Next Article in Journal
Brake Particle PN and PM Emissions of a Hybrid Light Duty Vehicle Measured on the Chassis Dynamometer
Previous Article in Journal
Extratropical Cyclone Response to Projected Reductions in Snow Extent over the Great Plains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Revisiting the Influence of ENSO on the Arctic Stratosphere in CMIP5 and CMIP6 Models

1
Key Laboratory of Meteorological Disaster, Ministry of Education (KLME), Nanjing University of Information Science and Technology, Nanjing 210044, China
2
Joint International Research Laboratory of Climate and Environmental Change (ILCEC), Nanjing University of Information Science and Technology, Nanjing 210044, China
3
Collaborative Innovation Center on Forecast and Evaluation of Meteorological Disasters (CIC-FEMD), Nanjing University of Information Science and Technology, Nanjing 210044, China
4
School of Longshan, Nanjing University of Information Science and Technology, Nanjing 210044, China
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(5), 785; https://doi.org/10.3390/atmos14050785
Submission received: 20 March 2023 / Revised: 17 April 2023 / Accepted: 25 April 2023 / Published: 26 April 2023
(This article belongs to the Section Climatology)

Abstract

:
The Arctic stratospheric response to El Niño–Southern Oscillation (ENSO) is assessed using the historical simulations provided by the Coupled Model Intercomparison Project Phases 5 and 6 (CMIP5 and CMIP6, respectively). CMIP6 models can well reproduce the ENSO signals in the Arctic stratosphere and have an ameliorated performance compared to CMIP5 models. Specifically, El Niño is associated with an intensified Pacific–North American pattern that leads to a considerable enhancement of planetary wavenumber 1 but a small reduction of planetary wavenumber 2, and thus, a warm and weakened stratospheric polar vortex. The case for La Niña is nearly the opposite, with a cool and strengthened stratospheric polar vortex. In CMIP6, the ENSO-related stratospheric signal matures in the February–March–April season and increases with ENSO magnitude, regardless of the ENSO phase. However, the stratospheric response to strong El Niño (La Niña) is weaker (stronger) than that which should be achieved if the response changes linearly with the amplitude of El Niño (La Niña). An asymmetric time evolution of stratospheric signals exists between strong El Niño and La Niña events. The stratospheric response caused by strong El Niño is weaker from late winter to early spring but stronger in middle and late spring compared to that caused by strong La Niña. By contrast, the Arctic stratospheric signal in moderate El Niño events is larger than that in moderate La Niña. Compared to ENSO-neutral winters, stratospheric sudden warming occurs more (less) frequently in El Niño (La Niña), as simulated by CMIP6 high-top models.

1. Introduction

El Niño–Southern Oscillation (ENSO), which occurs in the tropical Pacific Ocean, is one of the most important ocean–atmosphere coupled phenomena. It affects oceanic and atmospheric conditions in both the tropics and extratropics. Previous studies have pointed out that ENSO remarkably impacts the stratosphere in both hemispheres. In the Northern Hemisphere, the warm ENSO phase (i.e., El Niño) acts to warm and weaken the winter stratospheric polar vortex, as revealed by both reanalysis data and model simulations [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17].
El Niño can modulate the Arctic stratosphere mainly through its mid-tropospheric teleconnection in the North Pacific. Specifically, it can induce a positive Pacific–North American (PNA) pattern and deepen the Aleutian low that would, in turn, constructively interfere with the climatological stationary troughs [6,7,18]. As a result, during El Niño, the planetary wavenumber 1 enhances in the troposphere and propagates upward into the extratropical stratosphere. When the planetary wave breaks, it decelerates the stratospheric polar jet, warms the polar region, and favors more occurrences of stratospheric sudden warming (SSW) events [19]. Although planetary wavenumber 2 decreases simultaneously, its effect is much smaller than that of planetary wavenumber 1. This is also the case for La Niña (the cold ENSO phase) but with the opposite sign; that is, the planetary wavenumber 1 reduces, but the planetary wavenumber 2 increases in the North Hemisphere, and thereby the stratospheric polar vortex strengthens during La Niña winters. However, the intensity of the impact of La Niña on the Arctic stratosphere is far less than that of El Niño [2,4,20].
Several controversial issues remain regarding ENSO’s impact on the Arctic stratosphere. For example, the stratospheric responses to ENSO are asymmetric in the Northern Hemisphere, as commonly thought [2,9,10,13,21,22]; that is, the response to El Niño is stronger than that to La Niña. However, Weinberger et al. [23] argued that no asymmetry is found in the Arctic stratospheric responses to El Niño as compared to La Niña. Besides, Rao and Ren [9,10] pointed out that the stratospheric responses to strong El Niño and moderate La Niña are weaker than those to moderate El Niño and strong La Niña. In contrast, some studies argued that the influence of strong El Niño is stronger than that of moderate El Niño [23,24,25,26].
In addition, the time evolution of the stratospheric signals induced by ENSO is suggested to be different on sub-seasonal time scales. For instance, Calvo et al. [27] detected obvious signals of eastern Pacific El Niño in the boreal stratospheric polar region from November to March. Trascasa-Castro et al. [25] also found robust stratospheric responses to extreme El Niño as early as November. However, it is also argued that El Niño-related signals may not occur in the Arctic stratosphere earlier than January [2,22,24]. There are also controversies in the relationship between ENSO and SSW. It is revealed that El Niño tends to cause more SSW events while La Niña leads to fewer ones, compared to ENSO-neutral winters [19,23]. On the contrary, Butler and Polvani [28] and Butler et al. [29] reported that the occurrence frequency of SSW in La Niña winters is equal to that in El Niño winters in observations. Song and Son [30] indicated that the relationship between ENSO and SSW depends on the SSW identification.
It is suggested that sufficient composite samples are necessary to detect robust stratospheric response to ENSO [8,23,31]. The model simulations from the Coupled Model Intercomparison Project (CMIP) can provide large samples of ENSO events to examine their signals. For example, Hurwitz et al. [21] analyzed ENSO signals in the extratropical atmosphere using 13 high-top models from the CMIP Phases 5 (CMIP5). Calvo et al. [27] also assessed the Northern Hemisphere stratospheric response to different El Niño flavors based on 11 CMIP5 high-top models.
This study aims to revisit the influence of ENSO on the Arctic stratosphere in CMIP Phases 6 (CMIP6) models and present a comparison with CMIP5. We focus on the following two aspects: (1) the responses of the Arctic stratosphere to ENSO in CMIP6 models, such as the relationship between ENSO and SSW events, the asymmetry and nonlinearity of ENSO signals in the stratospheric polar region, and (2) comparisons of the model performance in simulating the ENSO’s impact on the boreal winter stratospheric polar vortex between CMIP6 and CMIP5 models and between CMIP6 high-top and low-top models.

