1. Introduction
Tropical convection anomalies, such as those driven by anomalous sea surface temperatures (SSTs) associated with El Niño–Southern Oscillation (ENSO) and the Indian Ocean dipole (IOD), have important influences on the atmospheric circulation at both tropical and extratropical latitudes. The Matsuno–Gill (Matsuno 1966; Gill 1980) model describes the equatorially trapped response to steady equatorial heating, consisting of a Kelvin wave to the east of the heating and Rossby waves to the west. This thermally direct Rossby wave response has a deep baroclinic structure with upper-level anticyclones and lower-level cyclones straddling the equator (e.g., Jin and Hoskins 1995). Away from the equator, equivalent-barotropic Rossby waves propagate poleward and eastward (e.g., Hoskins and Karoly 1981). They develop because tropical heating induces upper-level divergence that causes vortex stretching and also advects mean vorticity out of the tropics (e.g., Sardeshmukh and Hoskins 1985), thus producing a Rossby wave source (RWS) in the subtropical westerlies. Therefore, these wave trains enable anomalous tropical convection within the upper-tropospheric easterlies to influence extratropical climate (e.g., Karoly 1989; Li et al. 2015a).
The theory for the initiation (e.g., Sardeshmukh and Hoskins 1985) and propagation (e.g., Hoskins and Karoly 1981) of stationary Rossby waves, from a steady source, is well established for zonally symmetric flow but it also appears to work for zonally asymmetric flows (e.g., Hoskins and Ambrizzi 1993). During SH winter [June–August (JJA)], the major zonal asymmetry of the zonal winds is the subtropical jet (STJ), which extends from the eastern Indian Ocean (EIO) to the western Pacific near 27°S (e.g., Gillett et al. 2021). It strongly influences both the RWS due to the maximum in the meridional gradient of absolute vorticity
Closer inspection of the wintertime IOD-induced Rossby wave train (McIntosh and Hendon 2017) suggests that the group velocity goes to zero as it encounters the region where
Stationary Rossby wave propagation is also prohibited in the upper-tropospheric equatorial easterlies. Studies have demonstrated that cross-equatorial propagation can still occur due to the establishment of a RWS (e.g., Sardeshmukh and Hoskins 1985), and also by including the background meridional flow in the barotropic dispersion relation, which enables direct propagation when it is sufficiently strong by acting as a one-way tunnel (e.g., Schneider and Watterson 1984; Li et al. 2015c; Zhao et al. 2015).
The time-mean response to tropical heating is well established in linearized models (e.g., Hoskins and Karoly 1981; Branstator 1983) and in atmospheric general circulation models (AGCMs; e.g., Yiu and Maycock 2019). However, few studies have explored the time-dependent response in an AGCM, which has the advantage of revealing how the time-mean response is established. Therefore, we analyze the temporal evolution of the response to switched-on tropical heating in a comprehensive AGCM and a linear model. We focus on the austral winter months when the STJ is strongest to examine its role in producing the effective RWS and providing reflective effects. Our goal is to elucidate how the wave train that emerges from the tropical Indian Ocean can traverse the
This teleconnection is not just of academic interest as it promotes Australian drought (e.g., Ummenhofer et al. 2009; King et al. 2020), bushfires (e.g., Cai et al. 2009) and temperature extremes (e.g., McKay et al. 2021), variations in Antarctic sea ice extent (e.g., Meehl et al. 2019; Purich and England 2019; Wang et al. 2019), and influences the Antarctic stratosphere (e.g., Lim et al. 2020; Rao et al. 2020). Therefore, it is pertinent to have an improved understanding of its dynamics.
2. Data and numerical experiments
a. Observational and reanalysis datasets
Wintertime teleconnections between the tropical Indian Ocean and SH extratropics are analyzed using observational and reanalysis datasets from 1979 to 2019. Monthly mean heights and winds are taken from the Japanese 55-year Reanalysis (JRA-55; 1.25° grid; Kobayashi et al. 2015). Gridded (2.5°) monthly mean outgoing longwave radiation (OLR) is used to monitor tropical convection variations and is taken from Liebmann and Smith (1996).
Gridded (1°) monthly mean SST is obtained from the merged Hadley-NOAA/Optimal Interpolation SST analyses (hereafter Hadley-OI; Hurrell et al. 2008). The dipole mode index (DMI; Saji et al. 1999) is used as an indicator for the IOD and is calculated as the anomalous (from the monthly climatology) SST difference between the tropical western Indian Ocean (WIO; 10°S–10°N, 50°–70°E) and EIO (10°S–0°, 90°–110°E) using Hadley-OI. We do not linearly remove the effect of ENSO from the DMI as in Cai et al. (2011) because the indices are weakly correlated in winter, and the IOD wave train pattern is similar with and without the removal of ENSO (Cai et al. 2011).
To develop the observed IOD teleconnection, the detrended and normalized austral winter seasonal-mean DMI is regressed against detrended OLR, heights and winds. The regressed height and wind anomalies are also used to derive the diagnostics described in section 2d. Statistical significance of the regressed anomalies is assessed using a two-tailed t test assuming 40 degrees of freedom.
b. Community Atmosphere Model
1) Model description
We use the National Center for Atmospheric Research Community Atmosphere Model, version 5.1 (CAM5; Neale et al. 2012). CAM5 is the atmospheric component of the Community Earth System Model (CESM1.2.2; Hurrell et al. 2013) and is coupled to an active land model (Community Land Model CLM4; Oleson et al. 2010) with prescribed SSTs and sea ice concentrations from Hadley-OI. CAM5 is run with a 0.9° × 1.25° horizontal grid and 30 hybrid sigma–pressure levels.
