute vorticity, and . As the coefficients of the terms in (5) vary in both the x and y directions, it is difficult to derive the exact solution. Assuming a slowly varying medium, however, coefficients can be viewed as constants locally. Therefore, an elementary solution has the form
e6
where A(X, Y, T) and θ(x, y, t) are the wave amplitude and phase, respectively; (X, Y, T) = ε(x, y, t); and ε is a small parameter, indicating a length scale of perturbations far less than that of the basic flow. Substituting (6) into (5), a series of approximate equations at different orders of ε can be derived from the partial differential equation. Usually, the zeroth-order approximation is the frequency–wavenumber relation or dispersion relation describing the propagation characteristics of perturbations:
e7
where are the zonal wavenumber, meridional wavenumber, and the angular frequency, respectively.
The total wavenumber is defined as
e8
The zonal and meridional components of group velocity take the form
e9
e10
The longitudinal and latitudinal variation of the basic state means that both l and k change along the ray path of energy propagation: that is, along a trajectory locally tangential to the group velocity vector (Lighthill 1978). Their evolution is determined by kinematic wave theory (Whitham 1960), as follows:
e11
e12
where denotes the Lagrangian variation moving at the group velocity. Given the basic state, the forcing location, and the initial zonal wavenumber, the initial meridional wavenumber can be obtained through (7) or (8), and then (9)(12) can be integrated to derive the ray trajectory. These equations were first derived by Karoly (1983), but only limited discussion of the propagation behavior in the horizontally nonuniform flow was given. Li and Nathan (1997) discussed these equations in vector form but focused on the nonstationary waves.

b. The possibility of interhemispheric propagation

The stationary wavenumber can be written as
e13
where tanα is actually the slope of the wavenumber vector K = (k, l). If and are nonzero, (13) is equal to
e14
where and are the slopes of the basic-state absolute vorticity and streamfunction contours, respectively. If tanγ and tanβ are negligible, then
e15
where KSS denotes the stationary wavenumber in the basic state that excludes and , as in HA93. Generally, approximation (15) is satisfied for quasigeostrophic flow, since its absolute vorticity is quasi-parallel with the streamfunction. For small wavenumber slope (), (15) is also satisfied. However, the approximation breaks down
  1. over regions where the ageostrophic flow cannot be neglected, such as the tropics. Here, β and γ are different, indicating inhomogeneity of the basic state;

  2. for steeper waves that have a larger scale in the x direction than y. Here, tanα is large because of the structure of the waves themselves.

It can be seen from (14) that the wave steepness can magnify the effect of the inhomogeneity of the basic state. Therefore, KS and KSS may differ substantially, especially when both (i) and (ii) hold. To show the difference more clearly, (14) can be written as a cubic equation for tanα as
e16
Assuming the three solutions to (16) are p1, p2, and p3, then
e17
If is satisfied, there must be at least one positive solution, which indicates at least one propagating wave. The different propagation behaviors of stationary waves in basic states that exclude or include and (denoted as “without-υ” and “with-υ” flows, respectively) are discussed in the following cases:
  1. When , for stationary waves satisfying , gives , while for those satisfying , means . This indicates that in westerly flow, the propagating stationary waves in the without-υ case can still propagate southerly, while those that cannot propagate in the without-υ case can propagate northerly.

  2. When , is always satisfied, and gives . This indicates that in easterly flow, stationary waves are trapped in the without-υ case, while in the with-υ case, the southerly allows at least one propagating stationary wave.

  3. When , and are satisfied. In the without-υ case, both l and υg are zero, and this latitude is the turning latitude. In the with-υ case, there is also at least one zero-meridional-wavenumber solution, with group velocity given by
    e18
    e19
    It can be seen that ug has the same form in flows with and without υ, whereas the form of υg in the two flows is distinct. In the with-υ case, υg depends on the basic state, especially .
  4. When , , starting from (13), is satisfied. This indicates that the inclusion of allows at least one propagating wave in southerly (northerly) wind with a positive (negative) meridional absolute vorticity gradient across the traditional critical latitudes in the without-υ case.

It is thus clear that the meridional component of the basic flow makes the stationary wave propagating behavior very different from that in a purely zonal flow. Schneider and Watterson (1984) considered the meridional wind but ignored the zonal variation in the basic state. Compared with their results, the dispersion relation in (7) explicitly includes the role of the zonal variation in the basic state, which changes the necessary condition for three distinct propagating solutions into
e20
This condition depends on both the zonal wavelength and the basic state, which is different from that in Schneider and Watterson (1984). The condition sufficient for only one real solution can be written as
e21
With only one real solution to (16), the other two solutions must be a complex conjugate pair, taken to be p2 and p3 with a positive product C. Then we can obtain the real solution in the form of
e22
The numerator of the right-hand side in (22) tends to be positive in easterly; thus,
e23
This suggests that in the easterly region for , the lines of constant phase are oriented SE–NW (SW–NE), and for they are oriented SW–NE (SE–NW). The meridional wavelength is proportional to the magnitude of the meridional wind. A stronger meridional wind favors a larger wavelength. At the critical latitudes where , the real solution is
e24
The positive dependence of p1 on the meridional basic wind is also satisfied.
Considering the energy dispersion of this propagating wave, we can write its meridional group velocity as
e25
The weighting coefficients in (25) indicate the relative contributions of the four basic-state components: , , , and . At the critical latitudes, the first term is zero. The analysis in the last paragraph suggests that the remainder has the same sign as . Therefore, meridional energy propagation of stationary waves across critical latitudes tends to be in the same direction as the meridional wind. Even when the zonal wind is nonzero, a sufficiently strong meridional wind may dominate the meridional propagation of stationary waves.

5. Sensitivity of interhemispheric propagation to the meridional wind in idealized flows

To address how the meridional basic flow affects the interhemispheric propagation of stationary waves, an idealized flow dependent on latitude only, as used by Schneider and Watterson (1984), is applied:
e26
e27
where , E and F are parameters governing the distribution and strength of the zonal wind, y0 is the latitude where the southerly wind switches to northerly, and VS and VN satisfy
eq2

Five experiments are designed to investigate the dependence of the meridional wave propagation on the meridional basic wind. The basic state in experiment Exp0 includes only zonal flow (see Fig. 4a), the without-υ case. Exp1–Exp4 are the with-υ cases, with different meridional wind structures in the tropics (briefly introduced in Table 1 with different y0 and V0; see Figs. 4b,c). Given E = 18 m s−1 and F = 14 m s−1, the zonal wind is symmetric about the equator: westerly at middle to high latitudes with a maximum of 26 m s−1 and easterly in the tropics with a maximum about 4 m s−1. The meridional wind in Exp1 (Exp2) is northerly (southerly) in the tropics, while that in Exp3 (Exp4) is antisymmetric about the equator, southerly (northerly) in the NH, and northerly (southerly) in the SH. They all have the same maximum value of 2 m s−1. The Exp1 (Exp2) meridional wind profile crudely represents the climatological cross-equatorial flows at the upper troposphere in JJA (DJF) and lower troposphere in DJF (JJA). Exp4 crudely describes the trade wind in the intertropical convergence zone. Exp3 denotes the situation observed at DJF 300 hPa over the tropical Indian Ocean and Indonesian islands. The role of the meridional background wind in the cross-equatorial stationary wave propagation is isolated by comparing the results in these experiments.