2. Data and Methods

2.1. Data

In this study, we select 21 CMIP5 (Table 1) and 15 CMIP6 models (Table 2) with daily geopotential height, air temperature, zonal and meridional wind, and sea surface temperature (SST) fields available to us. The first member (r1i1p1 and r1i1p1f1) of the historical experiments for each model has been analyzed. The horizontal resolution of SST and atmospheric variables are interpolated to 1° × 1° and 2.5° × 2.5°, respectively, based on the bilinear interpolation method. Atmospheric variables include eight pressure levels (1000, 850, 700, 500, 250, 100, 50, and 10 hPa) in the vertical direction. The time period covered by each model is shown in Table 1 and Table 2.

2.2. Methods

To identify ENSO events, we first calculate the standardized detrended winter (December–January–February; DJF)-mean Niño3.4 index by averaging the SST anomalies in the Niño3.4 region (5° S–5° N, 170°–120° W). We focus on strong and moderate ENSO events. In each model, strong El Niño (La Niña) events are defined as the Niño3.4 index above 1.5 (below –1.5) standard deviations, and moderate El Niño (La Niña) events as the Niño3.4 index between 0.75 and 1.5 (–0.75 and –1.5) standard deviations. Other winters are classified as ENSO-neutral winters. The total number of strong and moderate ENSO events identified in the CMIP5 and CMIP6 models is shown in Table 3.
Following Charlton and Polvani [32], the SSW events are identified when the zonal-mean zonal wind at 10 hPa and 60° N reverses from a westerly to an easterly wind during the period from November to March. However, we exclude the stratospheric final warming events with a final direction reversal of circumpolar zonal-mean zonal wind. In addition, similar to Charlton and Polvani [32], there should be 20 consecutive days of westerly wind between two individual SSW events. Given this, the SSW events in this study represent the major SSW events, according to their definition from the World Meteorological Organization. However, for simplicity, the word major is omitted in the context.
We use two methods to diagnose planetary wave activities. One is a Fourier decomposition of the waves in geopotential height anomalies in the zonal direction. The other is the Eliassen–Palm (EP) flux. Under the quasi-geostrophic approximation, equations of EP flux in spherical coordinates are as follows [33]:
F ϕ = ρ a cos ϕ ( v u ¯ )
F z = ρ a f cos ϕ v θ ¯ ( θ ¯ z ) 1
F = ( a cos ϕ ) 1 ϕ ( F ϕ cos ϕ ) + F z z
where “ ¯ ”and “ ′ ” represent the zonal mean and the zonal deviation from the zonal mean, respectively; ρ, a and φ are the air density, the Earth’s radius and the latitude, respectively; θ and f represent the potential temperature and geostrophic parameters, respectively; Fφ is the horizontal component of EP flux, representing the eddy momentum flux; Fz is the vertical component of EP (EPz) flux, representing the eddy heat transport; and F is the divergence of EP flux.
We define the PNA index for each month in each CMIP6 model using the standardized geopotential height anomalies (represented by Z in the Equation (4)) at four points [34]:
PNA = 1/4 [Z (20° N, 160° W) − Z (45° N, 165° W) + Z (55° N, 115° W) − Z (30° N, 85° W)].
In Section 3.4, the Aleutian low strength and the stratospheric polar vortex strength are diagnosed. The strength of the Aleutian low is defined as 500-hPa geopotential height anomalies averaged over 30°–55° N, 170° E–140° W, and the strength of the stratospheric polar vortex is defined as 10-hPa geopotential height anomalies averaged poleward of 70° N.
For a given variable in each model, both long-term climatology and linear trend are removed to obtain anomalies before composite analyses. The climatology of an atmospheric variable is averaged over all years in all models; thus, there are 1734 and 1960 years for CMIP5 and CMIP6 models, respectively. The two-tailed Student’s t-test is used to test the significance of the composites.