2) Experimental setup
A control simulation uses repeating monthly varying SSTs and sea ice averaged over 1982–2001 to represent present-day climate. Other boundary conditions (greenhouse gases, ozone, aerosols, etc.) are prescribed to be year 2000 conditions. We generated a control ensemble by integrating CAM5 30 times for one year starting from 1 January 2000 initial conditions, with a small change to the initial atmospheric temperature applied on the first day using the “pertlim” parameter following Kay et al. (2015). We output a restart file on the first of June, July, and August for each of these 30 integrations. This provides 90 independent initial conditions across the three start dates and thus samples the full winter variability.
The CAM5 control simulates a reasonable wintertime STJ with similar magnitude and structure to that depicted with JRA-55 [see Fig. S1 in the online supplemental material, which displays the observed and simulated terms in Eq. (1)]. The large
The model responds to the imposed heating by producing additional enhanced rainfall primarily collocated with the imposed heating but also nonlocally. The total heating diabatic anomaly (i.e., sensitivity experiment minus control) averaged over the 30-day integration is shown in Fig. S2. Compensating downward motion occurs primarily to the west but also to the east (Fig. S2). The vertical profile of the induced diabatic heating anomaly averaged over the patch and 30-day period is overlaid in Fig. 1b (black curve). It shows that this induced heating is about twice as large as the imposed heating and is dominated by the induced latent heating anomaly (DTCOND; purple curve). Also important for our understanding of the induced RWS, these induced heating anomalies rapidly develop (even by day 1 it is of comparable magnitude as the imposed heating) and appears to saturate by day 3 (not shown). It is also clear in Fig. 1b that the vertical profile of the induced total heating anomaly has the shape of the induced DTCOND (i.e., it peaks above 500 hPa), although the imposed heating anomaly peaks below 500 hPa. Limited experimentation using a deeper prescribed heating profile (not shown) reveals little sensitivity of the response, presumably because the total heating anomaly is dominated by DTCOND that chooses its own profile irrespective of the prescribed profile.
The experiment design using 90 pairs of experiment and control simulations allows for the depiction of the daily (0000 UTC) evolution of the forced response to anomalous heating. However, later in the integration (beyond ∼2 weeks), the experiment and control responses for individual initial conditions become very different, therefore a much larger sample size is required to recover the forced component. Hence, we focus on the daily evolution during the first two weeks of the integration, which is the typical time for the high-latitude response to be established (e.g., Jin and Hoskins 1995), and average over the final two weeks to better extract the time-mean response.
Our focus is on the response to an imposed positive heating anomaly, but we also performed limited experimentation with a negative heating anomaly (not shown). For the first week, the negative heating produces a similar structure to the response to positive heating but with reversed sign. However, the magnitude of the induced negative heating anomaly is much weaker (∼2–3 times) than in the positive case, which results presumably from the positive-only nature of rainfall (i.e., the maximum negative rainfall anomaly is limited by the background rainfall rate). Hence, later in the integration, the negative heating results are noisier than the positive heating results and it is harder to recover the forced component. Therefore, we have not undertaken further analysis of potential nonlinearities in this study.
c. Linear simple general circulation model
1) Model description
The comprehensive AGCM (CAM5) results are compared to similar experiments using a linear simplified AGCM (SGCM; Hall 2000). The SGCM is a dry primitive equation atmospheric model that is linearized about an “observed” three-dimensional climatological basic state for the austral winter season. This basic state is the time mean of a long integration of the nonlinear version of the model with a prescribed constant forcing, which is calculated from observed daily data and acts to maintain the observed basic state. The basic state of the model is therefore close to the observed basic state. However, it may not be exactly the same because the time-mean of the eddy fluxes in the model is not constrained to be identical to the observations, thus resulting in slight biases in the mean flow. The SGCM is run at triangular 31 resolution (∼3.75°) and 10 sigma–vertical levels.
2) Experimental setup
The SGCM is converted into a linear perturbation model and is run on a time-independent basic state (perpetual JJA; e.g., Lin et al. 2010; Lin and Brunet 2018). Similar to the CAM5 simulations, a thermal forcing is applied centered on the equator that follows an elliptical form in the horizontal and half sine wave in the vertical, but which follows υ(σ) = (1 − σ)sin[π(1 − σ)] and peaks at σ = 0.35. The vertical profile is the same as in Lin and Brunet (2018; see their Fig. 2) but A = 4.5 K day−1. Here λw = 40° and ϕw = 11°, thus λw is double the length used in the CAM5 simulations. However, the similarity of the SGCM and CAM5 responses suggests that this difference has little impact on our results. Nonetheless, Ambrizzi and Hoskins (1997) found that zonal elongation of an applied forcing increased the amplitude of meridional propagation, which we have not explored further. As this is a linearized model, the response scales with the magnitude of the prescribed heating and is identical but opposite-signed to prescribed negative heating. Like the CAM5 simulation, we archive instantaneous daily output for 20 days after the heating is switched on and conduct experiments for a heating patch centered at (100°E, 0°).