Fig. 4.
Fig. 4.

(a) Zonal wind profile and (b) meridional wind profiles in Exp1 (dashed line) and Exp2 (solid line), and (c) Exp3 (dashed line) and Exp4 (solid line).

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

Table 1.

Sensitivity experiments with different meridional wind distributions.

Table 1.

a. WKB analysis

In the basic flows defined above, the equations in section 4 can be simplified to
e28
e29
The zonally invariant basic flow means that the zonal wavenumber k remains constant along a wave ray. For the without-υ case (Exp0) in particular, the equations are the same as in Hoskins and Karoly (1981). Schneider and Watterson (1984) showed a ray trajectory from NH to SH to identify the interhemispheric propagation behavior of stationary waves across tropical easterlies in a background flow similar to that used here. In this section, we examine the propagation behavior of different waves and the evolution of wavenumber along the ray paths and reemphasize the importance of the meridional ambient flow for stationary wave propagation through the critical latitudes.

The meridional wavenumbers for zonal wavenumber 1, 3, and 6 in Exp0, Exp1, Exp2, and Exp3 are shown in Fig. 5. In Exp0, there are two real solutions of l with opposite signs in the westerly flow, indicating two branches of wave energy dispersion, one equatorward and the other poleward. However, there is no real solution in the easterly flow, indicating that wave energy dissipates quickly. Hence, the easterly region is known as the dead zone for stationary wave propagation. The zero-zonal-wind latitudes at about 10° are critical latitudes, where the meridional wavenumbers tend to infinity. The magnitude of the meridional wavenumber decreases with latitude. For the waves with zonal wavenumbers 1, 3, and 6, the meridional wavenumber becomes zero at 80°, 60°, and 30°, respectively. This means that long waves propagate farther than short waves and play an important role in tropic–extratropic interactions. These results are similar to those obtained in previous studies (e.g., Hoskins and Karoly 1981) but differ over the region of meridional flow in Exp1, Exp2, and Exp3. In Exp1, the positive branch of the meridional wavenumber tends to infinity at different latitudes for different waves: about 30° for k = 1, 20° for k = 3, and 18° for k = 6. The singularity in stationary wavenumber occurs at the latitude where , as implied by (28). Therefore, the critical latitude in the with-υ flow depends on the basic zonal and meridional wind, as well as the wavenumber. At the tropics, there is only one negative meridional wavenumber indicating a SW–NE-tilted wave in Exp1, while there is only one positive meridional wavenumber indicating a SE–NW-tilted wave in Exp2. These single real solutions of the meridional wavenumber have the same sign as the meridional wind, as expected from the theory in section 4b. The magnitude of the single real solution for meridional wavenumber increases with the zonal wavenumber and at the equator is about 5 for k = 1, 8–10 for k = 3, and 10–15 for k = 6. These analyses suggest that stationary waves with certain wavelengths can propagate across the easterlies in the presence of the meridional background flow. In Exp3, the wavenumber of the single solution becomes so large that it is out of the plotted range at the equator (Figs. 5g–i), indicating that stationary waves cannot propagate across the easterlies in the absence of the meridional background flow.

Fig. 5.
Fig. 5.

The meridional wavenumber profiles for the without- and with-υ cases. (a)–(c) Exp1 (red) and Exp0 (gray) for zonal wavenumber k = 1, 3, and 6, respectively. (d)–(f),(g)–(i) As in (a)–(c), but for Exp2 and Exp0 and for Exp3 and Exp0, respectively.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

The field of meridional wavenumber shows the propagable and nonpropagable regions for stationary waves. Ray tracing results are presented in order to further understand how the wave energy disperses. Figure 6 shows the results from Exp0. Consistent with the meridional wavenumber field, one poleward and one equatorward ray are excited from wave sources at 30°N and 30°S, for each value of zonal wavenumbers 1–5. Both branches go eastward. The poleward branch of the wave with zonal wavenumber 1 reaches 80° after 8 days and then turns equatorward, while the other waves turn equatorward at lower latitudes. As they propagate, their wavelength first increases and then decreases after the waves reach the turning latitudes, with wavenumber 4–6 (6000–10 000 km) at 20°–45°, 0–3 (>10 000 km) at higher latitudes, and 7–9 or even larger (<5000 km) at lower latitudes until the equatorward propagation terminates at 10° as the zonal wind approaches zero. Results of Exp1–Exp4 are shown in Fig. 7. The notable differences in meridional propagation occur at the tropics, where the meridional winds are different. In Exp1, the tropics are dominated by easterly and northerly flow. Both the equatorward branch of waves triggered at 30°N and the reflected waves from the poleward branch propagate across the easterly flow, and continue propagating out of the tropical SH, then along the great circle with longer wavelength again. The equatorward waves triggered at 30°S and the reflected waves from the poleward branch, however, are reduced by the north wind near 15°S. The meridional flow in Exp2 is antisymmetric to that in Exp1. As a result, the propagation behaviors are antisymmetric to those in Exp1. Cross-equatorial propagation from the SH to NH is allowed, while that from the NH to SH is impeded by the southerly flow. In Exp3, the equatorward propagation in both the NH and SH disappears at about 15°, where the northerly flow in the NH and the southerly flow in the SH are strong. In Exp4, the equatorward waves propagate symmetrically to the equator and then turn westward. These results suggest that the basic meridional wind is important for the meridional propagation of stationary waves. The direction of meridional propagation is largely dependent on that of the meridional wind. Results of Exp1 and Exp2 indicate that cross-equatorial stationary wave propagation is made possible but limited to one direction by the cross-equatorial flow. In addition, the stationary waves tend to have a shorter wavelength when they propagate across the easterlies.

Fig. 6.
Fig. 6.

Ray paths (curves) and the evolution of total wavenumber (color) along the paths for k = 1–5 waves in the without-υ basic flow (Exp0). The red dots denote the wave sources at 30°N and 30°S. The propagation speed is indicated by the 2-day interval marked by black solid dots for k = 1, triangles for k = 2, square boxes for k = 3, crosses for k = 4, and circles for k = 5. The paths are terminated when the total wavenumber exceeds 40.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

Fig. 7.
Fig. 7.