3. Results

3.1. SST Anomalies and Arctic Stratospheric Response Associated with ENSO Events

Figure 1 shows the composite DJF-mean SST anomalies (SSTa) associated with ENSO events with different magnitudes in CMIP6 and CMIP5 models. The spatial pattern of SSTa for the El Niño and La Niña events in both CMIP6 and CMIP5 models are similar to the observations. However, the magnitude of SSTa in CMIP6 is relatively larger than those in CMIP5 (Figure 1a,b,d,e,j,k,m,n) with the discrepancies of 0.71, 0.39, 0.47, and 0.25 K in the Niño3.4 region for strong El Niño, moderate El Niño, strong La Niña, and moderate La Niña, respectively (Table 3). This implies that the CMIP6 models’ capability is improved in simulating different ENSO events, given the fact that the ENSO intensity is underestimated by CMIP5 models [35].
Figure 2 shows the temporal evolution of composite geopotential height anomalies area-averaged over the polar cap (70°–90° N, 0–360°) for different ENSO groups in CMIP6 and CMIP5 models. Clearly, the CMIP6 models can well reproduce the response of the Arctic stratosphere to ENSO; namely, El Niño greatly weakens the stratospheric polar vortex, while La Niña strengthens it (Figure 2a–d). We also see that the stratospheric responses are larger in strong ENSO events (Figure 2a,c) than those in moderate ENSO events (Figure 2b,d), indicating a monotonic increase of ENSO impact with its amplitude. Significant ENSO signals mainly appear from February to April, although the signals can extend to early May in strong El Niño events (Figure 2a), and appear relatively early, in mid-January, in strong La Niña events (Figure 2c). Generally, the CMIP5 models can also reproduce the response of the Arctic stratosphere to ENSO (Figure 2e–h) but with a less significant and much smaller magnitude as compared to CMIP6 models.
Previous studies suggested that only CMIP5 high-top models can better simulate the Arctic stratospheric response to ENSO [21,27]. Following Charlton-Perez et al. [36] and Wilcox and Charlton-Perez [37], we also divide the CMIP6 models into high-top (with model top pressure ≤ 1 hPa) and low-top models (with the model top pressure > 1 hPa) (see Table 2). Figure 3 shows the temporal evolution of composite polar-cap mean geopotential height anomalies for different ENSO groups in CMIP6 high-top and low-top models. In high-top models, the stratospheric responses to strong and moderate El Niño and strong La Niña events resemble those averaged over all the CMIP6 models; however, little signals can be found in moderate La Niña events (Figure 3d). The reason for this remains unknown. Iza et al. [20] found that while strong La Niña has a robust signature in the winter Arctic stratosphere in reanalysis data, the signals from relatively weaker La Niña might be obscured by the impacts from other factors such as the quasi-biennial oscillation and SSW events. Therefore, it might be the relatively high SSW frequency (~0.62 yr–1, Table 4, see Section 3.2 for details) that would lead to the undistinguished stratospheric signals in moderate La Niña events. This needs further investigation. In low-top models, the stratospheric polar vortex is also significantly weak in El Niño and strong in La Niña events from February to April. For moderate El Niño and strong La Niña events (Figure 3f,g), the stratospheric signals in early winter (from mid-November to mid-January) are opposite to those in late winter and early spring. Similar opposite signals between early winter and late winter during El Niño are also emphasized in previous studies [2,22]. Thus, such sub-seasonal signals cannot be sufficiently captured by seasonal mean responses, highlighting the importance of winter period selection for investigating the stratospheric responses to ENSO. These results indicate that the capability of low-top models is significantly improved from CMIP5 to CMIP6 in simulating the response of the Northern Hemisphere stratospheric polar vortex to ENSO since the stratospheric dynamical variability is poorly reproduced in CMIP5 low-top models [36].

3.2. ENSO-SSW Relationship

Previous studies reported different relationships between ENSO and SSW [19,23,28,29,30]. Here, we revisit such relationships in CMIP5 and CMIP6 models. Table 3 provides the number of SSW events with their frequency in CMIP5 and CMIP6 models for ENSO and ENSO-neutral conditions. The SSW frequencies are 0.58 and 0.59 yr–1, respectively, for all years and the ENSO-neutral years in CMIP6 models, which is consistent with the results from Charlton and Polvani [32] with a frequency of about six events per decade in the reanalysis data. The SSW frequency is slightly enhanced under El Niño conditions and reduced under La Niña conditions compared to ENSO-neutral conditions. Overall, the SSW events occur a little more frequently in strong ENSO events than in moderate ENSO events. By contrast, the frequency of SSW in CMIP5 models is quite low, with only ~0.3 events per year. This agrees well with the results of Rao and Garfinkel [38] that most of the CMIP5 models underestimate the SSW frequency. This may be partly due to the weaker planetary wave activities in CMIP5 compared to CMIP6 models (see next subsection). Additionally, it may be because the winter stratospheric polar vortex simulated by CMIP6 models is closer to observations than CMIP5 models [39,40].
Table 4 lists the numbers of ENSO and SSW events in CMIP6 high-top and low-top models. In high-top models, the SSW frequencies under different ENSO conditions and ENSO-neutral winters are around 0.6–0.7 events per year, which are closer to the observations and much higher than those in the low-top models. This indicates a remarkable improvement of the CMIP6 high-top models in simulating the stratospheric polar variability. Nevertheless, CMIP6 low-top models still outperform CMIP5 models in simulating the boreal winter SSW events, with an evident increase in SSW frequency. In addition, compared to those in ENSO-neutral winters, the SSW frequency is slightly enhanced during El Niño and reduced during La Niña in CMIP6 high-top models.

3.3. Tropospheric Teleconnection and Planetary Wave Activities

As noted previously, the main pathway through which ENSO affects the stratospheric polar vortex is its tropospheric teleconnection in the North Pacific, which interferes with the climatological stationary waves and thus modulates the propagation of planetary waves from the troposphere into the extratropical stratosphere. Figure 4 and Figure 5 show the composite 500-hPa geopotential height anomalies averaged from January to March and their wavenumber 1 and 2 components for different ENSO groups in CMIP6 and CMIP5 models, respectively. In CMIP6 models, a canonical pattern can be clearly seen over the PNA region, with an anomalous low over the Northeastern Pacific and eastern United States, respectively, and an anomalous high over Canada during El Niño (Figure 4a,d). In contrast, La Niña induces a nearly opposite PNA pattern (Figure 4g,j). The intensity of ENSO-related signals in the middle troposphere increases with ENSO amplitude. In a total of 921 strong and moderate ENSO events, the correlation coefficient between the winter Niño3.4 SSTa and JFM PNA index reaches up to 0.63. As seen in the middle and right panels in Figure 4, this pattern has a constructive (destructive) interference with the climatological stationary wavenumber 1 and destructive (constructive) interference with wavenumber 2 during El Niño (La Niña), leading to enhanced (suppressed) planetary wave activities entering into the extratropical stratosphere [6]. The composite geopotential height anomalies in the upper troposphere and low stratosphere exhibit similar results (figures not shown). Overall, the CMIP5 models can also reproduce the ENSO teleconnection in the PNA region (Figure 5). However, such teleconnection exhibits a more typical spatial pattern and larger amplitude in CMIP6 models.
Next, we diagnose the EPz flux at 100 hPa, which roughly represents the upward propagation of planetary wave activities from the troposphere into the extratropical stratosphere. Figure 6 shows the temporal evolution of composite EPz flux anomalies of planetary waves (wavenumbers 1–3), and the wavenumber 1 and wavenumber 2 components averaged in 55°–75° N for different ENSO groups in CMIP6 and CMIP5 models. In CMIP6 models (Figure 6a–d), the wavenumber 1 (wavenumber 2) of EPz flux anomalies intensifies, and wavenumber 2 (wavenumber 1) reduces during El Niño (La Niña) mainly from January to March, matching the results in Figure 4. We further calculate the correlation coefficients of the PNA index with the wavenumber 1 and wavenumber 2 EPz flux during the ENSO events in CMIP6 models, which are 0.37 and –0.32 during January–February–March season, respectively, exceeding the 95% confidence level. In Figure 6a–d, the increase (decrease) of wavenumber 1 overwhelms the decrease (increase) of wavenumber 2 during El Niño (La Niña), resulting in active (dormant) upward propagation of planetary waves and thus a weakened (strengthened) stratospheric polar vortex. Also, the ENSO-related planetary wave activities enhance with ENSO amplitude. In CMIP5 models (Figure 6e–h), the temporal evolution of planetary wave activities resembles that in CMIP6 models. Nevertheless, the CMIP6 models can better characterize the planetary wave activities, at least in strong ENSO events.