d. Rossby wave diagnostics
1) Rossby wave source
2) Wave activity flux
e. Eddy feedback
3. Results
a. Observed Indian Ocean teleconnection
We first review the observed wintertime teleconnection driven by the IOD (e.g., Cai et al. 2011; McIntosh and Hendon 2017). Figure 2 shows the regression coefficients of austral winter mean fields onto the negative DMI (one standard deviation anomaly). The negative phase of the IOD, which has anomalously warm SSTs in the EIO and anomalously cool SSTs in the WIO, is associated with enhanced convection over the EIO (negative OLR values shaded in blue in Fig. 2a), flanked by weaker suppressed convection to the west and east (positive OLR values shaded in red). The enhanced EIO convection leads to increased upper-level divergent outflow toward the south (vectors in Fig. 2a) and an anticyclonic S1 RWS (solid blue contours in Fig. 2a) in the subtropics to the west of Australia. A weaker cyclonic S2 RWS (dashed orange contours) occurs farther to the south due to convergence of this upper-level outflow near 30°S. An opposite-signed S1–S2 couplet also forms downstream (east) of the first couplet over southeastern Australia, with a strong anticyclonic S2 RWS (solid orange contours), which is discussed in section 3b.
The anticyclonic S1 term maximizes to the west of Australia in association with large
In summary, the region where
b. Evolution of the response to eastern Indian Ocean heating in CAM5
To explore the development of the IOD wave train, we examine the daily evolution of the forced response to imposed, switched-on diabatic heating in the EIO in CAM5 (Fig. 1). The 250-hPa height (shading) and horizontal WAF (vectors) anomalies are shown in Fig. 4 for every other day. The direct tropical response resembles the Matsuno–Gill pattern with anticyclonic anomalies straddling the heating to the west and a Kelvin wave front emerging along the equator to the east. It is interesting to observe that the Kelvin wave takes ∼11 days to reach South America, whereas a dry first baroclinic wave should encircle the entire globe in ∼10 days (e.g., Wheeler and Nguyen 2015). Inspection of the rainfall anomalies along the equator (not shown) reveal that the leading edge of this Kelvin wave is associated with a negative rainfall anomaly, suggestive that it is better described as being convectively coupled and therefore travels eastward much slower than a dry wave with the same deep baroclinic structure (e.g., Wheeler et al. 2000).
Away from the equator the initial response in the first week appears as a wave train developing from west to east in the upper troposphere with a wavelength of approximately 90° longitude (Figs. 4a–c). This rapid eastward propagation occurs along the STJ waveguide (Fig. S1h). The wave appears to respond to the presence of the
A second more poleward wave train with equivalent-barotropic structure becomes prominent in the second week of the integration (Figs. 4–6d–e). It appears to escape into the extratropics, arcing poleward and eastward to higher latitudes and then reflecting back to lower latitudes, reminiscent of a great circle path. This wave train is substantially established by day 11 (Fig. 4e), and the day 17–30 mean (Fig. 4f) resembles the observed IOD wave train (Fig. 2b).
The RWS reveals the apparent origin of the extratropical height response. The S1 and S2 RWS at day 3 are displayed in Fig. 7a, along with the anomalous divergent wind and OLR. We examine the RWS early in the integration (e.g., Li et al. 2015a) to highlight the direct forcing from the imposed heating instead of the vorticity response that develops as the Rossby wave modifies the divergence and vorticity fields later in the integration (altering S2, in particular, as discussed below). Decreased OLR (shading) in the EIO develops over the imposed heating, which is locally largely balanced by upward motion, as confirmed in Fig. 8 that displays the vertical velocity at 400 hPa early in the integration (negative omega i.e., upward motion is shaded in blue). Due to the presence of the STJ and the associated maximum in
Subsequently by days 3–5, a positive S2 anomaly develops over southern and eastern Australia, which acts to extend eastward the anticyclonic S1 anomaly. This anticyclonic S2 term is associated with upward motion that develops over southern Australia (Fig. 8), which can be deduced to occur from the omega equation (e.g., Hoskins and James 2014). As the wave train disperses eastward, a cyclonic center in the upper troposphere develops over southeastern Australia (Fig. 4). This cyclonic anomaly in the presence of the strong vertical shear of the mean zonal wind in the STJ will induce upward motion in the midtroposphere. Therefore, this S2 anomaly appears to develop in response to strong induced ascent over eastern Australia (Fig. 8b) rather than a direct RWS associated with the imposed heating.
The largest total diabatic heating anomaly (Fig. S2) occurs in the same location as the imposed heating but there is also anomalous negative heating in the equatorial WIO, creating a zonal dipole-like pattern, similar to the pattern expected during the negative IOD. In the zonal mean, the imposed heating causes the tropics and SH subtropics to warm and subsequently appears to drive a local Hadley response that shifts the STJ poleward over the Indian Ocean (not shown). We have conducted an experiment where the zonal-mean heating is prescribed to be zero (i.e., by applying compensating cooling in the latitude band of the heated region but outside the source region) and the response is similar, indicating that this mean-state change does not substantially affect wave propagation.