As in Fig. 6, but for the with-υ flows: (a) Exp1, (b) Exp2, (c) Exp3, and (d) Exp4.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

b. Barotropic model results

The results of the WKB analysis are examined further using the barotropic model linearized on the idealized flow in Exp0–Exp4. The parameters controlling the source strength and width are set as (x0, y0) = (180°, 30°), xd = 12°, yd = 7°, and S0 = 1 × 10−10 s−2 as in Wang et al. (2005). Figure 8 shows the steady response of the streamfunction to the sources placed in the NH and SH. In Exp0, a negative streamfunction response occurs over the wave source. Positive and negative centers are distributed at intervals northeast of the local response, tracing out the great circle trajectory of energy dispersion. Cross-equatorial responses are not possible in this experiment with easterly flow in the tropics in the absence of meridional wind. In Exp1 where the north wind dominates, there are clear wave train responses in the SH to the NH source but almost no responses in the NH to the SH source. In contrast, in Exp2 where the south wind dominates, there are strong responses in the NH to the SH source but no response in the SH to the NH source. These results suggest that a meridional wind allows interhemispheric response in the same direction across the tropical easterlies. In addition, the spatial structures of the cross-equatorial responses in Exp1 (Exp2) are tilted SW–NE (SE–NW) and are of narrower spatial extent than those at middle to high latitudes. These features are consistent with the theoretical analysis and wave ray results. For Exp3 and Exp4, there are almost no responses to the sources in the other hemisphere. The streamfunction responses in Exp3 terminate at about 15°. These features are also similar to the results of the wave ray trajectories.

Fig. 8.
Fig. 8.

Steady eddy streamfunction responses (thin contours, 106 m2 s−1) to the wave sources (thick lines, 10−11 s−2) for (top)–(bottom) Exp0–Exp4 with sources in the (a),(c),(e),(g),(i) NH and (b),(d),(f),(h),(j) SH. Contours are from −3 to 3 with an interval 0.6.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

The above analysis presents the sensitivity of the interhemispheric stationary wave propagation to the distribution of the basic meridional wind. Wang et al. (2005) examined the northern extratropical responses to tropical forcing through a southerly conveyor. Their results show that a stronger southerly wind contributes to stronger extratropical responses. Similarly, in order to further illustrate the effect of the basic meridional wind on the interhemispheric responses, we take Exp1 as the control experiment and perform another two sensitivity experiments in which the basic meridional wind has the same distribution but different strength. The parameters V0 in the experiments are set to be 1 and 4 m s−1, corresponding to the cases with weaker and stronger meridional wind, respectively. Figure 9 shows the steady streamfunction responses to the vorticity source located in the NH in these two experiments and their differences with respect to the result in Exp1. The SH responses in Exp1 are definitely weaker than those in the case with stronger meridional wind but stronger than those in the case with weaker meridional wind. This result indicates that the strength of the meridional wind affects the interhemispheric responses by modulating the energy dispersion velocity.

Fig. 9.
Fig. 9.

Dependence of the streamfunction responses to the NH source on the strength of the meridional wind: (a),(b) as in Fig. 8c, but for weaker and stronger meridional basic winds with V0 = 1 and 4 m s−1, respectively; (c) difference between the steady responses in the Exp1 flow with V0 = 1 and 2 m s−1; (d) as in (c), but for V0 = 4 and 2 m s−1. Contours in (a),(b) are from −3 to 3, with an interval 0.6, and in (c),(d), they are from −1 to 1 with an interval 0.2.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

6. Interhemispheric propagation of stationary waves in realistic flows

a. Wave ray results

In section 5, the sensitivity of the interhemispheric propagation of stationary waves to the basic meridional wind is examined, and the interhemispheric propagation behavior of stationary waves is shown to be strongly dependent on the meridional background wind. In this section, we apply the ray tracing method and the barotropic model to climatological flows. Cross-equatorial flow is observed in both the lower and upper troposphere. The climatological cross-equatorial flow in DJF is from the SH to NH in the upper troposphere and from the NH to SH in the lower troposphere. It is opposite in JJA. These seasonally and vertically opposite flows over the tropics would provide us good examples to compare the stationary wave propagation behaviors under them. Thus, we set the smoothed winter and summer climatological flows at 300 and 850 hPa as basic states to examine the possibility of stationary Rossby waves propagating through the tropical easterlies from one hemisphere to the other under these horizontally nonuniform flows.

Usually, the barotropic model is applied to one appropriate level. The upper troposphere is often used in many previous studies for the larger stationary wave component (e.g., Simmons 1982; HA93). Ting (1996) suggested that the barotropic model is able to describe the real atmosphere at the equivalent barotropic level and identified the equivalent barotropic level around 350 hPa in winter and 500 hPa in summer by comparing the rotational response of the barotropic linearized model with divergence at different vertical levels and that of the baroclinic model. However, the basic state in her work is zonal mean. The results would be changed when the horizontally nonuniform flow is taken into consideration. Therefore, the best choice of the vertical level for the barotropic model is still uncertain. The discussion about the extratropical responses to tropical heating by Held et al. (2002) indicates that the tentative work on the application of the barotropic model to the low level may be worthy. Hence, we involve the wave ray equations and the barotropic model to both the upper and lower troposphere.

A series of wave sources in the tropical NH and SH separately and the wave ray trajectories for each source are calculated. Since the wavenumber varies along the trajectory, the initial zonal wavenumber should be given before the calculation. We have compared the trajectories for the initial zonal wavenumber-1–6 waves. They have different behaviors over some regions, which seem to depend on the configuration of the basic state. For the interhemispheric propagation, they are generally similar. Here, we take zonal wavenumber 4 as an example, and derive the initial meridional wavenumber by solving the dispersion relation. Where there are three real solutions, the one with the smallest absolute value is chosen. Only waves with meridional wavenumber less than 40 (corresponding to wavelengths longer than 1000 km) are thought to propagate.

Figure 10 shows the stationary wave ray paths and the associated total wavenumber based on the smoothed DJF 300- and 850-hPa climatological flows for sources in the tropical NH and SH. Under the 300-hPa flow, rays originating from the tropical NH mostly follow the great circle route: that is, poleward from the tropics and then reflecting equatorward. The wavenumber along these rays is about 4–9, which means they are synoptic scale (5000–10 000 km). At the entrance region of the African–Asian subtropical jet, the rays originating over North Africa and South Asia with a wavenumber of about 7–9 propagate eastward and northward and reflect equatorward at 45°N, where the wavenumber is larger. This is consistent with the waveguide effect of the westerly jet (HA93). At the exit region of the jet, the wave rays take diverse paths, some poleward reaching to 80°N and some equatorward across the equator to SH middle to high latitudes. The cross-equatorial rays appear over the east tropical Pacific, where the so-called westerly duct is embedded in the tropical easterlies. This is in agreement with previous work (Hsu and Lin 1992; HA93). The ray paths originating from the tropical SH to the middle to high latitudes exhibit a more remarkable great circle feature. The difference with respect to the rays originating in the NH is that the cross-equatorial propagation occurs not only in the westerly duct, but also in the easterly region over the Maritime Continent from the northeast of Australia to East Asia. Compared with the situation under 300-hPa flow, the propagation of steady waves based on 850-hPa flow is slower, and the scales are shorter, mainly as a result of the slower and less uniform climatological flow in the lower troposphere. The flow in the tropics is predominantly easterly. Stationary waves forced by the sources in the positions chosen here can only propagate very weakly in the traditional theory. However, in the horizontally nonuniform flow, waves with scales less than 4000 km can propagate westward and southward from the NH tropical easterlies across the embedded westerlies as far as 20°S. The cross-equatorial propagation from the NH to SH is evident over Africa, the Maritime Continent, and South America. The meridional basic wind over the northern tropics is predominantly northerly. Stationary wave propagation from the southern tropics to the NH seems to be prevented. Thus, to summarize, the interhemispheric propagation of stationary waves in DJF across the tropical easterlies tends to be restricted to one direction by the cross-equatorial flow, while that through the westerly duct in the upper troposphere is bidirectional.