3.4. Asymmetry and Nonlinearity

ENSO signals exhibit obvious asymmetry and nonlinearity in both the troposphere and stratosphere. We have shown that the Arctic stratospheric and mid-tropospheric responses to El Niño and La Niña are symmetric in CMIP6 in terms of the sign. Now we turn to the amplitude of the response. Figure 7 shows the ratios of the Aleutian low strength and the stratospheric polar vortex strength averaged in two adjacent months between two different ENSO groups in CMIP6 models. The ratios of winter SSTa in the Niño3.4 region in different ENSO groups are also shown. It is found that the asymmetries in terms of the SSTa amplitude are relatively weak between strong El Niño and strong La Niña (~1.15) and between moderate El Niño and moderate La Niña (~0.94). We note that large SSTa can lead to even stronger responses if the ratio of Aleutian low strength or stratospheric polar vortex strength between two ENSO groups exceeds that of SSTa. Otherwise, large SSTa can only result in relatively weak responses.
It is seen that strong El Niño has a smaller winter response in the middle troposphere and Arctic stratosphere as compared to strong La Niña (purple bars in Figure 7), confirming the results of Rao and Ren [9,10]. In the stratosphere, the average strength of stratospheric polar vortex in strong El Niño is only 50–60% of that in strong La Niña from January to March. However, in spring, the response to strong El Niño is much larger than that of strong La Niña. This is because significant La Niña signals in the Arctic stratosphere can be found as early as mid-January and peak in February and March, while strong El Niño signals peak in March and April (Figure 2a,c). By contrast, for moderate ENSO events (green bars in Figure 7), the responses to El Niño in both the troposphere and stratosphere are larger than those to La Niña: on average, 1.65 times those to La Niña in the Arctic stratosphere from February to April, which is also consistent with that in Rao and Ren [9,10]. However, for all the moderate and strong ENSO events, the responses to El Niño and La Niña are almost equal in magnitude (the ratio is around ~1 for both the DJFM-mean Aleutian low strength and JFMA-mean stratospheric polar vortex strength). This indicates fairly symmetric responses between El Niño and La Niña, which is identical to the results in Weinberger et al. [23]. In addition, although strong El Niño leads to a stronger Aleutian low and stratospheric polar vortex than moderate El Niño (red bars in Figure 7), such an increase is not linear. That is, nearly double SSTa during strong El Niño do not generate double responses in the middle troposphere and Arctic stratosphere, verifying the findings of Weinberger et al. [23]. For La Niña, such nonlinearity seems to be a little bit complex. Less large Aleutian low in strong La Niña is associated with a much larger response in the stratosphere as compared to moderate La Niña (blue bars in Figure 7). The stratospheric polar vortex strength in strong La Niña is around a factor of 2.3 larger than that in moderate La Niña on average from January to April, agreeing well with the findings of Iza et al. [20] that robust polar stratospheric responses can only be found in strong La Niña events.

4. Summary

The amplitude and zonal location of the maximum SSTa are usually different among ENSO events, leading to conspicuous diversity or so-called asymmetry or nonlinearity of ENSO signals. In this study, we reexamine the ENSO’s impact on the Arctic stratosphere based on large samples of ENSO events provided by historical simulations from CMIP6 and CMIP5 models. The main conclusions are as follows.
CMIP6 models can well reproduce the Arctic stratospheric responses to ENSO. Specifically, El Niño leads to a deepened low in the North Pacific and an enhanced high over Canada in the middle troposphere, which constructively interferes with the climatological stationary wave. Although the planetary wavenumber 2 is relatively reduced, the wavenumber 1 is considerably reinforced, resulting in a net effect of increased vertically-propagating planetary waves from the troposphere into the extratropical stratosphere and thus a warm and weakened stratospheric polar vortex in the Northern Hemisphere. The case for La Niña is nearly the opposite, with a cool and strengthened stratospheric polar vortex. Compared to the CMIP5 models, the CMIP6 models’ capability is remarkably improved in simulating ENSO and its Arctic stratospheric response, manifesting as the larger amplitude of SSTa in Niño3.4 region, stronger and more typical PNA pattern, more active planetary waves, and intensified response of the stratospheric polar vortex.
The occurrence of SSW is 0.58 events per year in CMIP6 models, which is quite close to the observations. The SSW frequency is slightly increased during El Niño and decreased during La Niña relative to ENSO-neutral winters in CMIP6 high-top models. However, this is not true for the CMIP5 models and CMIP6 low-top models.
In CMIP6 models, ENSO signals in the Arctic stratosphere mature in late winter and early spring, mainly from February to April, irrespective of the ENSO phase. Overall, the responses of the 10-hPa stratospheric polar vortex to El Niño and La Niña are symmetric, but not for those induced by strong and moderate ENSO events. The asymmetry of the response to strong El Niño and La Niña is characterized by a different temporal evolution in the polar stratosphere. Significant signals associated with strong La Niña appear from early January and peak in February and early March, while strong El Niño-related signals peak in March and April. For moderate ENSO, the stratospheric response to El Niño is much larger than that to equivalent magnitude La Niña.
The Arctic stratospheric responses increase monotonically with ENSO magnitude in CMIP6 models, whereas the increase is not proportional to that of Niño3.4 SSTa. The response to strong El Niño events is weaker than that which should be achieved if the response changes linearly with the amplitude of El Niño. In contrast, strong La Niña leads to a much larger Arctic stratospheric response, around 2.3 times those to moderate La Niña.
The analyses in this study are based on a large number of ENSO samples provided by CMIP6 and CMIP5 models. This overcomes some uncertainties using limited samples in the reanalysis data and meanwhile helps to minimize the contamination of ENSO signals by other factors. For example, the main results in this study remain unchanged if we remove the linear effect of the stratospheric quasi-biennial oscillation on the stratospheric circulation (figures not shown). As CMIP6 models outperform CMIP5 models in simulating the polar stratospheric response to ENSO, similar approaches could also be used to examine the stratospheric response to eastern and central Pacific El Niño, respectively, in both hemispheres. Relevant studies are ongoing and will be reported in the future.