c. Possible pathways to overcome the undefined region
After examining the initiation of the wave response to imposed EIO diabatic heating and the evolution of the upper-tropospheric response, we now explore possible pathways that could enable the Rossby wave train to overcome the
The first consideration is that the
Second, it is possible that the wave train propagated largely unimpeded at lower levels of the troposphere where
The emergence of an equivalent-barotropic structure in the SH extratropics poleward of the
These ideas are summarized in Fig. 9, which shows the vertical structure of the height anomalies approximately following the ray path by tracing the centers of the anomalies from Figs. 4–6. The subtropical-midlatitude anticyclonic anomaly (shading) at 110°E is primarily confined to the mid- to upper troposphere (changes sign below 500 hPa), with the maximum anomaly between 150 and 200 hPa. The mid–high-latitude cyclonic anomaly that develops at 130°E by days 5 and 7 appears to do so mainly below the
The third possibility is that the wave propagated directly across
We also consider the possibility that the inclusion of the mean meridional wind could allow the Rossby wave to propagate directly across the
We also computed the WAF using the Plumb (1985) definition, which does not include variations in the background meridional wind, and they are nearly identical to the WAF in Fig. 4, thus providing further evidence that the background meridional wind is not important for poleward propagation in our experiments. In this section, we have provided evidence of possible pathways to explain this apparent disagreement between theory and observations and three more possible pathways are described below.
Another possibility is that the extratropical wave train can emerge from the tropics on the upstream side (to the west) of the
The pathway around to the west of the
It is also possible that the wave train itself acts to reduce the
Finally, the possibility proposed by McIntosh and Hendon (2017) is considered: the transient eddy feedback on the 250-hPa flow could act as an effective RWS on the poleward side of the
d. Comparison with a linearized simple model
Some of the ideas addressed above can be assessed by comparison to the linear SGCM. In this model, wave–mean-flow interactions and baroclinic eddy feedbacks do not occur. Figures 12 and 13 show the evolution of the 250- and 550-hPa height and WAF anomalies for the anomalous forcing at 100°E. We only display up to day 11 after which point the high-latitude response evolves very little and is maintained by dissipation. Like the CAM5 experiment, the mid–high-latitude wave train emerges after ∼5 days and has an equivalent-barotropic structure. In contrast to CAM5, in the upper troposphere, the equatorial Kelvin wave propagates faster but the rapid wave train response along the STJ waveguide and the apparent reflection from the
Nonetheless, by day 9 (Fig. 12e), a wave train similar to the observed (Fig. 2b) and simulated CAM5 wave trains (Fig. 4f) has developed. It is also apparent at the 550- (Fig. 13) and 750-hPa (not shown) levels, which closely resemble the CAM5 results (Figs. 5 and 6). Furthermore, the RWS anomalies in the SGCM (Fig. 14) are similar to CAM5 (Fig. 7), demonstrating that the wave train in the SGCM is also generated by the anticyclonic S1 RWS upstream of Australia.
Compared to 250 hPa (Fig. 12), where strong WAF emanates from the tropical Indian Ocean but appears to have no direct pathway across
In terms of a pathway around the western end of the
The SGCM appears to capture the extratropical teleconnection associated with the IOD, suggesting that the transient eddy feedback is not required. However, there is some loss of magnitude in the height anomalies between Australian latitudes and farther south in the SGCM, indicating that the transient eddy feedback may still be important. This can be demonstrated by computing the efficiency of the tropical heating for producing the extratropical response, which we estimate by the magnitude of the cyclonic anomaly south of Australia on day 20 divided by the vertically- and patch-averaged anomalous heating in the eastern Indian Ocean. This ratio is 70 and 52 m K−1 day−1 in CAM5 and the SGCM, respectively. Thus, the nonlinear mode produces an ∼25% bigger extratropical response, which we tentatively attribute to the induction of the transient eddy feedback in the nonlinear model. However, other factors, such as nonlinear advection (e.g., Hendon 1986; Ting and Yu 1998) and the biased western end of the
In summary, the transient eddy feedback appears not to be required to establish the extratropical wave train across the
4. Conclusions and discussion
This study has investigated the dynamics of the wintertime teleconnection from the tropical Indian Ocean into the SH extratropics. Using both a linear model (SGCM) and an AGCM (CAM5), we conducted experiments with an imposed local atmospheric diabatic heating anomaly in the equatorial EIO at 100°E and examined the daily evolution from when the heating is switched on. In this way, we were able to understand the development of the tropically forced Rossby wave and how it can overcome the apparent barrier to stationary wave propagation caused by the strong wintertime STJ, where Ks vanishes on its poleward flank due to
The development of the simulated wave train in the 90-member CAM5 ensemble appears to follow a two-stage process. An initial Rossby wave train in the upper troposphere rapidly develops with primarily zonal propagation and is trapped along the STJ waveguide. There is clear evidence of the
Analysis of the CAM5 simulations indicates that the Rossby wave train forced by imposed heating in the EIO appears to traverse the region of
First, we show evidence of uninhibited propagation of WAFs underneath the
Second, the Rossby wave appears to partly propagate directly through the
Third, the Rossby wave propagates around the upstream region of
The strong coherence between the observational and CAM5 analysis and the consistent results with the SGCM provides evidence that the mechanisms outlined above operate linearly. McIntosh and Hendon (2017) proposed that a feedback from the transient eddies was the key mechanism to allow the stationary Rossby wave to appear on the poleward side of the
Although we have assumed that the waves are stationary, our results should still be valid if we had assumed nonstationary wave propagation (Kω) because the only difference between stationary and nonstationary waves in Eq. (1) is in the denominator which becomes (
While there are still some gaps between theory and observations, this study helps bring some of these ideas closer together and has helped refine linear stationary wave theory ideas. Studies have typically focused on horizontal propagation of stationary Rossby waves in the upper troposphere (e.g., Hoskins and Karoly 1981). By resolving vertical changes in the forced anomaly, this study provides evidence that the wave can penetrate under the
Finally, our results demonstrate that teleconnections strongly depend on the mean state. The mean state can be defined in different ways, typically using a seasonal mean, and we have defined it on a daily scale using an ensemble mean. Ks is sensitive to the record length for the mean winds. When high temporal resolution data are used, there may be instances when wave propagation is possible in a region where it is usually prohibited (e.g., Rudeva and Simmonds 2021). This leaky behavior of the
There are several possible extensions to this study. Future work could isolate the key pathway, such as by applying a localized sponge relaxation scheme outside the diabatic heating region, which dampens wave propagation in a selected direction (e.g., Shaman and Tziperman 2016). Furthermore, during boreal winter, a similar large region of undefined Ks exists on the poleward flank of the STJ in the west Pacific [e.g., see Fig. 2a in Soulard et al. (2021)]. It would be interesting to explore the dynamics of teleconnections to this region (e.g., Seo and Lee 2017).