Fig. 10.
Fig. 10.

Ray paths (curves) and the evolution of total wavenumber (colors) along the rays in the smoothed DJF (a),(b) 300- and (c),(d) 850-hPa flows (gray vectors), with sources in the (a),(c) NH and (b),(d) SH tropics. The light gray shading indicates the easterly wind regions. The initial zonal wavenumber is 4. The initial meridional wavenumber is derived from the dispersion relation. Where there are three solutions, the one with the largest wavelength is chosen. Propagation is terminated when the total wavenumber exceeds 40. Red points denote wave sources. The climatological flows are smoothed using a two-dimensional Fourier series truncated at zonal and meridional wavenumber 5.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

The interhemispheric stationary wave propagation behavior in JJA is shown in Fig. 11. At 300 hPa, the tropical easterlies occupy a belt at 10°S–20°N, while strong northerly cross-equatorial flows prevail over Africa, the Indian Ocean, the east Pacific, and the Atlantic. Wave rays initiated in the NH, especially over central Africa and the Asian monsoon region as well as the north and east of South America, propagate along these cross-equatorial flows and then reach middle and high latitudes in the SH. They propagate westward and southward in the easterly belt with wavelength smaller than 4000 km and then turn eastward with wavelength increasing to synoptic scale after entering the westerlies. The wave propagation initiated at the SH sources is predominantly poleward and eastward, along arc-like routes. There is no cross-equatorial propagation into the NH because of the impediment of the northerly. At the lower troposphere, significant cross-equatorial flow is observed from the south tropical Indian Ocean and Australia to South Asia and northward to East Asia. Guided by this flow, the stationary waves triggered over the south tropical Indian Ocean and north Australia propagate across the equator to South and East Asia. These waves are scaled 2000–4000 km (Fig. 11d). Additionally, evident meridional propagation of stationary waves across the easterly is found over the northwest Pacific and Atlantic, steered by the southwest flows of the subtropical highs (Fig. 11c). These eastward and northward trajectories of stationary waves will maintain an obvious meridional structure of the responses to the tropical convective heating. Meridional patterns over these two regions at low level have been observed (Nitta 1987; Kosaka and Nakamura 2006, 2010a,b). Since the cross-equatorial flow is from the SH to NH, the abovementioned meridional wave propagation is almost poleward, with negligible equatorward propagation. Consequently, it can be deduced that the interhemispheric propagation of stationary waves in JJA follows the direction of the cross-equatorial flow.

Fig. 11.
Fig. 11.

As in Fig. 10, but for the smoothed JJA flow.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

We examined the teleconnectivity of the JJA eddy streamfunction in section 2. The wave train NAA pattern is found (Figs. 1b, 2c,d). Here, we plot wave ray trajectories to determine whether this pattern results from energy dispersion by Rossby waves. Figure 12 shows the trajectories and energy dispersion speeds of stationary waves starting from 15°S, 150°E with different initial zonal wavenumbers. The supposed source is located in the region of easterlies. Based on the theory with only zonal wind in the basic-state, stationary waves from such a source cannot propagate through the easterly belt. This means that the circulation anomalies over North Africa cannot affect the SH circulation. However, the significant correlation centers in the SH suggest that there is, in reality, a close relationship between North African and SH circulation. Ray trajectories based on the horizontally nonuniform basic flow match the teleconnection pattern very well. The propagation behaviors of the different initial zonal wavenumber waves diverge south of 10°S. The wavenumber-1 wave propagates westward and poleward to 30°S, 34°W, and then turns eastward and propagates along about 27°S. The wavenumber-2–5 waves propagate along a similar trajectory, eastward and poleward until they arrive at polar regions. The trajectory connects the five correlation centers. Wavenumber-6–8 waves are reflected at lower latitudes. It seems that the conclusion that long waves propagate farther and faster than short waves is still satisfied, but it is not as strict as the situation in the zonally symmetric flow, since the zonal wavenumber varies along the ray in the horizontally nonuniform flow.

Fig. 12.
Fig. 12.

Ray paths (curves) of initial zonal wavenumber-1–8 waves (indicated by color) in the smoothed JJA 300-hPa flow with a source at 5°N, 20°E (large red dot). The markers on the rays indicate 2-day interval. Contours, shading, and filling by little red dots are as in Fig. 2b.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

b. Barotropic model results

In section 6a, wave ray tracing was employed to examine the possibility of interhemispheric propagation of stationary waves in nearly realistic climatological flows. Interhemispheric propagation across the tropical easterlies is mainly determined by the cross-equatorial flow. The westerly duct over the eastern tropical Pacific in DJF in the upper troposphere behaves as a two-way waveguide for stationary wave propagation between two hemispheres. In this section, we attempt to examine these results further using the barotropic model.

Figure 13 is the steady eddy streamfunction response to the tropical NH sources for the DJF 300-hPa flow. When the source is located in the entrance region of the North African–Asian jet, positive and negative centers occur at intervals over Eurasia and downstream. When the source is at 90°E, the downstream responses are stronger. The cross-equatorial signals can be detected over the east Pacific and are enhanced when the source moves to 150° and 210°E. The effect of the westerly duct on the waves propagating from the tropical NH to the SH is clear. In addition, the PNA-like pattern is clearly seen when the sources are at 90°, 150°, and 210°E. Figure 14 shows the eddy streamfunction responses to the SH tropical sources. The east tropical Pacific is still an important area for interhemispheric responses, but 100°–150°E is also key for cross-equatorial propagation from North Australia to Asia. This corresponds to the wave ray results in Fig. 10b. Note that the climatological flow over the Maritime Continent is easterly, while to its north the flow is southerly. This tends to favor the northward propagation of stationary waves and contributes to the interhemispheric responses. At 850 hPa, the interhemispheric responses are generally weak (figures not shown) and are effectively absent in the NH when the sources are in the tropical SH, but there are still some cross-equatorial signals when the sources are in the tropical NH.

Fig. 13.
Fig. 13.

Steady eddy streamfunction responses (thin contours, 106 m2 s−1) to the sources (thick lines, 10−10 s−2) located at (a) 15°N, 30°E, (b) 15°N, 90°E, (c) 15°N, 150°E, and (d) 15°N, 210°E in the smoothed DJF 300-hPa flow. Contour values and intervals are shown below each panel.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

Fig. 14.
Fig. 14.

