1. Introduction
a. Background
During El Niño–Southern Oscillation (ENSO) winters an equivalent barotropic stationary wave train pattern dominates over Pacific and North American region (Horel and Wallace 1981) and this pattern can be interpreted as an external Rossby wave forced by equatorial Pacific heating (Hoskins and Karoly 1981; Held and Kang 1987; Trenberth et al. 1998). Rossby wave generation and propagation provides the basis for many theories of how the tropics influence midlatitudes (Hoskins et al. 1977; Held and Kang 1987; Branstator 1983). A process-oriented diagnostic (POD; Maloney et al. 2019; Annamalai 2020) package that addresses the chain of processes, that is, intermediate between equatorial Pacific heating and the circulation pattern over the Pacific and North American regions that are usually not addressed in model evaluations (Deser et al. 2016), is developed. The POD will help to address this critical model evaluation gap, by quantifying the roles of anomalous upper-tropospheric divergent wind patterns in the generation of stationary Rossby waves, and the zonally and meridionally varying ambient flow properties that determine the horizontal propagation of these waves (e.g., Branstator 1985; Ting and Sardeshmukh 1993; Di Carlo et al. 2022). As part of model evaluation, we apply the POD to a subset of climate models that participated in phases 5 and 6 of the Atmospheric General Circulation Model Intercomparison Project (AMIP). Besides assessing the representation of relevant processes in individual models, improvements or degradations in models that participated in both AMIP5 to AMIP6 versions are identified, and their implications to ENSO-induced teleconnections are discussed.
During El Niño winters, positive sea surface temperature (SST) anomalies along the central-eastern equatorial Pacific (e.g., Rasmusson and Carpenter 1982) favor local enhancement of moist static energy, promoting deep convection and increasing the release of latent heating throughout the troposphere (see Wallace et al. 1998 for a review). The resultant anomalous vertical velocity and upper-level divergence alters the generation of the horizontal component of atmospheric vorticity (Hoskins et al. 1977; Hoskins and Karoly 1981), giving rise to anomalous Rossby wave sources (RWS′; Sardeshmukh and Hoskins 1988). These RWS′ excite normal modes of the zonally varying ambient flow (Simmons et al. 1983). Kang and Held (1986) and Sardeshmukh and Hoskins (1988) emphasized that at upper-tropospheric levels, the contribution by advection of the climatological meridional gradient in absolute vorticity by the anomalous divergent winds can also lead to a significant RWS′ in westerly wind regions, in which Rossby wave motion is possible. Trenberth et al. (1998) provide a comprehensive review on all aspects of tropical–extratropical interactions. Atmospheric general circulation models (AGCMs) forced with ENSO-related SST anomalies exhibit various levels of success in representing this circulation pattern (Shukla and Wallace 1983; Lau and Nath 1994; Hoerling et al. 1997; Kumar and Hoerling 1997; Hoerling and Kumar 2002; Shukla et al. 2000; Barsugli and Sardeshmukh 2002; Peng and Kumar 2005; Annamalai et al. 2007; Peng et al. 2014; Deser et al. 2016; Li et al. 2020, among others).
During boreal winters, a prominent midlatitude circulation over the Pacific–North American region is the so-called Pacific–North American (PNA) pattern defined by Wallace and Gutzler (1981). One established interpretation is that PNA is an internal mode of the midlatitude circulation (Lau 1981) that is amplified by ENSO (Horel and Wallace 1981; Kumar and Hoerling 1995; Lau and Nath 1994; Hoerling et al. 1997; Hoerling and Kumar 2002; Peng and Kumar 2005; Li et al. 2020), that is, ENSO can create preferred teleconnection response patterns, such as the PNA (Trenberth et al. 1998). An alternative interpretation is based on differences in the position and structures of the upper-level height anomalies, particularly during El Niño (e.g., Lopez and Kirtman 2019). Based on this paradigm, Straus and Shukla (2002) suggest that ENSO does not force PNA. Intriguingly, the height anomalies bear close similarity to traditional PNA during La Niña winters (Fig. 1 in Lopez and Kirtman 2019). In the following subsection, we clarify our approach with reference to the above interpretations.
b. Present study
Guided by linear and nonlinear model results available in the literature, we hypothesize that the efficacy of climate models’ ENSO-induced teleconnection can be assessed by their fidelity in representing anomalous tropical precipitation and associated upper-level divergence, RWS′, and ambient flow characteristics. To test this hypothesis, we have developed a diagnostics package (section 2). Besides assessing climatological basic flow properties and seasonal anomalous conditions during ENSO winters, the barotropic vorticity equation is solved at mid-to-upper-troposphere levels, and terms contributing to the total RWS′ are quantified. Here, to circumvent coupled models’ simulated SST biases cascading into anomalous precipitation and RWS′ characteristics, we focus only on the AMIP-type experiments in which all the participating models are forced by observed SSTs (Gates et al. 1992). This limits the attribution of errors solely to limitations in the representation of atmospheric model processes and their interdependencies. However, the presence of error compensation can make such direct attribution challenging.