Acknowledgments.
We acknowledge support from the Australian Research Council through the Centre of Excellence for Climate Extremes (CLEX; CE170100023). Z.E.G. was supported by an Australian Government Research Training Program Scholarship and a CLEX/Australian Bureau of Meteorology PhD Top-up scholarship. J.M.A. was partially supported by the Regional and Global Model Analysis component of the Earth and Environmental System Modeling Program of the U.S. Department of Energy’s Office of Biological and Environmental Research via National Science Foundation IA 1947282. Computing resources and services were provided by the National Computational Infrastructure, which is supported by the Australian government. The NCAR Command Language (http://www.ncl.ucar.edu) version 6.6.2 was used for data analysis and visualization. Z.E.G. thanks Vishal Dixit for providing the code to implement the heating in CAM5, and Deepashree Dutta for assistance in setting up CESM. The authors thank Steven Sherwood for discussions about experiment design, and Brian Hoskins and Isaac Held for insightful discussions about the results. We also thank Irina Rudeva and Matthew Wheeler for their comments on the first draft.
Data availability statement.
The JRA-55 dataset is available through the Collaborative REAnalysis Technical Environment–Intercomparison Project (CREATE-IT) website (https://esgf-node.llnl.gov/projects/create-ip/). The monthly mean OLR dataset is from the NOAA website (https://psl.noaa.gov/data/gridded/data.olrcdr.interp.html). The merged Hadley-NOAA/Optimal Interpolation SST dataset is available from its website (ftp://ftp.cgd.ucar.edu/archive/SSTICE/). Model output is available upon request.
REFERENCES
Ambrizzi, T., and B. J. Hoskins, 1997: Stationary Rossby-wave propagation in a baroclinic atmosphere. Quart. J. Roy. Meteor. Soc., 123, 919–928, https://doi.org/10.1002/qj.49712354007.
Bao, M., and D. L. Hartmann, 2014: The response to MJO-like forcing in a nonlinear shallow-water model. Geophys. Res. Lett., 41, 1322–1328, https://doi.org/10.1002/2013GL057683.
Barsugli, J., and P. Sardeshmukh, 2002: Global atmospheric sensitivity to tropical SST anomalies throughout the Indo-Pacific basin. J. Climate, 15, 3427–3442, https://doi.org/10.1175/1520-0442(2002)015<3427:GASTTS>2.0.CO;2.
Branstator, G., 1983: Horizontal energy propagation in a barotropic atmosphere with meridional and zonal structure. J. Atmos. Sci., 40, 1689–1708, https://doi.org/10.1175/1520-0469(1983)040<1689:HEPIAB>2.0.CO;2.
Branstator, G., 2014: Long-lived response of the midlatitude circulation and storm tracks to pulses of tropical heating. J. Climate, 27, 8809–8826, https://doi.org/10.1175/JCLI-D-14-00312.1.
Cai, W., T. Cowan, and M. Raupach, 2009: Positive Indian Ocean dipole events precondition southeast Australia bushfires. Geophys. Res. Lett., 36, L19710, https://doi.org/10.1029/2009GL039902.
Cai, W., P. van Rensch, T. Cowan, and H. H. Hendon, 2011: Teleconnection pathways of ENSO and the IOD and the mechanisms for impacts on Australian rainfall. J. Climate, 24, 3910–3923, https://doi.org/10.1175/2011JCLI4129.1.
Ding, Q. H., E. J. Steig, D. S. Battisti, and J. M. Wallace, 2012: Influence of the tropics on the southern annular mode. J. Climate, 25, 6330–6348, https://doi.org/10.1175/JCLI-D-11-00523.1.
Dixit, V., S. Sherwood, O. Geoffroy, and D. Mantsis, 2018: The role of nonlinear drying above the boundary layer in the mid-Holocene African monsoon. J. Climate, 31, 233–249, https://doi.org/10.1175/JCLI-D-17-0234.1.
Duchon, C., 1979: Lanczos filtering in one and two dimensions. J. Appl. Meteor., 18, 1016–1022, https://doi.org/10.1175/1520-0450(1979)018<1016:LFIOAT>2.0.CO;2.
Gill, A., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106, 447–462, https://doi.org/10.1002/qj.49710644905.