As in Fig. 13, but for sources located at 15°S.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

The JJA results are shown as Figs. 15 and 16. At 300 hPa, the steady responses to sources at 20°N, 60°E and 20°N, 190°E exhibit a distinctive wave-train pattern in the SH. The forcing located over the Indian summer monsoon region can trigger a cyclonic and anticyclonic anomaly west and east of it, respectively, and then one branch of wave trains propagates northward to high latitudes in NH, one branch propagates southward into the subtropic SH. The anomalies centered over North Africa, South Africa, the south Indian Ocean, and the South Ocean, as well as the South Pole, agree with the NAA pattern indicated in Figs. 2a and 2b. The barotropic responses in subtropical SH in the model are more robust than the observation, and they have almost the same pattern with globally zonal wavenumber 3 when the forcing moves over the tropical Pacific. Compared with the NAA pattern, the meridional structure of the NPSA is less described by the barotropic model. When the source moved to the tropical SH, the zonal wavenumber-3 pattern is stronger, especially when the source is at the entrance region of the jet. However, the responses in the NH to the SH forcings are much weaker than those in the SH to the NH forcings. The interhemispheric responses only occur from the NH to the SH at the upper troposphere in JJA in the same direction as the cross-equatorial flow.

Fig. 15.
Fig. 15.

As in Fig. 13, but for the JJA flow with sources at (a) 20°N, 60°E, (b) 20°N, 190°E, (c) 20°S, 60°E, and (d) 20°S, 210°E.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

Fig. 16.
Fig. 16.

As in Fig. 15, but for the 850-hPa flow with sources at (a) 15°N, 150°E, (b) 15°N, 300°E, (c) 15°S, 90°E, and (d) 15°S, 150°E.

Citation: Journal of the Atmospheric Sciences 72, 8; 10.1175/JAS-D-14-0239.1

At 850 hPa, a pattern remarkably similar to the PJ pattern can be seen when the source is at 15°N, 150°E, and another wave train pattern is found over North America in response to the source at 15°N, 300°E. The latter is in good agreement with the meridional teleconnection described by Kosaka and Nakamura (2010b). Nevertheless, few signals are found in the SH when sources are set in the NH. However, significant positive responses are found over South Asia to the north of the local anomalies triggered by the vorticity sources at 15°S, 90°E and 15°S, 150°E. When the source moves to other longitudes in the tropical SH, there is negligible cross-equatorial response. These longitude-dependent responses are attributed to the strong cross-equatorial flows from the SH to NH over the Australian–Asian monsoon region, which may transmit the wave energy across the easterlies in their directions.

7. Discussion and conclusions

In this study, we examined the interhemispheric propagation of stationary Rossby waves using statistical and theoretical analysis. Teleconnections of the JJA streamfunction field at 300 hPa were investigated using a correlation method (Wallace and Gutzler 1981). Significant correlations between the streamfunction in the NH and tropical SH were found. Two wave train patterns in the eddy streamfunction were identified: one over North Africa–Antarctica and another over the North Pacific–Atlantic sector. The former links the circulation anomalies over North Africa and the South Pole and is known as the NAA pattern. The latter is similar to the PSA pattern, but with an additional center located in the north tropical Pacific, and is referred to as the NPSA pattern. Both patterns straddle the tropical easterly belt. It is implied by our results that the direct propagation of stationary waves from one hemisphere to the other may be responsible for these interhemispheric teleconnections.

The possibility of interhemispheric propagation for stationary waves is presented based on two-dimensional spherical Rossby wave theory under the horizontally nonuniform basic state. The meridional group velocity at the traditional critical latitudes (where ) is shown in the same direction as the meridional background wind. This means that southerly background flow would transmit Rossby wave energy from the SH to NH and impede transmission from the NH to SH. A northerly background wind would give the opposite response. Wave ray tracing and barotropic model results strongly support the wave theory. Although Schneider and Watterson (1984) and Watterson and Schneider (1987) have reported similar findings, they focused on the zonal mean Hadley circulation, neglected the zonal variation in the basic state. The meridional wind displays significant regional features, differing much in the monsoon and nonmonsoon regions. Hence, the extension of previous studies into a realistic basic state would provide many insights into the interhemispheric teleconnections in the real atmosphere.

The paths of interhemispheric stationary wave propagation in the climatological upper- and lower-troposphere flows are investigated using the ray tracing method and barotropic model. In DJF, tunnels of interhemispheric stationary wave propagation are found in the upper troposphere over the east tropical Pacific and the Maritime Continent. The east Pacific tunnel is the westerly duct, which supports bidirectional propagation. However, propagation through the Maritime Continent is unidirectional because of the basic meridional wind. In JJA, the cross-equatorial propagation is from the NH to SH in the upper troposphere and from the SH to NH in the lower troposphere, consistent with the cross-equatorial flows. Because of the remarkable cross-equatorial flows, the Australian–Asian monsoon region is key for interhemispheric stationary wave propagation, which implies that the stationary waves may transmit the monsoonal heating far away from the monsoon region, even to the middle to high latitudes in the other hemisphere. Several modeling works recently suggest that stationary barotropic waves play an important role in the influence of summertime monsoon heating in the tropical NH on the circulations in the SH (Kraucunas and Hartmann 2007; Liu and Wang 2013; Lee et al. 2013). Therefore, this study provides an insight into global responses to the Asian monsoon (Lin 2009). Note that these results are based on a barotropic nondivergent model. How the tropical divergence will change the results needs to be investigated, and verification using a baroclinic model is desired in the future.

We emphasized the modulation by the meridional background wind of the interhemispheric propagation of stationary waves in the text. Since the zonal background wind is an order of magnitude larger than the meridional background wind at middle and high latitudes, the waveguide effect of the westerly jets may dominate the wave propagation (Naoe et al. 1997; Enomoto and Matsuda 1999; Müller and Ambrizzi 2007). Even so, our results suggest that the inclusion of the meridional background wind in the wave ray theory may better describe the actual teleconnection patterns along the jets. Shaman et al. (2012) extended the wave ray method to complex wavenumbers, but here we still consider real wavenumbers.

Previous studies (Dima et al. 2005; Kraucunas and Hartmann 2007) suggested the important role of the cross-equatorial momentum flux by the mean meridional circulation in accelerating the tropical easterly and the formation of the equatorial symmetric Rossby wave response to off-equatorial forcing. Hence, the transport of the momentum and energy by the interhemispheric propagating Rossby waves is an important and essential topic for understanding the general atmospheric circulation variations. An elaborate investigation would be worthwhile in the future.

Acknowledgments

The author Yanjie Li acknowledges useful discussions with David Karoly, Fred Kucharski, and In-Sik Kang during her visit to the Abdus Salam International Centre for Theoretical Physics (ICTP). We thank the European Centre for Medium-Range Weather Forecasts for providing the reanalysis data. This work is supported by the National Natural Science Foundation of China (Grants 41030961, 41205034, and 91437216).