The POD is applied to five reanalysis products to assess inherent uncertainties among them, and for the purposes of model assessment, robustness in their results is sought. The reason being that in data sparse regions such as the deep tropics, estimates of diabatic processes are not well constrained by in situ measurements. Then, they depend on the first guess provided by the forecast models that is sensitive to the physical parameterizations employed (Annamalai et al. 1999; Zhang et al. 2008).
Figure 1a shows El Niño winter (DJF) composites of geopotential height anomalies at 200 hPa (HGT200) constructed from ERA5 reanalysis (details are given in section 2). One notes alternating signs in HGT200 that arch from the tropical Pacific poleward to the North Pacific, and then eastward and southward across North America exhibiting a wave train pattern. One clear difference from the traditional PNA pattern of Wallace and Gutzler (1981) is the longitudinal shift of centers (∼10°) over North Pacific and North America (boxes 2–4 in Fig. 1a), in agreement with Lopez and Kirtman (2019). Guided by Fig. 1a and following the Wallace and Gutzler (1981) approach, a circulation index termed El Niño–induced circulation index (ENCI) over the Pacific–North America is calculated using standardized HGT200 as 0.25 × (HGT200REG1 − HGT200REG2 + HGT200REG3 − HGT200REG4), and the results are shown for the period 1979–2019 (Fig. 1b). For boreal winters, the traditional PNA index obtained from the Climate Prediction Center (CPC) is also shown (Fig. 1c). Both indices, barring differences in magnitude, show positive (negative) values during warm (cold) phases of ENSO, and the cross-correlation between them is 0.8. Our goal here is not to examine how accurately El Niño forces PNA. Our intention is to assess the model processes that are deemed necessary for the forced midlatitude circulation over the Pacific–North American regions during El Niño winters.
In this model evaluation study, we assess the models’ fidelity in accurately representing (i) strength and location of RWS′ that depend on anomalous tropical precipitation and associated divergent flow properties, and climatological vorticity gradients in the subtropical jets, and (ii) local vorticity gradient of the ambient zonal flow (
The remainder of the paper is organized as follows: Section 2 deals with the POD and some details of the AMIP simulations and reanalysis products. In section 3, models’ ability in representing equatorial Pacific precipitation (EPP) anomalies to generation of RWS′ are discussed. Section 4 deals with the models’ fidelity in simulating basic-state properties. El Niño–induced teleconnections are interpreted in section 5. In section 6, model biases are quantified, along with improvements or degradations in the transitions from AMIP5 to AMIP6 are shown, and POD’s efficiency to model developers along with key implications are presented. Conclusions and future directions are highlighted in section 7. Further discussions on model interpretations on midlatitude circulation are summarized in appendix.
2. ENSO Rossby wave source POD, model solutions and reanalysis products
a. ENSO_RWS
The diagnostics package termed ENSO_RWS consists of four levels or steps that are sequentially performed with monthly data either from reanalysis or model integrations. To attain robust composite results a reasonable sample of ENSO winters is needed. However, the POD can be applied even for a single El Niño winter (e.g., seasonal prediction solutions). Similarly, the POD is applicable to any number of pressure levels (e.g., to identify the level at which maximum upper-level divergence and associated RWS′ are located).
In level 4, results from levels 1–3 are condensed into scatterplots. Specifically, the sequential plots illustrate the model’s ability in representing the chain of processes. The POD is expected to inform the sensitivity of RWS′ to EPP anomalies and ambient flow characteristics and the importance of
b. Model solutions and reanalysis
The AMIP5/6 suite consists of solutions from a total of 55 models, and the simulation period is 1979–2005 for AMIP5, and 1979–2014 for AMIP6, respectively. Compared to AMIP5, there are additional models in AMIP6 pool (Eyring et al. 2016). The POD is applied to solutions available from all AMIP5/6 models, and results are summarized for AMIP5 and AMIP6 separately as well as for the models that participated in both phases.
To validate model results, similar diagnostics are applied to multiple reanalysis products for the period 1979–2014. We emphasize that reanalysis uncertainties need to be taken into account while validating models. The products we diagnose include the European Centre for Medium-Range Weather Forecasts (ECMWF) interim reanalysis (ERAi; Dee et al. 2011); the fifth major global reanalysis produced by ECMWF (ERA5; Herbasch et al. 2020); the Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2; Gelaro et al. 2017); the Japanese 55-year Reanalysis (JRA-55; Kobayashi et al. 2015); and the Climate Forecast System reanalysis (CFS-R; Saha et al. 2010).
For models and reanalysis, variables diagnosed include precipitation, three-dimensional winds and geopotential height. All the diagnostics are performed for composite El Niño winters [years when Niño-3.4 (5°S–5°N, 120°–170°W) SST anomalies exceed 1.0 standard deviation]. Since observed SST is prescribed in AMIP simulations, El Niño winters are identical in both reanalysis and models, and thus offer direct comparisons. Due to space constraints, however, spatial plots from only two reanalysis, CFS-R and ERAi, are shown, since in most of the diagnostics presented here, results from these two reanalyses show lower/upper bounds on the five reanalyses considered.