Gillett, Z. E., H. H. Hendon, J. M. Arblaster, and E.-P. Lim, 2021: Tropical and extratropical influences on the variability of the Southern Hemisphere wintertime subtropical jet. J. Climate, 34, 4009–4022, https://doi.org/10.1175/JCLI-D-20-0460.1.
Hall, N. M. J., 2000: A simple GCM based on dry dynamics and constant forcing. J. Atmos. Sci., 57, 1557–1572, https://doi.org/10.1175/1520-0469(2000)057<1557:ASGBOD>2.0.CO;2.
Held, I. M., R. L. Panetta, and R. T. Pierrehumbert, 1985: Stationary external Rossby waves in vertical shear. J. Atmos. Sci., 42, 865–883, https://doi.org/10.1175/1520-0469(1985)042<0865:SERWIV>2.0.CO;2.
Hendon, H. H., 1986: The time-mean flow and variability in a nonlinear model of the atmosphere with tropical diabatic forcing. J. Atmos. Sci., 43, 72–89, https://doi.org/10.1175/1520-0469(1986)043<0072:TTMFAV>2.0.CO;2.
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, 1179–1196, https://doi.org/10.1175/1520-0469(1981)038<1179:TSLROA>2.0.CO;2.
Hoskins, B. J., and T. Ambrizzi, 1993: Rossby wave propagation on a realistic longitudinally varying flow. J. Atmos. Sci., 50, 1661–1671, https://doi.org/10.1175/1520-0469(1993)050<1661:RWPOAR>2.0.CO;2.
Hoskins, B. J., and I. N. James, 2014: The omega equation. Fluid Dynamics of the Midlatitude Atmosphere, B. J. Hoskins and I. N. James, Eds., John Wiley & Sons, 225–243.
Hoskins, B. J., I. N. James, and G. H. White, 1983: The shape, propagation and mean-flow interaction of large-scale weather systems. J. Atmos. Sci., 40, 1595–1612, https://doi.org/10.1175/1520-0469(1983)040<1595:TSPAMF>2.0.CO;2.
Hurrell, J. W., J. J. Hack, D. Shea, J. M. Caron, and J. Rosinski, 2008: A new sea surface temperature and sea ice boundary dataset for the Community Atmosphere Model. J. Climate, 21, 5145–5153, https://doi.org/10.1175/2008JCLI2292.1.
Hurrell, J. W., and Coauthors, 2013: The Community Earth System Model: A framework for collaborative research. Bull. Amer. Meteor. Soc., 94, 1339–1360, https://doi.org/10.1175/BAMS-D-12-00121.1.
Jin, F. F., and B. J. Hoskins, 1995: The direct response to tropical heating in a baroclinic atmosphere. J. Atmos. Sci., 52, 307–319, https://doi.org/10.1175/1520-0469(1995)052<0307:TDRTTH>2.0.CO;2.
Karoly, D. J., 1989: Southern Hemisphere circulation features associated with El Niño–Southern Oscillation events. J. Climate, 2, 1239–1252, https://doi.org/10.1175/1520-0442(1989)002<1239:SHCFAW>2.0.CO;2.
Kay, J. E., and Coauthors, 2015: The Community Earth System Model (CESM) large ensemble project: A community resource for studying climate change in the presence of internal climate variability. Bull. Amer. Meteor. Soc., 96, 1333–1349, https://doi.org/10.1175/BAMS-D-13-00255.1.
King, A. D., A. J. Pitman, B. J. Henley, A. M. Ukkola, and J. R. Brown, 2020: The role of climate variability in Australian drought. Nat. Climate Change, 10, 177–179, https://doi.org/10.1038/s41558-020-0718-z.
Kobayashi, S., and Coauthors, 2015: The JRA-55 reanalysis: General specifications and basic characteristics. J. Meteor. Soc. Japan, 93, 5–48, https://doi.org/10.2151/jmsj.2015-001.
Li, X., E. P. Gerber, D. M. Holland, and C. Yoo, 2015a: A Rossby wave bridge from the tropical Atlantic to West Antarctica. J. Climate, 28, 2256–2273, https://doi.org/10.1175/JCLI-D-14-00450.1.
Li, X., D. M. Holland, E. P. Gerber, and C. Yoo, 2015b: Rossby waves mediate impacts of tropical oceans on West Antarctic atmospheric circulation in austral winter. J. Climate, 28, 8151–8164, https://doi.org/10.1175/JCLI-D-15-0113.1.
Li, Y., J. Li, F. F. Jin, and S. Zhao, 2015c: Interhemispheric propagation of stationary Rossby waves in a horizontally nonuniform background flow. J. Atmos. Sci., 72, 3233–3256, https://doi.org/10.1175/JAS-D-14-0239.1.
Liebmann, B., and C. Smith, 1996: Description of a complete (interpolated) outgoing longwave radiation dataset. Bull. Amer. Meteor. Soc., 77, 1275–1277.
Lim, E.-P., and Coauthors, 2020: The 2019 Antarctic sudden stratospheric warming. SPARC Newsletter, No. 54, 10–13, https://www.sparc-climate.org/wp-content/uploads/sites/5/2017/12/SPARCnewsletter_Jan2020_WEB.pdf.
Lim, H., and C. Chang, 1986: Generation of internal- and external-mode motions from internal heating: Effects of vertical shear and damping. J. Atmos. Sci., 43, 948–957, https://doi.org/10.1175/1520-0469(1986)043<0948:GOIAEM>2.0.CO;2.