Author Contributions

Conceptualization, J.H. and Y.S.; methodology, J.H., Y.S. and J.D.; software, Y.S., J.D., Y.J. and Z.W.; validation, J.H., Z.W. and A.L.; formal analysis, J.H. and Y.S.; investigation, J.H., Y.S. and Y.J.; resources, J.H., Y.S. and J.D.; data curation, Y.S. and A.L.; writing—original draft preparation, J.H.; writing—review and editing, Y.S., J.D. and Z.W.; visualization, J.H. and Y.S.; supervision, J.H. and J.D.; project administration, J.H.; funding acquisition, J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (41975048, 42030605, 42175069) and the Natural Science Foundation of Jiangsu province (BK20191404).

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Acknowledgments

We thank the four anonymous reviewers for their assistance in evaluating this paper. We thank the ESGF (https://esgf-node.llnl.gov/projects/esgf-llnl/, accessed on 1 March 2022) for providing the CMIP5 and CMIP6 simulations. We also thank Xiao Zhang for the helpful discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sassi, F.; Kinnison, D.; Boville, B.A.; Garcia, R.R.; Roble, R. Effect of El Niño–Southern Oscillation on the dynamical, thermal, and chemical structure of the middle atmosphere. J. Geophys. Res. Atmos. 2004, 109, D17108. [Google Scholar] [CrossRef]
  2. Manzini, E.; Giorgetta, M.A.; Esch, M.; Kornblueh, L.; Roeckner, E. The influence of sea surface temperatures on the Northern winter stratosphere: Ensemble simulations with the MAECHAM5 model. J. Clim. 2006, 19, 3863–3881. [Google Scholar] [CrossRef]
  3. Camp, C.D.; Tung, K.K. Stratospheric polar warming by ENSO in winter: A statistical study. Geophys. Res. Lett. 2007, 34, L04809. [Google Scholar] [CrossRef]
  4. Free, M.; Seidel, D.J. Observed El Niño-Southern Oscillation temperature signal in the stratosphere. J. Geophys. Res. 2009, 114, D23108. [Google Scholar] [CrossRef]
  5. Garfinkel, C.I.; Hartmann, D.L. Effects of the El Niño-Southern Oscillation and the Quasi-Biennial Oscillation on polar temperatures in the stratosphere. J. Geophys. Res. 2007, 112, D19112. [Google Scholar] [CrossRef]
  6. Garfinkel, C.I.; Hartmann, D.L. Different ENSO teleconnections and their effects on the stratospheric polar vortex. J. Geophys. Res. Atmos. 2008, 113, D18114. [Google Scholar] [CrossRef]
  7. Garfinkel, C.I.; Hartmann, D.L.; Sassi, F. Tropospheric precursors of anomalous Northern Hemisphere stratospheric polar vortices. J. Clim. 2010, 23, 3282–3299. [Google Scholar] [CrossRef]
  8. Garfinkel, C.I.; Hurwitz, M.M.; Waugh, D.W.; Butler, A.H. Are the teleconnections of Central Pacific and Eastern Pacific El Niño distinct in boreal wintertime? Clim. Dyn. 2013, 41, 1835–1852. [Google Scholar] [CrossRef]
  9. Rao, J.; Ren, R. Asymmetry and nonlinearity of the influence of ENSO on the northern winter stratosphere: 1. Observations. J. Geophys. Res. Atmos. 2016, 121, 9000–9016. [Google Scholar] [CrossRef]
  10. Rao, J.; Ren, R. Asymmetry and nonlinearity of the influence of ENSO on the northern winter stratosphere: 2. Model study with WACCM. J. Geophys. Res. Atmos. 2016, 121, 9017–9032. [Google Scholar] [CrossRef]
  11. Xie, F.; Li, J.; Tian, W.; Feng, J.; Huo, Y. Signals of El Niño Modoki in the tropical tropopause layer and stratosphere. Atmos. Chem. Phys. 2012, 12, 5259–5273. [Google Scholar] [CrossRef]
  12. Xie, F.; Li, J.; Tian, W.; Zhang, J.; Sun, C. The relative impacts of El Niño Modoki, canonical El Niño, and QBO on tropical ozone changes since the 1980s. Environ. Res. Lett. 2014, 9, 064020. [Google Scholar] [CrossRef]
  13. Xie, F.; Zhou, X.; Li, J.; Sun, C.; Feng, J.; Ma, X. The key role of background sea surface temperature over the cold tongue in asymmetric responses of the Arctic stratosphere to El Niño–Southern Oscillation. Environ. Res. Lett. 2018, 13, 114007. [Google Scholar] [CrossRef]
  14. Hu, J.; Li, T.; Xu, H.; Yang, S. Lessened response of boreal winter stratospheric polar vortex to El Niño in recent decades. Clim. Dyn. 2017, 49, 263–278. [Google Scholar] [CrossRef]
  15. Yang, S.; Li, T.; Hu, J.; Shen, X. Decadal variation of the impact of La Niña on the winter Arctic stratosphere. Adv. Atmos. Sci. 2017, 34, 679–684. [Google Scholar] [CrossRef]
  16. Domeisen, D.I.V.; Garfinkel, C.I.; Butler, A.H. The teleconnection of El Niño Southern Oscillation to the stratosphere. Rev. Geophys. 2019, 57, 5–47. [Google Scholar] [CrossRef]
  17. Mezzina, B.; Palmeiro, F.M.; García-Serrano, J.; Bladé, I.; Batté, L.; Benassi, M. Multi-model assessment of the late-winter stratospheric response to El Niño and La Niña. Clim. Dyn. 2021, 58, 1987–2007. [Google Scholar] [CrossRef]