3. Anomalous Rossby wave sources during El Niño winters
First, we assess anomalous tropical precipitation and associated upper-level circulation since model biases in them (e.g., Kumar et al. 2005) modulate the generation of the horizontal component of atmospheric vorticity giving rise to RWS′ (e.g., Jin and Hoskins 1995; Ting and Sardeshmukh 1993). Second, RWS′ spatial patterns in the subtropical North Pacific are discussed. To condense all the results from all the models, scatterplots are shown. In them, vertical and horizontal dashed lines correspond to reanalysis uncertainties measured as lower and upper bound values among them (e.g., Figs. 4, 7 and 13), and a model simulation is interpreted to be realistic if its values fall within these uncertainties. Discussions on the quantification of model biases, and improvements/degradations from AMIP5 to AMIP6 are deferred to section 6.
a. El Niño–related precipitation and circulation anomalies at 200 hPa
Figures 2a–e show anomalous precipitation (shaded), 200-hPa divergence/convergence (contours/hatching) and divergent winds (vector) from reanalyses. Observations (Fig. 2f) show that in response to warm SST anomalies (contours) along the central and eastern equatorial Pacific, precipitation (shaded) is increased with a local maximum around the date line (10°S–0°, 160°E–140°W), termed the EPP region. Besides differences in intensity, all the reanalyses represent the EPP pattern. In response to increased convection, locally concentrated anomalous upper-level divergence and divergent winds are seen. Poleward, the meridional component of the divergent wind anomalies converges over the subtropical North Pacific (STNP; 25°–40°N, 150°E–160°W) and along the South Pacific convergence zone, implying a strengthened local overturning Hadley circulation. In the equatorial plane, zonal weakening of the Walker circulation with enhanced precipitation over the EPP and suppressed precipitation over the Maritime Continent–tropical western Pacific (MC-TWP), resulting in convergence anomalies over MC-TWP are also apparent. Thus, two well-separated upper-level convergence anomalies, one over the STNP and another over the MC-TWP are evident. In the reanalyses, while the structures of divergent wind anomalies are in good agreement, their intensities differ, which is consistent with the varying precipitation anomalies. Note that a direct comparison of divergent winds in reanalyses will depend on the first guess supplied by the forecast models (Dee et al. 2011). This is sensitive to the physical parameterizations used in each reanalysis model (Annamalai et al. 1999).
Figure 3 shows results from select AMIP5 (left) and their corresponding AMIP6 (right) models while Fig. 4 summarizes all AMIP models’ fidelity in capturing anomalous precipitation and STNP convergence. Compared to observations (Fig. 2f), both in GFDL-AM3 (Fig. 3a) and CAM5 (Fig. 3c), EPP anomalies are meridionally confined but zonally elongated, extending well into the western equatorial Pacific (green vertical lines) with implications for westward shifts in anomalous convergence in the STNP. Also, suppressed precipitation and convergence anomalies protrude into the eastern equatorial Indian Ocean. In contrast, EPP anomalies in MIROC5 and MPI-ESM-LR are stronger than observations (Fig. 4c), but STNP convergence anomalies are stronger in MPI-ESM-LR (Fig. 4a) while they are unorganized with multiple local maxima in MIROC5 (Fig. 3e). In GFDL-AM4 (Fig. 3b) and CAM6 (Fig. 3d), notable improvements include a reduced westward extension of EPP anomalies and spatially organized convergence anomalies in the STNP. Based on EPP and MC-TWP precipitation anomalies (Figs. 3b,d and 4d), the anomalous Walker circulation is realistic in GFDL-AM4 while both precipitation anomalies are weaker than observations in CAM6. The reduced precipitation anomalies protruding into the eastern equatorial Indian Ocean still persist from CAM5, and a narrow tongue of unrealistic convergence anomalies extends along the subtropical Asian jet (around 30°N). This is a feature present in all CAM6 model versions including Finite Volume 2 degree resolution (FV2), WACCM, and WACCM-FV2 (not shown). Compared to MIROC5 (Fig. 3e), considerable improvements in MIROC6 (Fig. 3f) include STNP convergence and positive precipitation anomalies along the East Asian monsoon (EAM; 20°–30°N, 100°–140°E). Of all the AMIP5 models, anomalous EPP and STNP convergence are by far the strongest in MPI-ESM-LR (Fig. 4a), while they are weaker in its AMIP6 version (Fig. 4b).