Lin, H., and G. Brunet, 2018: Extratropical response to the MJO: Nonlinearity and sensitivity to the initial state. J. Atmos. Sci., 75, 219–234, https://doi.org/10.1175/JAS-D-17-0189.1.
Lin, H., G. Brunet, and R. Mo, 2010: Impact of the Madden–Julian Oscillation on wintertime precipitation in Canada. Mon. Wea. Rev., 138, 3822–2839, https://doi.org/10.1175/2010MWR3363.1.
Lindzen, R. S., and K. K. Tung, 1978: Wave overreflection and shear instability. J. Atmos. Sci., 35, 1626–1632, https://doi.org/10.1175/1520-0469(1978)035<1626:WOASI>2.0.CO;2.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44, 25–43, https://doi.org/10.2151/jmsj1965.44.1_25.
McIntosh, P. C., and H. H. Hendon, 2017: Understanding Rossby wave trains forced by the Indian Ocean Dipole. Climate Dyn., 50, 2783–2798, https://doi.org/10.1007/s00382-017-3771-1.
McKay, R. C., J. M. Arblaster, P. Hope, and E.-P. Lim, 2021: Exploring atmospheric circulation leading to three anomalous Australian spring heat events. Climate Dyn., 56, 2181–2198, https://doi.org/10.1007/s00382-020-05580-0.
Meehl, G. A., J. M. Arblaster, G. Branstator, and H. van Loon, 2008: A coupled air–sea response mechanism to solar forcing in the Pacific region. J. Climate, 21, 2883–2897, https://doi.org/10.1175/2007JCLI1776.1.
Meehl, G. A., J. M. Arblaster, C. M. Bitz, C. T. Y. Chung, and H. Teng, 2016: Antarctic sea-ice expansion between 2000 and 2014 driven by tropical Pacific decadal climate variability. Nat. Geosci., 9, 590–595, https://doi.org/10.1038/ngeo2751.
Meehl, G. A., J. M. Arblaster, C. T. Y. Chung, M. M. Holland, A. DuVivier, L. Thompson, D. Yang, and C. M. Bitz, 2019: Sustained ocean changes contributed to sudden Antarctic sea ice retreat in late 2016. Nat. Commun., 10, 14, https://doi.org/10.1038/s41467-018-07865-9.
Neale, R. B., and Coauthors, 2012: Description of the NCAR Community Atmosphere Model (CAM 5.0). NCAR Tech. Note NCAR/TN-486+STR, 274 pp., www.cesm.ucar.edu/models/cesm1.0/cam/docs/description/cam5_desc.pdf.
Nie, Y., Y. Zhang, X. Yang, and H. Ren, 2019: Winter and summer Rossby wave sources in the CMIP5 models. Earth Space Sci., 6, 1831–1846, https://doi.org/10.1029/2019EA000674.
O’Kane, T. J., D. P. Monselesan, and J. S. Risbey, 2017: A multiscale reexamination of the Pacific–South American pattern. Mon. Wea. Rev., 145, 379–402, https://doi.org/10.1175/MWR-D-16-0291.1.
Oleson, K. W., and Coauthors, 2010: Technical description of version 4.0 of the Community Land Model (CLM). NCAR Tech. Note NCAR/TN-478+STR, 257 pp., https://doi.org/10.5065/D6FB50WZ.
Plumb, R. A., 1985: On the three-dimensional propagation of stationary waves. J. Atmos. Sci., 42, 217–229, https://doi.org/10.1175/1520-0469(1985)042<0217:OTTDPO>2.0.CO;2.
Purich, A., and M. H. England, 2019: Tropical teleconnections to Antarctic Sea ice during austral spring 2016 in coupled pacemaker experiments. Geophys. Res. Lett., 46, 6848–6858, https://doi.org/10.1029/2019GL082671.
Qin, J., and W. Robinson, 1993: On the Rossby wave source and the steady linear response to tropical forcing. J. Atmos. Sci., 50, 1819–1823, https://doi.org/10.1175/1520-0469(1993)050<1819:OTRWSA>2.0.CO;2.
Rao, J., C. I. Garfinkel, I. P. White, and C. Schwartz, 2020: The Southern Hemisphere minor sudden stratospheric warming in September 2019 and its predictions in S2S models. J. Geophys. Res. Atmos., 125, e2020JD032723, https://doi.org/10.1029/2020JD032723.
Rudeva, I., and I. Simmonds, 2021: Midlatitude winter extreme temperature events and connections with Anomalies in the Arctic and tropics. J. Climate, 34, 3733–3749, https://doi.org/10.1175/JCLI-D-20-0371.1.
Saji, N. H., B. N. Goswami, P. N. Vinayachandran, and T. Yamagata, 1999: A dipole mode in the tropical Indian Ocean. Nature, 401, 360–363, https://doi.org/10.1038/43854.
Sardeshmukh, P. D., and B. J. Hoskins, 1984: Spatial smoothing on the sphere. Mon. Wea. Rev., 112, 2524–2529, https://doi.org/10.1175/1520-0493(1984)112<2524:SSOTS>2.0.CO;2.
Sardeshmukh, P. D., and B. J. Hoskins, 1985: Vorticity balances in the tropics during the 1982‐83 El Niño‐Southern oscillation event. Quart. J. Roy. Meteor. Soc., 111, 261–278, https://doi.org/10.1256/smsqj.46801.