  18. Ineson, S.; Scaife, A.A. The role of the stratosphere in the European climate response to El Niño. Nat. Geosci. 2009, 2, 32–36. [Google Scholar] [CrossRef]
  19. Taguchi, M.; Hartmann, D.L. Increased occurrence of stratospheric sudden warmings during El Niño as simulated by WACCM. J. Clim. 2006, 19, 324–332. [Google Scholar] [CrossRef]
  20. Iza, M.; Calvo, N.; Manzini, E. The stratospheric pathway of La Niña. J. Clim. 2016, 29, 8899–8914. [Google Scholar] [CrossRef]
  21. Hurwitz, M.M.; Calvo, N.; Garfinkel, C.I.; Butler, A.H.; Ineson, S.; Cagnazzo, C.; Manzini, E.; Pena-Ortiz, C. Extra-tropical atmospheric response to ENSO in the CMIP5 models. Clim. Dyn. 2014, 43, 3367–3376. [Google Scholar] [CrossRef]
  22. Hardiman, S.C.; Dunstone, N.J.; Scaife, A.A.; Smith, D.M.; Ineson, S.; Lim, J.; Fereday, D. The impact of strong El Niño and La Niña events on the North Atlantic. Geophys. Res. Lett. 2019, 46, 2874–2883. [Google Scholar] [CrossRef]
  23. Weinberger, I.; Garfinkel, C.I.; White, I.P.; Oman, L.D. The salience of nonlinearities in the boreal winter response to ENSO: Arctic stratosphere and Europe. Clim. Dyn. 2019, 53, 4591–4610. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, X.; Li, J.; Xie, F.; Chen, Q.; Ding, R.; Zhang, W.; Li, Y. Does extreme El Niño have a different effect on the stratosphere in boreal winter than its moderate counterpart? J. Geophys. Res. Atmos. 2018, 123, 3071–3086. [Google Scholar] [CrossRef]
  25. Trascasa-Castro, P.; Maycock, A.C.; Yiu, Y.Y.S.; Fletcher, J.K. On the linearity of the stratospheric and Euro-Atlantic sector response to ENSO. J. Clim. 2019, 32, 6607–6626. [Google Scholar] [CrossRef]
  26. Richter, J.H.; Deser, C.; Sun, L. Effects of stratospheric variability on El Niño teleconnections. Environ. Res. Lett. 2015, 10, 124021. [Google Scholar] [CrossRef]
  27. Calvo, N.; Iza, M.; Hurwitz, M.M.; Manzini, E.; Pena-Ortiz, C.; Butler, A.H.; Cagnazzo, C.; Ineson, S.; Garfinkel, C.I. Northern Hemisphere stratospheric pathway of different El Niño flavors in stratosphere-resolving CMIP5 Models. J. Clim. 2017, 30, 4351–4371. [Google Scholar] [CrossRef]
  28. Butler, A.H.; Polvani, L.M. El Niño, La Niña, and stratospheric sudden warmings: A reevaluation in light of the observational record. Geophys. Res. Lett. 2011, 38, L13807. [Google Scholar] [CrossRef]
  29. Butler, A.H.; Polvani, L.M.; Deser, C. Separating the stratospheric and tropospheric pathways of El Niño-Southern Oscillation teleconnections. Environ. Res. Lett. 2014, 9, 024014. [Google Scholar] [CrossRef]
  30. Song, K.; Son, S.-W. Revisiting the ENSO–SSW relationship. J. Clim. 2018, 31, 2133–2143. [Google Scholar] [CrossRef]
  31. Garfinkel, C.I.; Weinberger, I.; White, I.P.; Oman, L.D.; Aquila, V.; Lim, Y.-K. The salience of nonlinearities in the boreal winter response to ENSO: North Pacific and North America. Clim. Dyn. 2018, 52, 4429–4446. [Google Scholar] [CrossRef]
  32. Charlton, A.J.; Polvani, L.M. A new look at stratospheric sudden warmings. Part I: Climatology and modeling benchmarks. J. Clim. 2007, 20, 449–469. [Google Scholar] [CrossRef]
  33. Andrews, D.G.; Holton, J.R.; Leovy, C.B. Middle Atmosphere Dynamics; Academic Press: San Diego, CA, USA, 1987; p. 489. [Google Scholar]
  34. Wallace, J.M.; Gutzler, D.S. Teleconnections in the geopotential height field during the Northern Hemisphere Winter. Mon. Wea. Rev. 1981, 109, 784–812. [Google Scholar] [CrossRef]
  35. Kim, S.T.; Yu, J.-Y. The two types of ENSO in CMIP5 models. Geophys. Res. Lett. 2012, 39, L11704. [Google Scholar] [CrossRef]
  36. Charlton-Perez, A.J.; Baldwin, M.P.; Birner, T.; Black, R.X.; Butler, A.H.; Calvo, N.; Davis, N.A.; Gerber, E.P.; Gillett, N.; Hardiman, S.; et al. On the lack of stratospheric dynamical variability in low-top versions of the CMIP5 models. J. Geophys. Res. Atmos. 2013, 118, 2494–2505. [Google Scholar] [CrossRef]
  37. Wilcox, L.J.; Charlton-Perez, A.J. Final warming of the Southern Hemisphere polar vortex in high- and low-top CMIP5 models. J. Geophys. Res. Atmos. 2013, 118, 2535–2546. [Google Scholar] [CrossRef]
  38. Rao, J.; Garfinkel, C.I. CMIP5/6 models project little change in the statistical characteristics of sudden stratospheric warmings in the 21st century. Environ. Res. Lett. 2021, 16, 034024. [Google Scholar] [CrossRef]
  39. Rao, J.; Garfinkel, C.I.; Wu, T.; Lu, Y.; Chu, M. Mean state of the Northern Hemisphere stratospheric polar vortex in three generations of CMIP models. J. Clim. 2022, 35, 4603–4625. [Google Scholar] [CrossRef]