REFERENCES

  • Alexander, M. A., I. Bladé, M. Newman, J. R. Lanzante, N.-C. Lau, and J. D. Scott, 2002: The atmospheric bridge: The influence of ENSO teleconnections on air–sea interaction over the global oceans. J. Climate, 15, 22052231, doi:10.1175/1520-0442(2002)015<2205:TABTIO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Ambrizzi, T., and B. J. Hoskins, 1997: Stationary rossby-wave propagation in a baroclinic atmosphere. Quart. J. Roy. Meteor. Soc., 123, 919928, doi:10.1002/qj.49712354007.

    • Search Google Scholar
    • Export Citation
  • Ambrizzi, T., B. J. Hoskins, and H.-H. Hsu, 1995: Rossby wave propagation and teleconnection patterns in the austral winter. J. Atmos. Sci., 52, 36613672, doi:10.1175/1520-0469(1995)052<3661:RWPATP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Barnston, A. G., and R. E. Livezey, 1987: Classification, seasonality and persistence of low-frequency atmospheric circulation patterns. Mon. Wea. Rev., 115, 10831126, doi:10.1175/1520-0493(1987)115<1083:CSAPOL>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Berbery, E. H., J. Nogués-Paegle, and J. D. Horel, 1992: Wavelike Southern Hemisphere extratropical teleconnections. J. Atmos. Sci., 49, 155177, doi:10.1175/1520-0469(1992)049<0155:WSHET>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Blackmon, M. L., Y.-H. Lee, and J. M. Wallace, 1984a: Horizontal structure of 500 mb height fluctuations with long, intermediate and short time scales. J. Atmos. Sci., 41, 961980, doi:10.1175/1520-0469(1984)041<0961:HSOMHF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Blackmon, M. L., Y.-H. Lee, J. M. Wallace, and H.-H. Hsu, 1984b: Time variation of 500 mb height fluctuations with long, intermediate and short time scales as deduced from lag-correlation statistics. J. Atmos. Sci., 41, 981991, doi:10.1175/1520-0469(1984)041<0981:TVOMHF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Branstator, G., 1983: Horizontal energy propagation in a barotropic atmosphere with meridional and zonal structure. J. Atmos. Sci., 40, 16891708, doi:10.1175/1520-0469(1983)040<1689:HEPIAB>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Chen, W. Y., 1982: Fluctuations in Northern Hemisphere 700-mb height field associated with the Southern Oscillation. Mon. Wea. Rev., 110, 808823, doi:10.1175/1520-0493(1982)110<0808:FINHMH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dima, I. M., J. M. Wallace, and I. Kraucunas, 2005: Tropical zonal momentum balance in the NCEP reanalyses. J. Atmos. Sci., 62, 24992513, doi:10.1175/JAS3486.1.

    • Search Google Scholar
    • Export Citation
  • Ding, Q., and B. Wang, 2005: Circumglobal teleconnection in the Northern Hemisphere summer. J. Climate, 18, 34833505, doi:10.1175/JCLI3473.1.

    • Search Google Scholar
    • Export Citation
  • Ding, Q., B. Wang, J. M. Wallace, and G. Branstator, 2011: Tropical–extratropical teleconnections in boreal summer: Observed interannual variability. J. Climate, 24, 18781896, doi:10.1175/2011JCLI3621.1.

    • Search Google Scholar
    • Export Citation
  • Enomoto, T., and Y. Matsuda, 1999: Rossby wavepacket propagation in a zonally-varying basic flow. Tellus, 51A, 588602, doi:10.1034/j.1600-0870.1999.00004.x.

    • Search Google Scholar
    • Export Citation
  • Enomoto, T., B. J. Hoskins, and Y. Matsuda, 2003: The formation mechanism of the Bonin high in August. Quart. J. Roy. Meteor. Soc., 129, 157178, doi:10.1256/qj.01.211.

    • Search Google Scholar
    • Export Citation
  • Esbensen, S. K., 1984: A comparison of intermonthly and interannual teleconnections in the 700 mb geopotential height field during the Northern Hemisphere winter. Mon. Wea. Rev., 112, 20162032, doi:10.1175/1520-0493(1984)112<2016:ACOIAI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Esler, J. G., L. M. Polvani, and R. A. Plumb, 2000: The effect of a Hadley circulation on the propagation and reflection of planetary waves in a simple one-layer model. J. Atmos. Sci., 57, 15361556, doi:10.1175/1520-0469(2000)057<1536:TEOAHC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Farrell, B., and I. G. Watterson, 1985: Rossby waves in opposing currents. J. Atmos. Sci., 42, 17461755, doi:10.1175/1520-0469(1985)042<1746:RWIOC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Gong, D., and S. Wang, 1999: Definition of Antarctic Oscillation Index. Geophys. Res. Lett., 26, 459462, doi:10.1029/1999GL900003.

  • Hartmann, D. L., 2007: The atmospheric general circulation and its variability. J. Meteor. Soc. Japan, 85B, 123143.

  • He, J. H., 1990: Discussion of meridional propagation mechanism of quasi-40-day oscillation. Adv. Atmos. Sci., 7, 7886, doi:10.1007/BF02919170.

    • Search Google Scholar
    • Export Citation
  • Held, I. M., M. Ting, and H. Wang, 2002: Northern winter stationary waves: Theory and modeling. J. Climate, 15, 21252144, doi:10.1175/1520-0442(2002)015<2125:NWSWTA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hoskins, B. J., and D. J. Karoly, 1981: The steady linear response of a spherical atmosphere to thermal and orographic forcing. J. Atmos. Sci., 38, 11791196, doi:10.1175/1520-0469(1981)038<1179:TSLROA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hoskins, B. J., and T. Ambrizzi, 1993: Rossby wave propagation on a realistic longitudinally varying flow. J. Atmos. Sci., 50, 16611661, doi:10.1175/1520-0469(1993)050<1661:RWPOAR>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hoskins, B. J., A. J. Simmons, and D. G. Andrews, 1977: Energy dispersion in a barotropic atmosphere. Quart. J. Roy. Meteor. Soc., 103, 553567, doi:10.1002/qj.49710343802.

    • Search Google Scholar
    • Export Citation
  • Hsu, H.-H., and S.-H. Lin, 1992: Global teleconnections in the 250-mb streamfunction field during the Northern Hemisphere winter. Mon. Wea. Rev., 120, 11691190, doi:10.1175/1520-0493(1992)120<1169:GTITMS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Huang, R., and L. Lu, 1989: Numerical simulation of the relationship between the anomaly of subtropical high over East Asia and the convective activities in the western tropical Pacific. Adv. Atmos. Sci., 6, 202214, doi:10.1007/BF02658016.

    • Search Google Scholar
    • Export Citation
  • Huang, R., and F. Sun, 1992: Impacts of the tropical western Pacific on the East Asian summer monsoon. J. Meteor. Soc. Japan, 70, 243256.