Schneider, E. K., and I. G. Watterson, 1984: Stationary Rossby wave propagation through easterly layers. J. Atmos. Sci., 41, 2069–2083, https://doi.org/10.1175/1520-0469(1984)041<2069:SRWPTE>2.0.CO;2.
Seo, K.-H., and H.-J. Lee, 2017: Mechanisms for a PNA-like teleconnection pattern in response to the MJO. J. Atmos. Sci., 74, 1767–1781, https://doi.org/10.1175/JAS-D-16-0343.1.
Shaman, J., and E. Tziperman, 2016: The superposition of eastward and westward Rossby waves in response to localized forcing. J. Climate, 29, 7547–7557, https://doi.org/10.1175/JCLI-D-16-0119.1.
Simpson, I. R., and Coauthors, 2020: An evaluation of the large-scale atmospheric circulation and its variability in CESM2 and other CMIP models. J. Geophys. Res. Atmos., 125, e2020JD032835, https://doi.org/10.1029/2020JD032835.
Soulard, N., H. Lin, J. Derome, and B. Yu, 2021: Tropical forcing of the circumglobal teleconnection pattern in boreal winter. Climate Dyn., 57, 865–877, https://doi.org/10.1007/s00382-021-05744-6.
Takaya, K., and H. Nakamura, 2001: A formulation of a phase-independent wave-activity flux for stationary and migratory quasigeostrophic eddies on a zonally varying basic flow. J. Atmos. Sci., 58, 608–627, https://doi.org/10.1175/1520-0469(2001)058<0608:AFOAPI>2.0.CO;2.
Ting, M., 1996: Steady linear response to tropical heating in barotropic and baroclinic models. J. Atmos. Sci., 53, 1698–1709, https://doi.org/10.1175/1520-0469(1996)053<1698:SLRTTH>2.0.CO;2.
Ting, M., and L. Yu, 1998: Steady response to tropical heating in wavy linear and nonlinear baroclinic models. J. Atmos. Sci., 55, 3565–3582, https://doi.org/10.1175/1520-0469(1998)055<3565:SRTTHI>2.0.CO;2.
Trenberth, K. E., Y. Zhang, and J. T. Fasullo, 2015: Relationships among top-of-atmosphere radiation and atmospheric state variables in observations and CESM. J. Geophys. Res. Atmos., 120, 10 074–10 090, https://doi.org/10.1002/2015JD023381.
Ummenhofer, C. C., M. H. England, P. C. McIntosh, G. A. Meyers, M. J. Pook, J. S. Risbey, A. S. Gupta, and A. S. Taschetto, 2009: What causes southeast Australia’s worst droughts? Geophys. Res. Lett., 36, L04706, https://doi.org/10.1029/2008GL036801.
Wang, G., H. H. Hendon, J. M. Arblaster, E.-P. Lim, S. Abhik, and P. van Rensch, 2019: Compounding tropical and stratospheric forcing of the record low Antarctic sea-ice in 2016. Nat. Commun., 10, 13, https://doi.org/10.1038/s41467-018-07689-7.
Wheeler, M. C., and H. Nguyen, 2015: Tropical meteorology and climate: Equatorial waves. Encyclopedia of Atmospheric Sciences, 2nd ed. G. R. North, J. Pyle, and F. Zhang, Eds., Academic Press, 102–112.
Wheeler, M. C., G. N. Kiladis, and P. J. Webster, 2000: Large-scale dynamical fields associated with convectively coupled equatorial waves. J. Atmos. Sci., 57, 613–640, https://doi.org/10.1175/1520-0469(2000)057<0613:LSDFAW>2.0.CO;2.
Wirth, V., 2020: Waveguidability of idealized midlatitude jets and the limitations of ray tracing theory. Wea. Climate Dyn., 1, 111–125, https://doi.org/10.5194/wcd-1-111-2020.
Yang, G.-Y., and B. J. Hoskins, 1996: Propagation of Rossby waves of nonzero frequency. J. Atmos. Sci., 53, 2365–2378, https://doi.org/10.1175/1520-0469(1996)053<2365:PORWON>2.0.CO;2.
Yang, G.-Y., and B. J. Hoskins, 2017: The equivalent barotropic structure of waves in the tropical atmosphere in the Western Hemisphere. J. Atmos. Sci., 74, 1689–1704, https://doi.org/10.1175/JAS-D-16-0267.1.
Yiu, Y. Y. S., and A. C. Maycock, 2019: On the seasonality of the El Niño teleconnection to the Amundsen Sea region. J. Climate, 32, 4829–4845, https://doi.org/10.1175/JCLI-D-18-0813.1.
Zhao, M., and H. H. Hendon, 2009: Representation and prediction of the Indian Ocean dipole in the POAMA seasonal forecast model. Quart. J. Roy. Meteor. Soc., 135, 337–352, https://doi.org/10.1002/qj.370.
Zhao, S., J. Li, and Y. Li, 2015: Dynamics of an interhemispheric teleconnection across the critical latitude through a southerly duct during boreal winter. J. Climate, 28, 7437–7456, https://doi.org/10.1175/JCLI-D-14-00425.1.
Zhao, S., J. Li, Y. Li, F. Jin, and J. Zheng, 2019: Interhemispheric influence of Indo-Pacific convection oscillation on Southern Hemisphere rainfall through southward propagation of Rossby waves. Climate Dyn., 52, 3203–3221, https://doi.org/10.1007/s00382-018-4324-y.