  40. Hu, J.; Liu, Z.; Xu, H.; Ren, R.; Jin, D. On the spring stratospheric final warming in CMIP5 and CMIP6 models. Sci. China Earth Sci. 2023, 66, 129–145. [Google Scholar] [CrossRef]
Figure 1. Composite of winter SST anomalies (SSTa; unit: °C) for (a,j) strong El Niño, (b,k) strong La Niña, (d,m) moderate El Niño, and (e,n) moderate La Niña years, respectively, in (ai) CMIP6 and (jr) CMIP5 models. (c,f,i) Composite of winter SSTa for (c) both strong El Niño and La Niña years, (f) both moderate El Niño and La Niña years, and (i) all ENSO years in CMIP6 models, respectively. (g,h) The SSTa differences (g) between strong and moderate El Niño years and (h) between strong and moderate La Niña years in CMIP6 models, respectively. (l,o,r), (p,q) As in (c,f,i) and (g,h), respectively, but for CMIP5 models. Dotted areas mark the 95% confidence level of the anomalies or differences based on the two-tailed Student’s t-test. EN = El Niño, LN = La Niña, and the same below.
Figure 1. Composite of winter SST anomalies (SSTa; unit: °C) for (a,j) strong El Niño, (b,k) strong La Niña, (d,m) moderate El Niño, and (e,n) moderate La Niña years, respectively, in (ai) CMIP6 and (jr) CMIP5 models. (c,f,i) Composite of winter SSTa for (c) both strong El Niño and La Niña years, (f) both moderate El Niño and La Niña years, and (i) all ENSO years in CMIP6 models, respectively. (g,h) The SSTa differences (g) between strong and moderate El Niño years and (h) between strong and moderate La Niña years in CMIP6 models, respectively. (l,o,r), (p,q) As in (c,f,i) and (g,h), respectively, but for CMIP5 models. Dotted areas mark the 95% confidence level of the anomalies or differences based on the two-tailed Student’s t-test. EN = El Niño, LN = La Niña, and the same below.
Atmosphere 14 00785 g001
Figure 2. Temporal evolution of composite geopotential height anomalies area-averaged over the polar cap (70°–90° N, 0–360°) for different ENSO groups in (ad) CMIP6 and (eh) CMIP5 models. Dotted areas mark the 95% confidence level of the anomalies. The number in the upper right corner of each figure represents the sample size for ENSO events in each ENSO group.
Figure 2. Temporal evolution of composite geopotential height anomalies area-averaged over the polar cap (70°–90° N, 0–360°) for different ENSO groups in (ad) CMIP6 and (eh) CMIP5 models. Dotted areas mark the 95% confidence level of the anomalies. The number in the upper right corner of each figure represents the sample size for ENSO events in each ENSO group.
Atmosphere 14 00785 g002
Figure 3. As in Figure 2, but for (ad) CMIP6 high-top and (eh) low-top models.
Figure 3. As in Figure 2, but for (ad) CMIP6 high-top and (eh) low-top models.
Atmosphere 14 00785 g003
Figure 4. Composite (left panels: a,d,g,j) 500-hPa geopotential height anomalies averaged from January to March and (middle panels: b,e,h,k) the wavenumber 1 and (right panels: c,f,i,l) wavenumber 2 components for different ENSO groups in CMIP6 models. Dotted areas mark the 95% confidence level of the anomalies. Black contours represent the corresponding climatological components at ±25, ±50, and ±75 gpm, respectively. Green boxes outline the region used for defining the Aleutian low strength (see the text for details).
Figure 4. Composite (left panels: a,d,g,j) 500-hPa geopotential height anomalies averaged from January to March and (middle panels: b,e,h,k) the wavenumber 1 and (right panels: c,f,i,l) wavenumber 2 components for different ENSO groups in CMIP6 models. Dotted areas mark the 95% confidence level of the anomalies. Black contours represent the corresponding climatological components at ±25, ±50, and ±75 gpm, respectively. Green boxes outline the region used for defining the Aleutian low strength (see the text for details).
Atmosphere 14 00785 g004
Figure 5. As in Figure 4, but for CMIP5 models.
Figure 5. As in Figure 4, but for CMIP5 models.
Atmosphere 14 00785 g005
Figure 6. Temporal evolution of composite anomalies of the vertical component of 100-hPa Eliassen–Palm (EPz) flux averaged in 55°–75° N for different ENSO groups in (ad) CMIP6 and (eh) CMIP5 models. An 11-day running mean is applied. The thick segments indicate the anomalies significant at the 95% confidence level.
Figure 6. Temporal evolution of composite anomalies of the vertical component of 100-hPa Eliassen–Palm (EPz) flux averaged in 55°–75° N for different ENSO groups in (ad) CMIP6 and (eh) CMIP5 models. An 11-day running mean is applied. The thick segments indicate the anomalies significant at the 95% confidence level.
Atmosphere 14 00785 g006