Scatterplots (Fig. 4) summarize models’ abilities in representing anomalous local Hadley (Figs. 4a,b) and planetary Walker (Figs. 4c,d) circulations. Compared to reanalysis, simulated anomalous tropical precipitation and associated upper-level circulation in most models are weaker. The intermodel spreads between EPP (∼2.5 to 8.0 mm day−1) and STNP convergence (0 to −2.2 × 10−6 s−1), as well as a varied convergence response for similar EPP forcing and vice versa are readily apparent (Figs. 4a,b). The strong association (both in sign and amplitude) between the EPP wetness and MC-TWP dryness (Figs. 4c,d) can be interpreted as follows: The anomalous Walker circulation with ascent over the equatorial central-eastern Pacific leads to descent and low-level divergence over the MC-TWP. Subsequently, the low-level anomalous westerlies advect climatological moisture from the equatorial west to central Pacific further feeding the convection and enhancing ascent (Annamalai 2020). Figures 4e and 4f summarizes the linkage between MC-TWP dryness and regional divergence anomalies along the EAM front. The low-level anomalous anticyclonic circulation that develops in response to reduced MC-TWP precipitation, through horizontal moist advection, leads to a strengthened EAM and, hence, upper-level divergence (e.g., Annamalai et al. 2005). However, large scatter between MC-TWP dryness and EAM-divergence may be attributed to models’ inability in representing latitudinally confined EAM anomalous precipitation (Figs. 2 and 3). Within reanalysis uncertainties, it must be emphasized that GFDL-AM4 captures realistic simulations of both anomalous EPP and STNP convergence (Fig. 4b), and MC-TWP dryness (Fig. 4d). As will be shown next, both convergence over the STNP and divergent wind anomalies along the EAM contribute to
b. Anomalous Rossby wave sources at 200 hPa
Figures 5 and 6 show the geographical distributions of RWS′ (shading) from two reanalysis (Figs. 5a and 6a) and a selection of AMIP5 (Figs. 5b–e) and AMIP6 (Figs. 6b–e) models. In the right panels, 200-hPa climatological
In GFDL-AM3, the STNP maxima in
To assess the role of simulated biases in anomalous precipitation, divergence, and climatological
In summary, the POD identifies model STNP RWS′ deficiencies as being due to the following:
-
Limitations in
-
Westward extension of EPP anomalies and associated anomalous convergence, accounting for spatially incoherent and zonally elongated RWS′ along the jet.
-
Limitations in representing the structure of
4. Ambient flow properties at 200 hPa
In this section, we assess models’ fidelity in accurately representing climatological flow characteristics since model biases in them influence the propagation of stationary Rossby waves, and subsequently influence ENSO-induced teleconnections (e.g., Branstator 1985; Palmer and Mansfield 1986; Ting and Sardeshmukh 1993). Figure 8 shows 200-hPa climatological flow properties of
To illustrate the challenge of choosing a reanalysis for model validation of high-order dynamical flow properties, GFDL-AM3 and GFDL-AM4 are compared in Fig. 9. In GFDL-AM3, the spatial variations in
5. Circulation response during El Niño winters
In models, we test the hypothesis that realistic simulations of primary RWS′ (Figs. 5–7),
a. Circulation response over North Pacific and North America
Figures 10 and 11 show 200-hPa boreal winter climatological distribution of
-
The coherent versus diffuse nature of the primary RWS′ region and its location in the STNP (west or east of the date line or zonally elongated along the jet).
-
Spatial distributions of
Of the AMIP5/6 models, GFDL-AM4 and CAM6 capture the most realistic El Niño–induced circulation anomalies (Figs. 11g,h) with a characteristic meridional propagation from the tropics to North Pacific and then arching northeastwards covering North America, and finally southern parts of the United States of America. Over the north-central Pacific, compared to AMIP5 versions (Figs. 10b,c), there are improvements in representing
In many models, a very common erroneous feature is zonally elongated (northwest–southeast oriented) HGT200 over the Aleutian low region. If the primary RWS′ is positioned to the west of the date line (e.g., MIROC5; Fig. 10d) then owing to the presence of positive values in
In GFDL-AM3 (Fig. 10b), negative values in
b. Metrics to assess models’ circulation performance
We identify two metrics, namely, RWS′ and
6. Model biases, improvements, and implications for model developers
a. Quantification of model biases
Figure 14 shows model biases (expressed in %) with respect to multireanalysis mean anomalies. AMIP5 (AMIP6) model biases are shown in the left (right) panels, and the key variables discussed in previous sections are shown. EPP wetness and MC-TWP dryness, in about two-thirds of all AMIP models, are weaker than observations. Irrespective of a model’s ability to represent equatorial precipitation anomalies, biases in STNP convergence (and associated RWS′) are systematically weaker than in reanalysis. This implies a role for other processes in determining actual STNP convergence. In some models, the maximum level of anomalous divergence/convergence may not be at 200 hPa, an issue currently being investigated. As regards to
b. Improvements or degradations from AMIP5 to AMIP6
Model results of the chain of processes presented above suggest that some of them are improved in AMIP6 and some are degraded. To highlight improvements for each modeling center, Fig. 15 shows results from models that participated in both AMIP versions. Here, biases are estimated with respect to their corresponding multireanalysis means. If the AMIP6 bias falls closer to reanalyses (boxed area in Fig. 15) compared to AMIP5 biases, it is interpreted as improved (left panels). Conversely, if the bias falls more distant, it is interpreted as degraded (right panels). In all panels, number 5 stands for AMIP5 and 6 refers to AMIP6, and colors correspond to the model’s name. If only one of the two variables in Fig. 15 is either improved or degraded, that model’s results are not shown.
AMIP5 model values that lie in the bottom left quadrants (weak in both variables; Figs. 15a,c), and more eastward values of
In considering
c. Implications for model developers
In models, both in basic-state and anomalous conditions, changes in the response of moist processes either to parameterization modifications or tuning and calibration, can often change the nature of the seasonal distributions of tropical precipitation and heating, and by association moistening and divergence profiles. While validation of an integrated scalar quantity such as precipitation is mostly straightforward, an understanding of the circulation consequences in the tropics, and particularly the nonlinear wave–mean flow interactions of the extratropics is not. The intention would be to monitor simulation improvements over time. Equally, however, when inadvertent or unintended degradations are discovered, decisions need to be made on how to move forward. This could involve either a concerted effort to understand the processes leading to the degradation or accepting the response and moving forward regardless. By having a suite of PODs as an automated tool, a broad range of simulation features could be invoked that span the range of performance features for which a model is being designed. In the case of climate models this would include large-scale circulation features (the focus here), modes of variability and climate feedback processes (e.g., cloud feedbacks, precipitation, and temperatures). Thus, the assessment presented here is intended to provide useful feedback to model developers where POD performance changes can be linked to parameterization improvements, even though these linkages can often be circuitous and difficult to disentangle.
Model developers primarily focus on a very limited set of requirements or metrics in order to produce a useful model, often for CMIP (and also AMIP) purposes (Schmidt et al. 2017; Hourdin et al. 2017). The focus is often biased toward the larger climate change requirements, in order to capture historical, and therefore reliable, future evolution of a warming world, even though there is disagreement in priorities across the modeling centers. The analysis presented here is different, in that it captures a succession of processes critical to both climate change and Earth system prediction in the Northern Hemisphere, but it does so with a process-oriented focus that follows this succession from the tropical Pacific to the West Coast of the United States. Applying this particular POD continually through the development process will keep a regular check on the fidelity of the modeled response to anomalous El Niño convection (such a progression in coupled models is obviously a much greater challenge given the variation in Pacific SSTs and associated model biases). Applying a set of PODs is expected to reduce the frequent occurrence in model development of a particular process or mode of variability degrading over time. This degradation can be unknown to the developers earlier due to lack of availability of appropriate PODs. The hope is that it will become a standard method to assess models, in the same way that ENSO, the North Atlantic Oscillation (Simpson et al. 2020), and Madden–Julian oscillation (Chen et al. 2022) have become to date.
7. Conclusions
During boreal winters, to assess climate models’ fidelity in representing ENSO-induced teleconnection over North Pacific and North America, as part of NOAA Model Diagnostics Task Force efforts, a process-oriented diagnostic (POD) package termed ENSO_RWS is developed. The POD assesses the chain of processes, that is, intermediate between equatorial Pacific precipitation (EPP) and the midlatitude circulation pattern that are rarely addressed in model development and evaluations. Ambient upper-tropospheric flow properties such as local vorticity gradient of the ambient zonal flow (
Our POD identifies two important metrics, and they are as follows:
-
Spatial coherence in the primary RWS′ in the STNP.
-
Spatial distributions in
Encouragingly, clear improvements in forcing (EPP) and response (STNP convergence) are noted in AMIP6 experiments leading to improvements in the simulation of primary RWS′ (Figs. 5 and 6) with subsequent improvements of
Comparing models from the same modeling groups (Figs. 15 and 16), we note that if both the metrics are improved then those models’ representation of PNA and ENCI are most commonly improved, perhaps for correct reasons based on the interpretations offered here. Even in these models, however, need for improvements on other aspects such as arresting westward extension of enhanced EPP anomalies and associated zonally elongated RWS′ (e.g., CAM6) are brought out by the POD. If only one of the metrics is improved (e.g., RWS′ in MIROC6) and other metrics are worsened (
Based on the lack of a one-to-one relationship between EPP and STNP convergence (Fig. 4), the suggestion is that the initial source of model error could very well lie in the vertical distribution of diabatic processes that largely determine divergent winds. A successive POD to assess vertical processes during ENSO is a logical next step and will be similarly applied to climate models.
Acknowledgments.
The authors acknowledge the support from NOAA-MAPP Award NA18OAR4310279 for developing process-oriented diagnostics. We acknowledge the World Climate Research Programme’s Working Group on Coupled Modelling, which is responsible for AMIP/CMIP, and we thank the climate modeling groups for producing and making available their model output. Authors sincerely acknowledge valuable comments from all the three anonymous reviewers that helped improve the manuscript. Critical comments from reviewer 3 were instrumental to improve the overall presentation of the manuscript. Dr. Neale also acknowledges the support by the U.S. Department of Energy, Office of Science, Office of Biological and Environmental Research (BER), Regional and Global Model Analysis (RGMA) component of the Earth and Environmental System Modeling Program under Award DE-SC0022070, and National Science Foundation (NSF) IA 1947282. This work was also supported by the National Center for Atmospheric Research (NCAR), which is a major facility sponsored by the NSF under Cooperative Agreement 1852977. This is IPRC Publication Number 1591 and SOEST Publication Number 11617.
Data availability statement.
Data sets analyzed in the study include reanalysis data that are existing data products available at the locations cited in the respective references. Furthermore, standard model outputs that are archived as part of the CMIP6 are diagnosed, and the data are available at https://esgf-node.llnl.gov/search/cmip5/ and https://esgf-node.llnl.gov/search/cmip6/.
APPENDIX
Further Model Interpretations
For active model developers, the simulated circulation response in AMIP6 is too strong or weak or realistic, and do the two identified metrics, RWS′ and
While the simulated PNA/ENCI is realistic in some models, it may be for incorrect reasons. As an example, results from ACCESS-ESM1-5 (Figs. A1 and A2; bottom panels) show weakened Asian jet and weaker
For differing RWS′ strength, CanESM5, INM-CM-4-8, and NorCPM1 simulate PNA/ENCI values that are too strong (Figs. 13c,e). Are there common errors across these models? In all of them,
On the other hand, simulated PNA/ENCI is weaker than reanalysis in about 50% of the AMIP6 models (Fig. 13). They are ACCESS-CM2, BCC-CSM2-MR, CESM-FV2, CESM2-WACCM, CESM2-WACCM-FV2, E3SM1.0, FGOALS-f3-L, GISS-E2-1-G, IPSL-CM6A-LR, SAM0-UNICON, MIROC6, MPI-ESM1-2-HR, MRI-ESM2-0, and Tai-ESM-1. This occurs despite both metrics being realistically represented in some of them (e.g., BCC-CSM2-MR and Tai-ESM-1). Common features in all these models include zonally elongated EPP, RWS′ and HGT200 over the North Pacific, radiation and propagation of multiple wave trains and spatially incoherent
In BCC-CSM2-MR, ambient flow properties are realistic but RWS′ is zonally elongated with signatures along the Asian jet and has a pronounced north–south orientation (reaching 55°–60°N) in central-North Pacific. HGT200 patterns are nearly circular over central-North Pacific (instead of elliptical as in reanalysis) and multiple wave trains are seen in the Northern Hemisphere. Annamalai et al. (2007) suggested that HGT200 due to short wavelength Rossby waves trapped by the jets, after exiting the jet region, are out-of-phase with HGT200 in the North Pacific with waves of longer wavelength (or lower Ks) forced by EPP-induced heating, and thus their destructive interference weakens PNA/ENCI. While results presented in Fig. 15 suggest an overall improvement from its AMIP5 version, caution needs to exercised based on the interpretations offered here. In TaiESM-1, EPP anomalies extend across the entire Pacific basin with consequences for zonally elongated primary RWS′ and HGT200 over the Aleutian region. Compared to reanalysis (Fig. 11b), two well-separated wave trains over North Pacific, one in midlatitudes (∼40°N) and another in high latitudes (60°–70°N), are discernible.
Briefly, besides EPP forcing and ambient flow properties on planetary wave propagation, a number of other nonlinear processes are involved in accurately determining the magnitude of extratropical response. They include interactions with transients (e.g., storm-track changes; Held et al. 1989) and interference from midlatitude chaotic internal variability (Lau 1981), and horizontal and vertical distributions of diabatic heating anomalies over the tropical Indian Ocean region (Simmons et al. 1983; Ting and Sardeshmukh 1993; Lau 1997; Barsugli and Sardeshmukh 2002; Annamalai et al. 2007). In AGCMs, besides primary RWS′ in STNP, due to wave propagation new convergence and divergence centers can form leading to secondary RWS′ elsewhere (Figs. 5 and 6). Consequently, radiation of various wave trains with different wavelengths and great circle paths are expected. Therefore, a detailed assessment of all factors impacting PNA/ENCI amplitude is not feasible and is also beyond the scope of the present research.
REFERENCES
Annamalai, H., 2020: ENSO precipitation anomalies along the equatorial Pacific: Moist static energy framework diagnostics. J. Climate, 33, 9103–9127, https://doi.org/10.1175/JCLI-D-19-0374.1.
Annamalai, H., J. M. Slingo, K. R. Sperber, and K. Hodges, 1999: The mean evolution and variability of the Asian summer monsoon: Comparison of ECMWF and NCEP/NCAR reanalyses. Mon. Wea. Rev., 127, 1157–1186, https://doi.org/10.1175/1520-0493(1999)127<1157:TMEAVO>2.0.CO;2.
Annamalai, H., P. Liu, and S.-P. Xie, 2005: Southwest Indian Ocean SST variability: Its local effect and remote influence on Asian monsoons. J. Climate, 18, 4150–4167, https://doi.org/10.1175/JCLI3533.1.
Annamalai, H., H. Okajima, and M. Watanabe, 2007: Possible impact of Indian Ocean SST on the Northern Hemisphere circulation during El Niño. J. Climate, 20, 3164–3189, https://doi.org/10.1175/JCLI4156.1.
Barsugli, J. J., and P. D. 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. W., 1983: Horizontal energy propagation in a baroclinic 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. W., 1985: Analysis of general circulation model sea-surface temperature anomaly experiments using a linear model. Part I: Forced solutions. J. Atmos. Sci., 42, 2225–2241, https://doi.org/10.1175/1520-0469(1985)042<2225:AOGCMS>2.0.CO;2.
Chen, G., J. Ling, R. Zhang, Z. Xiao, and C. Li, 2022: The MJO from CMIP5 to CMIP6: Perspectives from tracking MJO precipitation. Geophys. Res. Lett., 49, e2021GL095241, https://doi.org/10.1029/2021GL095241.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, https://doi.org/10.1002/qj.828.
Deser, C., L. Terray, and A. S. Phillips, 2016: Forced and internal components of winter air temperature trends over North America during the past 50 years: Mechanisms and implications. J. Climate, 29, 2237–2258, https://doi.org/10.1175/JCLI-D-15-0304.1.
Di Carlo, E., and P. Ruggieri, P. Davini, S. Tibaldi and S. Corti, 2022: ENSO teleconnections and atmospheric mean state in idealised simulations. Climate Dyn., 59, 3287–3304, https://doi.org/10.1007/s00382-022-06261-w.
Eyring, V., S. Bony, G. A. Meehl, C. A. Senior, B. Stevens, R. J. Stouffer, and K. E. Taylor, 2016: Overview of the Coupled Model Intercomparison Project phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev., 9, 1937–1958, https://doi.org/10.5194/gmd-9-1937-2016.
Gates, W. L., 1992: AMIP: The Atmospheric Model Intercomparison Project. Bull. Amer. Meteor. Soc., 73, 1962–1970, https://doi.org/10.1175/1520-0477(1992)073<1962:ATAMIP>2.0.CO;2.
Gelaro, R., and Coauthors, 2017: The Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2). J. Climate, 30, 5419–5454, https://doi.org/10.1175/JCLI-D-16-0758.1.
Held, I. M., and I.-S. Kang, 1987: Barotropic models of the extratropical response to El Niño. J. Atmos. Sci., 44, 3576–3586, https://doi.org/10.1175/1520-0469(1987)044<3576:BMOTER>2.0.CO;2.
Held, I. M., S. W. Lyons, and S. Nigam, 1989: Transients and the extratropical response to El Niño. J. Atmos. Sci., 46, 163–174, https://doi.org/10.1175/1520-0469(1989)046<0163:TATERT>2.0.CO;2.
Herbasch, H., and Coauthors, 2020: ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803.
Hoerling, M. P., and A. Kumar, 2002: Atmospheric response patterns associated with tropical forcing. J. Climate, 15, 2184–2203, https://doi.org/10.1175/1520-0442(2002)015<2184:ARPAWT>2.0.CO;2.
Hoerling, M. P., A. Kumar, and M. Zhong, 1997: El Niño, La Niña, and the nonlinearity of their teleconnections. J. Climate, 10, 1769–1786, https://doi.org/10.1175/1520-0442(1997)010<1769:ENOLNA>2.0.CO;2.
Horel, J. D., and J. M. Wallace, 1981: Planetary-scale atmospheric phenomena associated with the Southern Oscillation. Mon. Wea. Rev., 109, 813–829, https://doi.org/10.1175/1520-0493(1981)109<0813:PSAPAW>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 in 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., A. J. Simmons, and D. G. Andrews, 1977: Energy dispersion in a barotropic atmosphere. Quart. J. Roy. Meteor. Soc., 103, 553–567, https://doi.org/10.1002/qj.49710343802.
Hourdin, F., and Coauthors, 2017: The art and science of climate model tuning. Bull. Amer. Meteor. Soc., 98, 589–602, https://doi.org/10.1175/BAMS-D-15-00135.1.
Jin, 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.
Kang, I.-S., and I. M. Held, 1986: Linear and nonlinear diagnostic models of stationary eddies in the upper troposphere during northern summer. J. Atmos. Sci., 43, 3045–3057, https://doi.org/10.1175/1520-0469(1986)043<3045:LANDMO>2.0.CO;2.
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.
Kumar, A., and M. P. Hoerling, 1995: Prospects and limitations of seasonal atmospheric GCM predictions. Bull. Amer. Meteor. Soc., 76, 335–345, https://doi.org/10.1175/1520-0477(1995)076<0335:PALOSA>2.0.CO;2.
Kumar, A., and M. P. Hoerling, 1997: Interpretation and implications of the observed inter-El Niño variability. J. Climate, 10, 83–91, https://doi.org/10.1175/1520-0442(1997)010<0083:IAIOTO>2.0.CO;2.
Kumar, A., Q. Zhang, P. Peng, and B. Jha, 2005: SST-forced atmospheric variability in an atmospheric general circulation model. J. Climate, 18, 3953–3967, https://doi.org/10.1175/JCLI3483.1.
Lau, N.-C., 1981: A diagnostic study of recurrent meteorological anomalies appearing in a 15-year simulation with a GFDL general circulation model. Mon. Wea. Rev., 109, 2287–2311, https://doi.org/10.1175/1520-0493(1981)109<2287:ADSORM>2.0.CO;2.
Lau, N.-C., 1997: Interactions between global SST anomalies and the midlatitude atmospheric circulation. Bull. Amer. Meteor. Soc., 78, 21–34, https://doi.org/10.1175/1520-0477(1997)078<0021:IBGSAA>2.0.CO;2.
Lau, N.-C., and M. J. Nath, 1994: A modeling study of the relative roles of tropical and extratropical SST anomalies in the variability of the global atmosphere–ocean system. J. Climate, 7, 1184–1207, https://doi.org/10.1175/1520-0442(1994)007<1184:AMSOTR>2.0.CO;2.
Li, R. K. K., T. Woollings, C. O’Reilly, and A. A. Scaife, 2020: Effect of the North Pacific tropospheric waveguide on the fidelity of model El Niño teleconnections. J. Climate, 33, 5223–5237, https://doi.org/10.1175/JCLI-D-19-0156.1.
Lopez, H., and B. Kirtman, 2019: ENSO influence over the Pacific North American sector: Uncertainty due to atmospheric internal variability. Climate Dyn., 52, 6149–6172, https://doi.org/10.1007/s00382-018-4500-0.
Maloney, E. D., and Coauthors, 2019: Process-oriented evaluation of climate and weather forecasting models. Bull. Amer. Meteor. Soc., 100, 1665–1686, https://doi.org/10.1175/BAMS-D-18-0042.1.
Palmer, T. N., and D. A. Mansfield, 1986: A study of wintertime circulation anomalies during past El Niño events using a high resolution general circulation model. I: Influence of model climatology. Quart. J. Roy. Meteor. Soc., 112, 613–638, https://doi.org/10.1002/qj.49711247304.
Peng, P., and A. Kumar, 2005: A large ensemble analysis of the influence of tropical SSTs on seasonal atmospheric variability. J. Climate, 18, 1068–1085, https://doi.org/10.1175/JCLI-3314.1.
Peng, P., A. Kumar, and B. Jha, 2014: Climate mean, variability and dominant patterns of the Northern Hemisphere wintertime mean atmospheric circulation in the NCEP CFSv2. Climate Dyn., 42, 2783–2799, https://doi.org/10.1007/s00382-014-2116-6.
Rasmusson, E. M., and T. H. Carpenter, 1982: Variations in tropical sea surface temperature and surface wind fields associated with the Southern Oscillation/El Niño. Mon. Wea. Rev., 110, 354–384, https://doi.org/10.1175/1520-0493(1982)110<0354:VITSST>2.0.CO;2.
Saha, S. S., and Coauthors, 2010: The NCEP Climate Forecast System reanalysis. Bull. Amer. Meteor. Soc., 91, 1015–1058, https://doi.org/10.1175/2010BAMS3001.1.
Sardeshmukh, P. D., and B. J. Hoskins, 1988: The generation of global rotational flow by steady idealized tropical divergence. J. Atmos. Sci., 45, 1228–1251, https://doi.org/10.1175/1520-0469(1988)045<1228:TGOGRF>2.0.CO;2.
Schmidt, G. A., and Coauthors, 2017: Practice and philosophy of climate model tuning across six US modeling centers. Geosci. Model Dev., 10, 3207–3223, https://doi.org/10.5194/gmd-10-3207-2017.
Shukla, J., and J. M. Wallace, 1983: Numerical simulation of the atmospheric response to equatorial Pacific sea surface temperature anomalies. J. Atmos. Sci., 40, 1613–1630, https://doi.org/10.1175/1520-0469(1983)040<1613:NSOTAR>2.0.CO;2.
Shukla, J., and Coauthors, 2000: Dynamical seasonal prediction. Bull. Amer. Meteor. Soc., 81, 2593–2606, https://doi.org/10.1175/1520-0477(2000)081<2593:DSP>2.3.CO;2.
Simmons, A. J., J. M. Wallace, and G. W. Branstator, 1983: Barotropic wave propagation, and instability, and atmospheric teleconnections. J. Atmos. Sci., 40, 1363–1392, https://doi.org/10.1175/1520-0469(1983)040<1363:BWPAIA>2.0.CO;2.
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.
Straus, D. M., and J. Shukla, 2002: Does ENSO force the PNA? J. Climate, 15, 2340–2358, https://doi.org/10.1175/1520-0442(2002)015<2340:DEFTP>2.0.CO;2.
Ting, M., and P. D. Sardeshmukh, 1993: Factors determining the extratropical response to equatorial diabatic heating anomalies. J. Atmos. Sci., 50, 907–918, https://doi.org/10.1175/1520-0469(1993)050<0907:FDTERT>2.0.CO;2.
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, 14 291–14 324, https://doi.org/10.1029/97JC01444.
Wallace, J. M., and D. S. Gutzler, 1981: Teleconnections in geopotential height field during the Northern Hemisphere winter. Mon. Wea. Rev., 109, 784–812, https://doi.org/10.1175/1520-0493(1981)109<0784:TITGHF>2.0.CO;2.
Wallace, J. M., E. M. Rasmusson, T. P. Mitchell, V. E. Kousky, E. S. Sarachik, and H. von Storch, 1998: On the structure and evolution of ENSO-related climate variability in the tropical Pacific: Lessons from TOGA. J. Geophys. Res., 103, 14 241–14 259, https://doi.org/10.1029/97JC02905.
Zhang, C., D. S. Nolan, C. D. Thorncroft, and H. Nguyen, 2008: Shallow meridional circulations in the tropical atmosphere. J. Climate, 21, 3453–3470, https://doi.org/10.1175/2007JCLI1870.1.