    • Search Google Scholar
    • Export Citation
  • Ji, X., J. D. Neelin, S.-K. Lee, and C. R. Mechoso, 2014: Interhemispheric teleconnections from tropical heat sources in intermediate and simple models. J. Climate, 27, 684697, doi:10.1175/JCLI-D-13-00017.1.

    • Search Google Scholar
    • Export Citation
  • Jin, F., and B. J. Hoskins, 1995: The direct response to tropical heating in a baroclinic atmosphere. J. Atmos. Sci., 52, 307319, doi:10.1175/1520-0469(1995)052<0307:TDRTTH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Karoly, D. J., 1983: Rossby wave propagation in a barotropic atmosphere. Dyn. Atmos. Oceans, 7, 111125, doi:10.1016/0377-0265(83)90013-1.

    • Search Google Scholar
    • Export Citation
  • Karoly, D. J., 1989: Southern Hemisphere circulation features associated with El Niño–Southern Oscillation events. J. Climate, 2, 12391252, doi:10.1175/1520-0442(1989)002<1239:SHCFAW>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Karoly, D. J., 1990: The role of transient eddies in low-frequency zonal variations of the Southern Hemisphere circulation. Tellus, 42A, 4150, doi:10.1034/j.1600-0870.1990.00005.x.

    • Search Google Scholar
    • Export Citation
  • Kidson, J. W., 1999: Principal modes of Southern Hemisphere low-frequency variability obtained from NCEP–NCAR reanalyses. J. Climate, 12, 28082830, doi:10.1175/1520-0442(1999)012<2808:PMOSHL>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Kosaka, Y., and H. Nakamura, 2006: Structure and dynamics of the summertime Pacific–Japan teleconnection pattern. Quart. J. Roy. Meteor. Soc., 132, 20092030, doi:10.1256/qj.05.204.

    • Search Google Scholar
    • Export Citation
  • Kosaka, Y., and H. Nakamura, 2010a: Mechanisms of meridional teleconnection observed between a summer monsoon system and a subtropical anticyclone. Part I: The Pacific–Japan pattern. J. Climate, 23, 50855108, doi:10.1175/2010JCLI3413.1.

    • Search Google Scholar
    • Export Citation
  • Kosaka, Y., and H. Nakamura, 2010b: Mechanisms of meridional teleconnection observed between a summer monsoon system and a subtropical anticyclone. Part II: A global survey. J. Climate, 23, 51095125, doi:10.1175/2010JCLI3414.1.

    • Search Google Scholar
    • Export Citation
  • Kraucunas, I., and D. L. Hartmann, 2007: Tropical stationary waves in a nonlinear shallow-water model with realistic basic states. J. Atmos. Sci., 64, 25402557, doi:10.1175/JAS3920.1.

    • Search Google Scholar
    • Export Citation
  • Krishnamurti, T., C. Wagner, T. J. Cartwright, and D. Oosterhof, 1997: Wave trains excited by cross-equatorial passage of the monsoon annual cycle. Mon. Wea. Rev., 125, 27092715, doi:10.1175/1520-0493(1997)125<2709:WTEBCE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lau, K.-M., P.-J. Sheu, and I.-S. Kang, 1994: Multiscale low-frequency circulation modes in the global atmosphere. J. Atmos. Sci., 51, 11691193, doi:10.1175/1520-0469(1994)051<1169:MLFCMI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lau, N.-C., and M. J. Nath, 1996: The role of the “atmospheric bridge” in linking tropical Pacific ENSO events to extratropical SST anomalies. J. Climate, 9, 20362057, doi:10.1175/1520-0442(1996)009<2036:TROTBI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lee, S.-K., C. R. Mechoso, C. Wang, and J. D. Neelin, 2013: Interhemispheric influence of the northern summer monsoons on southern subtropical anticyclones. J. Climate, 26, 10 19310 204, doi:10.1175/JCLI-D-13-00106.1.

    • Search Google Scholar
    • Export Citation
  • Li, C., 1991: Global characteristics of 30–60 day atmospheric oscillation (in Chinese). Chin. J. Atmos. Sci., 15, 6676.

  • Li, J., and Coauthors, 2013: Progress in air–land–sea interactions in Asia and their role in global and Asian climate change (in Chinese). Chin. J. Atmos. Sci., 37, 518538.

    • Search Google Scholar
    • Export Citation
  • Li, L., and T. R. Nathan, 1994: The global atmospheric response to low-frequency tropical forcing: Zonally averaged basic states. J. Atmos. Sci., 51, 34123426, doi:10.1175/1520-0469(1994)051<3412:TGARTL>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Li, L., and T. R. Nathan, 1997: Effects of low-frequency tropical forcing on intraseasonal tropical–extratropical interactions. J. Atmos. Sci., 54, 332346, doi:10.1175/1520-0469(1997)054<0332:EOLFTF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Li, Y., and J. Li, 2012: Propagation of planetary waves in the horizonal non-uniform basic flow (in Chinese). Chin. J. Geophys., 55, 361371.

    • Search Google Scholar
    • Export Citation
  • Li, Y., J. Li, and J. Feng, 2013: Boreal summer convection oscillation over the Indo-Western Pacific and its relationship with the East Asian summer monsoon. Atmos. Sci. Lett., 14, 6671, doi:10.1002/asl2.418.

    • Search Google Scholar
    • Export Citation
  • Lighthill, J., 1978: Waves in Fluids. Cambridge University Press, 504 pp.

  • Lin, H., 2009: Global extratropical response to diabatic heating variability of the Asian summer monsoon. J. Atmos. Sci., 66, 26972713, doi:10.1175/2009JAS3008.1.

    • Search Google Scholar
    • Export Citation
  • Liu, F., and B. Wang, 2013: Mechanisms of global teleconnections associated with the Asian summer monsoon: An intermediate model analysis. J. Climate, 26, 17911806, doi:10.1175/JCLI-D-12-00243.1.

    • Search Google Scholar
    • Export Citation
  • Liu, Z., and M. Alexander, 2007: Atmospheric bridge, oceanic tunnel, and global climatic teleconnections. Rev. Geophys., 45, RG2005, doi:10.1029/2005RG000172.

    • Search Google Scholar
    • Export Citation
  • Longuet-Higgins, M., 1964: Planetary waves on a rotating sphere. Proc. Roy. Soc. London, 279A, 446473, doi:10.1098/rspa.1964.0116.

  • Longuet-Higgins, M., 1965: Planetary waves on a rotating sphere. II. Proc. Roy. Soc. London, 284A, 4068, doi:10.1098/rspa.1965.0051.

  • Lu, R., J.-H. Oh, and B.-J. Kim, 2002: A teleconnection pattern in upper-level meridional wind over the North African and Eurasian continent in summer. Tellus, 54A, 4455, doi:10.1034/j.1600-0870.2002.00248.x.

    • Search Google Scholar
    • Export Citation
  • Mo, K. C., and G. H. White, 1985: Teleconnection patterns in the Southern Hemisphere. Mon. Wea. Rev., 113, 2237, doi:10.1175/1520-0493(1985)113<0022:TITSH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Mo, K. C., and R. W. Higgins, 1998: The Pacific–South American modes and tropical convection during the Southern Hemisphere winter. Mon. Wea. Rev., 126, 15811596, doi:10.1175/1520-0493(1998)126<1581:TPSAMA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Müller, G., and T. Ambrizzi, 2007: Teleconnection patterns and Rossby wave propagation associated to generalized frosts over southern South America. Climate Dyn., 29, 633645, doi:10.1007/s00382-007-0253-x.

    • Search Google Scholar
    • Export Citation
  • Nakamura, H., M. Tanaka, and J. M. Wallace, 1987: Horizontal structure and energetics of Northern Hemisphere wintertime teleconnection patterns. J. Atmos. Sci., 44, 33773391, doi:10.1175/1520-0469(1987)044<3377:HSAEON>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Naoe, H., Y. Matsuda, and H. Nakamura, 1997: Rossby wave propagation in idealized and realistic zonally varying flows. J. Meteor. Soc. Japan, 75, 687700.

    • Search Google Scholar
    • Export Citation
  • Nitta, T., 1987: Convective activities in the tropical western Pacific and their impact on the Northern Hemisphere summer circulation. J. Meteor. Soc. Japan, 65, 373390.

    • Search Google Scholar
    • Export Citation
  • Nitta, T., 1989: Global features of the Pacific–Japan oscillation. Meteor. Atmos. Phys., 41, 512, doi:10.1007/BF01032585.

  • Rogers, J. C., and H. van Loon, 1982: Spatial variability of sea level pressure and 500 mb height anomalies over the Southern Hemisphere. Mon. Wea. Rev., 110, 13751392, doi:10.1175/1520-0493(1982)110<1375:SVOSLP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Rossby, C.-G., 1945: On the propagation of frequencies and energy in certain types of oceanic and atmospheric waves. J. Meteor., 2, 187204, doi:10.1175/1520-0469(1945)002<0187:OTPOFA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Sardeshmukh, P. D., and B. J. Hoskins, 1988: The generation of global rotational flow by steady idealized tropical divergence. J. Atmos. Sci., 45, 12281251, doi:10.1175/1520-0469(1988)045<1228:TGOGRF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Schneider, E. K., and I. G. Watterson, 1984: Stationary Rossby wave propagation through easterly layers. J. Atmos. Sci., 41, 20692083, doi:10.1175/1520-0469(1984)041<2069:SRWPTE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Shaman, J., and E. Tziperman, 2005: The effect of ENSO on Tibetan Plateau snow depth: A stationary wave teleconnection mechanism and implications for the South Asian monsoons. J. Climate, 18, 20672079, doi:10.1175/JCLI3391.1.

    • Search Google Scholar
    • Export Citation
  • Shaman, J., and E. Tziperman, 2007: Summertime ENSO–North African–Asian jet teleconnection and implications for the Indian monsoons. Geophys. Res. Lett., 34, L11702, doi:10.1029/2006GL029143.

    • Search Google Scholar
    • Export Citation
  • Shaman, J., R. M. Samelson, and E. Tziperman, 2012: Complex wavenumber Rossby wave ray tracing. J. Atmos. Sci., 69, 21122133, doi:10.1175/JAS-D-11-0193.1.

    • Search Google Scholar
    • Export Citation
  • Simmons, A., 1982: The forcing of stationary wave motion by tropical diabatic heating. Quart. J. Roy. Meteor. Soc., 108, 503534, doi:10.1002/qj.49710845703.

    • Search Google Scholar
    • Export Citation
  • Sun, C., and J. Li, 2012: Space–time spectral analysis of the Southern Hemisphere daily 500-hPa geopotential height. Mon. Wea. Rev., 140, 38443856, doi:10.1175/MWR-D-12-00019.1.

    • Search Google Scholar
    • Export Citation
  • Ting, M., 1996: Steady linear response to tropical heating in barotropic and baroclinic models. J. Atmos. Sci., 53, 16981709, doi:10.1175/1520-0469(1996)053<1698:SLRTTH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Tomas, R. A., and P. J. Webster, 1994: Horizontal and vertical structure of cross-equatorial wave propagation. J. Atmos. Sci., 51, 14171430, doi:10.1175/1520-0469(1994)051<1417:HAVSOC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Trenberth, K. E., G. W. Branstator, D. Karoly, A. Kumar, N.-C. Lau, and C. Ropelewski, 1998: Progress during TOGA in understanding and modeling global teleconnections associated with tropical sea surface temperatures. J. Geophys. Res., 103, 14 29114 324, doi:10.1029/97JC01444.

    • Search Google Scholar
    • Export Citation
  • Uppala, S. M., and Coauthors, 2005: The ERA-40 Re-Analysis. Quart. J. Roy. Meteor. Soc., 131, 29613012, doi:10.1256/qj.04.176.

  • Van den Dool, H. M., and R. E. Livezy, 1984: Geographical distribution and seasonality of month-to-month correlation of monthly mean 700 mb heights. Mon. Wea. Rev., 112, 610615, doi:10.1175/1520-0493(1984)112<0610:GDASOM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Wallace, J. M., and D. S. Gutzler, 1981: Teleconnections in the geopotential height field during the Northern Hemisphere winter. Mon. Wea. Rev., 109, 784812, doi:10.1175/1520-0493(1981)109<0784:TITGHF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Wang, Z., C.-P. Chang, B. Wang, and F.-F. Jin, 2005: Teleconnections from tropics to northern extratropics through a southerly conveyor. J. Atmos. Sci., 62, 40574070, doi:10.1175/JAS3600.1.

    • Search Google Scholar
    • Export Citation
  • Watterson, I. G., and E. K. Schneider, 1987: The effect of the Hadley circulation on the meridional propagation of stationary waves. Quart. J. Roy. Meteor. Soc., 113, 779813, doi:10.1002/qj.49711347706.

    • Search Google Scholar
    • Export Citation
  • 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, 722733, doi:10.1175/1520-0469(1982)039<0722:CERTML>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Whitham, G., 1960: A note on group velocity. J. Fluid Mech., 9, 347352, doi:10.1017/S0022112060001158.

  • Wu, Z., B. Wang, J. Li, and F.-F. Jin, 2009: An empirical seasonal prediction model of the East Asian summer monsoon using ENSO and NAO. J. Geophys. Res., 114, D18120, doi:10.1029/2009JD011733.

    • Search Google Scholar
    • Export Citation
  • Yang, G.-Y., and B. J. Hoskins, 1996: Propagation of Rossby waves of nonzero frequency. J. Atmos. Sci., 53, 23652378, doi:10.1175/1520-0469(1996)053<2365:PORWON>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Yeh, T.,