Figure 7. The ratios of (a) the Aleutian low strength and (b) the stratospheric polar vortex strength between two different ENSO groups in CMIP6 models. The dashed lines represent the ratios of winter Niño3.4 SSTa between two different ENSO groups. DJ = December–January, JF = January–February, FM = February–March, MA = March–April.
Figure 7. The ratios of (a) the Aleutian low strength and (b) the stratospheric polar vortex strength between two different ENSO groups in CMIP6 models. The dashed lines represent the ratios of winter Niño3.4 SSTa between two different ENSO groups. DJ = December–January, JF = January–February, FM = February–March, MA = March–April.
Atmosphere 14 00785 g007
Table 1. CMIP5 models used in this study.
Table 1. CMIP5 models used in this study.
No.CMIP5 ModelsResolution for
Atmospheric Variables (Lon × Lat)
Resolution for SST (Lon × Lat)Time Period
1ACCESS1-0192 × 144360 × 3001950–2005
2ACCESS1-3192 × 144360 × 3001950–2005
3CanESM2128 × 64256 × 1921950–2005
4CMCC-CESM96 × 48182 × 1491950–2005
5CMCC-CMS192 × 96182 × 1491950–2005
6CNRM-CM5256 × 128362 × 2921950–2005
7FGOALS-s2128 × 108360 × 1961950–2005
8GFDL-CM3144 × 90360 × 2001860–2005
9GFDL-ESM2G144 × 90360 × 2101861–2005
10IPSL-CM5A-LR96 × 96182 × 1491850–2005
11IPSL-CM5A-MR144 × 143182 × 1491850–2005
12IPSL-CM5B-LR96 × 96182 × 1491850–2005
13MIROC5256 × 128256 × 2241850–2005
14MIROC-ESM128 × 64256 × 1921950–2005
15MIROC-ESM-CHEM128 × 64256 × 1921950–2005
16MPI-ESM-LR192 × 96256 × 2201950–2005
17MPI-ESM-MR192 × 96802 × 4041950–2005
18MPI-ESM-P192 × 96256 × 2201950–2005
19MRI-CGCM3320 × 160360 × 3681950–2005
20MRI-ESM1320 × 160360 × 3681950–2005
21NorESM1-M144 × 96320 × 3841950–2005
Table 2. CMIP6 models used in this study.
Table 2. CMIP6 models used in this study.
No.CMIP6 ModelsResolution for
Atmospheric Variables (Lon × Lat)
Model TopResolution for SST (Lon × Lat)Time Period
1ACCESS-CM2192 × 14485 km (High-top)360 × 3001950–2014
2BCC-ESM1128 × 642.19 hPa (Low-top)360 × 2321950–2014
3CanESM5128 × 641 hPa (High-top)361 × 2901850–2014
4CESM2288 × 1922.25 hPa (Low-top)320 × 3841850–2014
5CESM2-FV2144 × 962.25 hPa (Low-top)320 × 3841850–2014
6CESM2-WACCM288 × 1924.5 × 10–6 hPa (High-top)320 × 3841850–2014
7EC-Earth3512 × 2560.01 hPa (High-top)362 × 2921850–2014
8GFDL-CM4360 × 18010 hPa (Low-top)1440 × 10801850–2014
9IPSL-CM6A-LR144 × 1430.01 hPa (High-top)362 × 3321850–2014
10MPI-ESM-1-2-HAM192 × 960.01 hPa (High-top)256 × 2201850–2014
11MPI-ESM1-2-HR384 × 1920.01 hPa (High-top)802 × 4041850–2014
12MPI-ESM1-2-LR192 × 960.01 hPa (High-top)256 × 2201850–2014
13MRI-ESM2-0320 × 1600.01 hPa (High-top)360 × 3641950–2014
14NorESM2-LM144 × 963 hPa (Low-top)360 × 3841950–2014
15NorESM2-MM288 × 1923 hPa (Low-top)360 × 3841950–2014
Table 3. Numbers of ENSO events and SSW events in different ENSO groups in CMIP5 and CMIP6 models.
Table 3. Numbers of ENSO events and SSW events in different ENSO groups in CMIP5 and CMIP6 models.
ModelsENSO MagnitudeWinter SST Anomalies in Niño3.4 RegionWinter NumbersSSW NumbersSSW Frequency
CMIP6Strong El Niño2.32158970.61
Moderate El Niño1.192991790.60
Strong La Niña–2.01108610.56
Moderate La Niña–1.263561870.53
Neutral10396180.59
All196011420.58
CMIP5Strong El Niño1.61155520.34
Moderate El Niño0.80273940.34
Strong La Niña–1.5474250.34
Moderate La Niña–1.01271800.30
Neutral9612560.27
All17345070.29
Table 4. Numbers of ENSO events and SSW events in different ENSO groups in CMIP6 high-top and low-top models.
Table 4. Numbers of ENSO events and SSW events in different ENSO groups in CMIP6 high-top and low-top models.
ModelsENSO MagnitudeWinter NumbersSSW NumbersSSW Frequency
High-topStrong El Niño101690.68
Moderate El Niño1881370.73
Strong La Niña66420.64
Moderate La Niña2341440.62
Neutral6874570.67
All12768490.67
Low-topStrong El Niño57280.49
Moderate El Niño111420.38
Strong La Niña42190.45
Moderate La Niña122430.35
Neutral3521610.46
All6842930.43
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, J.; Shen, Y.; Deng, J.; Jia, Y.; Wang, Z.; Li, A. Revisiting the Influence of ENSO on the Arctic Stratosphere in CMIP5 and CMIP6 Models. Atmosphere 2023, 14, 785. https://doi.org/10.3390/atmos14050785

AMA Style

Hu J, Shen Y, Deng J, Jia Y, Wang Z, Li A. Revisiting the Influence of ENSO on the Arctic Stratosphere in CMIP5 and CMIP6 Models. Atmosphere. 2023; 14(5):785. https://doi.org/10.3390/atmos14050785

Chicago/Turabian Style

Hu, Jinggao, Yifan Shen, Jiechun Deng, Yanpei Jia, Zixu Wang, and Anqi Li. 2023. "Revisiting the Influence of ENSO on the Arctic Stratosphere in CMIP5 and CMIP6 Models" Atmosphere 14, no. 5: 785. https://doi.org/10.3390/atmos14050785

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop