Next Article in Journal
Diverse Techniques in Estimating Integrated Water Vapor for Calibration and Validation of Satellite Altimetry
Previous Article in Journal
Land8Fire: A Complete Study on Wildfire Segmentation Through Comprehensive Review, Human-Annotated Multispectral Dataset, and Extensive Benchmarking
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Decadal Breakdown of Northeast Pacific SST–Arctic Stratospheric Ozone Coupling

1
College of Automation, Nanjing University of Information Science and Technology, Nanjing 210044, China
2
Institute of Meteorology and Oceanography, National University of Defense Technology, Changsha 410073, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2025, 17(16), 2777; https://doi.org/10.3390/rs17162777
Submission received: 29 May 2025 / Revised: 21 July 2025 / Accepted: 27 July 2025 / Published: 11 August 2025

Abstract

Using multiple reanalysis datasets, this study investigates the decadal variability in the relationship between Northeast Pacific Sea surface temperature (SST) and Arctic stratospheric ozone (ASO), with a focus on the role of atmospheric dynamics in mediating this connection. A significant decadal shift is identified around the year 2000, characterized by a weakening of the previously strong negative correlation between January–February SST anomalies and February–March ASO. Prior to 2000 (1980–2000), warm SST in the northeastern Pacific suppressed upward planetary wave propagation, resulting in decreased stratospheric wave activity and a weakened Brewer–Dobson circulation. The weakened BD circulation reduced poleward transport of tropical ozone and heat, yielding a colder, ozone-poor polar vortex. The strong relationship enabled skillful seasonal predictability of ASO using SST precursors in a linear regression model. However, post-2000 (2001–2022), the weakened planetary wave response to SST anomalies resulted in a breakdown of this relationship, yielding non-significant predictive skill. The findings highlight the non-stationary nature of ocean-stratosphere coupling and underscore the importance of accounting for such decadal shifts in climate models to improve projections of Arctic ozone recovery and its surface climate impacts.

Graphical Abstract

1. Introduction

Stratospheric ozone plays a dual critical role in the earth system, serving both as a protective shield against harmful ultraviolet radiation for life [1,2] and as a key modulator of global climate through radiative-dynamic coupling [3]. This coupling mechanism enables stratospheric ozone variability to exert downward influences on surface weather extremes [4,5,6,7,8,9,10,11,12,13,14,15]. For instance, the recent record extreme hot Australian summer of 2012/13 was linked to anomalously high ozone levels in the preceding November [8], while ozone depletion has been associated with the intensification of austral summer precipitation extremes [7]. Given the significant impact of stratospheric ozone on Earth’s climate, numerous studies have investigated its variability and explored the factors and processes that influence ozone concentrations [16,17,18,19].
Extensive research has established that stratospheric ozone variability is governed by both anthropogenic forcing and natural climate variability. Key anthropogenic factors include ozone-depleting substances (ODS) [20,21,22] and greenhouse gases (GHGs) [21,23], while natural modes of variability encompass the El Niño-Southern Oscillation (ENSO) [17,24,25], Quasi-Biennial Oscillation (QBO) [26,27], Indian Ocean Dipole (IOD) [28,29], and Pacific Decadal Oscillation (PDO) [30]. The pronounced global ozone decline observed prior to 1998 was primarily attributable to anthropogenic emissions. Specifically, ODS undergo photo-dissociation upon reaching the stratosphere under intense UV radiation, releasing chlorine and bromine radicals that catalytically destroy ozone molecules through chain reactions [31,32]. The implementation of the Montreal Protocol successfully averted uncontrolled global ozone depletion and consequent dramatic increases in surface UV-B radiation [33,34]. In recent years, ozone has begun to recover, but the trends in ozone concentration vary by altitude. While the upper stratosphere has demonstrated robust recovery trends [35,36], the lower stratosphere (60°S–60°N) has exhibited persistent ozone decreases since 1998 [37,38]. These complex spatial patterns likely reflect the dominant influence of internal climate variability. For instance, ENSO modulates tropical lower stratospheric ozone concentrations, particularly during boreal summer [39]. During western QBO phases, double peaks of positive ozone anomalies are observed in the tropics, accompanied by negative anomalies in the extra-tropics [27].
Although Arctic stratospheric ozone depletion is not as severe as in Antarctica, the variability in ozone loss has always been much greater in the Arctic due to variability in wave forcing [40]. It is well known that Arctic ozone has been depleting, but since the signing of the Montreal Protocol, it began to slowly recover after the 1990s. It is projected that Arctic ozone will return to 1980 levels by 2040 [41]. In addition to trend changes, inter-annual variability has been a major focus of current research and discussions, as recent studies have shown that inter-annual variations in Arctic ozone significantly influence surface weather and climate in the Northern Hemisphere [9,10,11,12,13,42,43,44,45,46]. Understanding the inter-annual variability of Arctic ozone and its potential influencing factors is of great practical importance. The inter-annual variations in polar ozone are intrinsically linked to stratospheric polar vortex dynamics. Previous studies have extensively examined the relationship between the Arctic polar vortex and Pacific sea surface temperature (SST) [47,48,49,50,51]. For instance, weak stratospheric polar vortex events occur more frequently during positive PDO phases compared to its negative phases [50,51], with approximately 25% of the polar vortex strength attributable to central North Pacific SST [48]. However, research on ozone-SST relationships remains relatively scarce, with limited studies suggesting that North Pacific warming contributes to Arctic lower stratospheric ozone depletion [52,53]. While some researchers pointed out that negative SST anomalies in the central North Pacific could enhance the upward propagation of planetary wavenumbers 1 and 2, weakening the Brewer–Dobson circulation and subsequently reducing ozone concentrations in the Arctic lower stratosphere [54].
Despite previous studies having explored Arctic stratospheric ozone variability and its potential drivers, few have discussed the decadal relationship between Arctic ozone concentration and SST in the Pacific. Recently, Wang et al. (2023) examined decadal changes in the connection between Arctic stratospheric ozone and SST responses [55]. It is worthwhile to explore whether there is a decadal relationship between Arctic stratospheric ozone and preceding SST. This study aims to characterize the decadal relationship between Arctic stratospheric ozone and Pacific SST variability, to identify key connection regions, and to elucidate the underlying physical mechanisms governing these relationships.

2. Data and Methods

This study employs three ozone datasets to ensure robust analysis: the monthly Stratospheric Water and Ozone Satellite Homogenized (SWOOSH) dataset at 2.5° longitude resolution, which provides a merged record of stratospheric ozone and water vapor measurements from limb sounding and solar occultation satellites and comprises data from the SAGE-II/III, UARS HALOE, UARS MLS, and Aura MLS instruments spanning from 1984 to present [56]; the monthly Modern-Era Retrospective analysis for Research and Applications, Version 2 (MERRA-2) ozone dataset with a uniform 2.5° latitude × 2.5° longitude spatial grid, produced by NASA’s Global Modeling and Assimilation Office (GMAO), which assimilates Solar Backscatter Ultraviolet Radiometer (SBUV) partial column ozone retrievals (1980–2004) and later incorporates Aura Microwave Limb Sounder (MLS) ozone profiles and Ozone Monitoring Instrument total column ozone (2004–present) [57]; and the monthly fifth-generation European Centre for Medium-Range Weather Forecasts (ERA5) reanalysis provided at 2° latitude × 2° longitude resolution, which uses Level 2 ozone from multiple instruments [58]. Additionally, atmospheric variables—including geopotential height, wind, and temperature—are also derived from ERA5 with a uniform 1° latitude × 1° longitude spatial grid, while monthly SST data are obtained from the Met Office Hadley Centre’s HadISST dataset on a 1° latitude × 1° longitude global grid [59]. In this study, we mainly focus on the period of 1980–2022.
The propagation of wave activity uses the following formulas according to Andrews et al. (1987) [60]. The meridional (Fy) and vertical (Fz) components of the Eliassen–Palm (EP) flux and the EP flux divergence (DF) are expressed as follows:
F y = ρ 0 a cos φ ( u ¯ z v θ ¯ / θ ¯ z u v ¯ )
F z = ρ 0 a cos φ { [ f a cos φ 1 u ¯ cos φ φ ] v θ ¯ / θ ¯ z }
D F = · F ρ 0 a cos φ = ( F y cos φ ) / a cos φ φ + F z / z ρ 0 a cos φ
where ρ0 is the density of air; a is the radius of Earth; φ is the latitude; f is the Coriolis parameter; u and υ are the zonal and meridional components of the wind, respectively; and θ is the potential temperature. The overline denotes the zonal mean. Superscript symbols indicate deviations from the zonal mean.
The EP flux indicates the direction of wave propagation, quantifying planetary wave activity in the stratosphere. Increased wave propagation from the troposphere causes stronger EP flux convergence in the stratosphere. The EP flux convergence drives a stronger Brewer–Dobson circulation via downward control arguments.
We employed standard linear regression models to examine the relationships between preceding SST and ASO. To determine the robustness of our linear model in predicting ASO and sidestep the issue of data overfitting. A leave-one-out cross-validation method is employed [9,61]. Specifically, a time point is selected from a time series of length N as a hindcasting target point, while the remaining (N–1) points are treated as training points to develop the linear model. Repeat the above procedure until each point in the time series has served as the hindcasting target point. Consequently, a set of N predictive models will be constructed, culminating in an ensemble hindcast.
Throughout our study, we exclusively employed Pearson’s correlation coefficients to assess linear relationships between variables. Given the relatively small sample size in our analysis, we employed a nonparametric bootstrap approach to assess the statistical significance of Pearson’s correlation coefficients [62]. We performed 1000 bootstrap resamples to generate robust confidence intervals while maintaining computational efficiency. The strength of correlation was determined by examining the 95% confidence interval derived from these resampled coefficients: if the interval excluded zero and the original correlation coefficient fell within this range, the correlation was considered statistically significant (p < 0.05, strong correlation coefficients). Conversely, if the 95% confidence interval included zero, the correlation was deemed weak or moderate.

3. The Weakening Relationship Between the Arctic Ozone and Northeast Pacific SST Since the 2000s

To investigate the ozone variability, ozone concentrations in ppmv and Dobson Units (DU) derived from the SWOOSH dataset are analyzed. Figure 1 presents the climatology of zonal-mean ozone mixing ratios (ppmv) and their corresponding DU values (Figure 1a,b). The results for the ozone mixing ratio (ppmv) reveal a pronounced maximum centered in the tropical stratosphere near 10 hPa, with values decreasing progressively toward higher latitudes and altitudes (Figure 1a). Notably, the standard deviation is higher in the lower tropical stratosphere and the lower stratosphere in the polar regions (Figure 1c). Conversion to DU yields distinct spatial patterns compared to mixing ratios. The DU distribution exhibits an arch-like structure, with peak ozone occurring in the Arctic lower stratosphere (Figure 1b). Maximum variability is observed between 30 and 150 hPa over both polar regions (Figure 1d). These findings demonstrate that the Arctic lower stratosphere contains not only the highest DU-based ozone but also exhibits substantial variability. This large Arctic ozone variability has important implications, as growing evidence suggests that Arctic stratospheric ozone anomalies can significantly influence surface weather and climate across northern mid-to-high latitudes [9,10,11,12,13,42,43,44,45,46].
To further characterize the evolution of Arctic stratospheric ozone, we define an Arctic Stratospheric Ozone (ASO) index. This index is computed as the area-weighted average over the polar cap region (60–90°N) and incorporates mass-weighting through vertical integration across the 1–150 hPa pressure. Figure 2a presents the time series of ASO anomalies derived from three independent datasets (SWOOSH, MERRA2, and ERA5), which shows remarkable consistency. The records reveal a pronounced ozone depletion trend prior to 1997, consistent with increasing atmospheric concentrations of ODS during this period, followed by a stabilization after the implementation of the Montreal Protocol. To exclude the influences from the long-term ODS-driven trend, Figure 2b displays the detrended ASO anomalies, which exhibit substantial seasonal variability. To further investigate these seasonal differences, Figure 2c presents the standard deviation of the detrended ASO anomalies from January to December. The analysis reveals maximum variability during winter months (December–March), with peak values occurring in February and March. This seasonal variability likely reflects enhanced dynamical variability associated with a stronger polar vortex during winter in early spring [63]. The pronounced winter/spring variability carries particular significance, as numerous studies have demonstrated that late winter/early spring Arctic ozone anomalies can influence tropospheric temperature and East Asian precipitation variability [12,13].
Figure 3 presents the February–March ASO anomaly time series. Similarly to the annual dataset, a significant ozone depletion trend is evident before 1997, while the post-1997 period shows continued but weaker depletion. The detrended anomalies in Figure 3b exhibit pronounced inter-annual variability, with amplitude exceeding 20 DU during extreme years. Previous studies have established decadal linkage between ozone variability and SSTs [55]. They have documented a weakened connection between March Arctic ozone and April North Pacific SST since 2000. This raises the question of whether the influence of preceding SST on Arctic ozone also exhibits inter-decadal variations. To address this issue, and considering that the impact of SST on the polar circulation may have a lead of about one month, Figure 3c,d displays the correlation between the detrended February–March ASO anomalies and January–February SST anomalies, with the pre-2000 period (1980–2000) showing a significant negative correlation over the Northeast Pacific, while no significant positive correlation is observed post-2000. We further examine the relationship through 21-year running correlation between January–February SST and February–March ASO (Figure 3e), which reveals a significant shift in their correlation structure around the year 2000. This suggests that the influence of SST on late winter or early spring ozone also exhibits decadal changes. To further analyze this relationship, the time series of detrended February–March ASO anomalies and January–February SST anomalies, averaged over 20–40°N and 105–127°W (SST_NEP), are shown in Figure 4a,b. It is clear that they are significantly negatively correlated before 2000, with a correlation coefficient of −0.57, while after 2000, no significant correlation is observed. Figure 4c,d present scatter plots of standardized detrended ASO versus SST_NEP anomalies. Similarly to Figure 4a,b, SST_NEP and ASO show a significant negative correlation before 2000, but no significant linear correlation is found post-2000.

4. Diagnosing the Possible Mechanisms Behind the Decadal Changes

To further investigate the physical processes through which SST influences ASO variability, Figure 5 presents the correlation analyses between the SST_NEP index and zonal-mean zonal winds at 500 hPa and 200 hPa for both pre- and post-2000 periods. During the pre-2000 era, significant negative correlations prevailed across mid-latitudes at both pressure levels, indicating a pronounced weakening of westerly winds that would have suppressed vertical propagation of planetary waves into the stratosphere (Figure 5a,c). This large-scale circulation response would favor a stronger, more persistent polar vortex, creating conditions that inhibit the meridional transport of ozone from tropical to polar regions, consistent with the strong negative ASO-SST relationships observed during this era. In contrast, the post-2000 correlations exhibit a markedly different pattern, with negative correlations confined primarily to the North Pacific sector while positive correlations emerge across other mid-latitude regions. Such changes in the circulation field likely have a weaker impact on wave propagation, corresponding with the documented weakening of ASO-SST relationships (Figure 4b). These results demonstrate that, while pre-2000 SST variability exerted a strong hemispheric-scale influence on circulation patterns, significantly affecting Arctic ozone, the post-2000 regime has shifted to a state where SST forcing produces more limited and localized dynamical responses, with a reduced potential to influence stratospheric ozone variability.
Furthermore, to better understand the process through which planetary waves forced by tropical SST anomalies modulate polar ozone, we employed EP flux to examine stratospheric wave activity propagation. The stratospheric polar vortex variability is intrinsically linked to vertically propagating planetary waves originating from tropospheric sources. Figure 6 presents daily computed planetary wave responses between positive and negative SST_NEP events, revealing distinct decadal changes around 2000. During the pre-2000 period (Figure 6a), we observed reduced upward wave propagation at 60°N accompanied by EP flux divergence in the high-latitude stratosphere (10–30 hPa), a configuration that would cool and strengthen the northern polar vortex by reducing Brewer–Dobson circulation-induced adiabatic heating. In contrast, the post-2000 period shows weaker EP flux convergence in the polar stratosphere, with enhanced equatorward EP flux transport from high latitudes but no significant changes in vertical wave (Figure 6b). Further decomposition into stationary and transient wave components (Figure 7) reveals more complex dynamics. The pre-2000 period demonstrates fewer planetary waves transported upward into the stratosphere in both stationary and transient wave at mid-latitudes, though with opposing signs in EP flux divergence (Figure 7a,c), indicating that the stratospheric EP flux propagation and divergence were jointly influenced by both wave types. After 2000, similar to the overall changes in EP flux (Figure 6b), the EP flux did not exhibit significant vertical changes, while the divergence (Figure 7b) and convergence (Figure 7d) counteracted each other, leading to a weaker response in the total EP flux divergence (Figure 6b). To quantitatively assess the relative contributions of stationary waves and transient waves, we calculated the vertical flux averaged over the 45–75°N region at 50 hPa. The results show that during 1980–2000, stationary waves accounted for 68% of the total wave activity, while transient waves contributed 32%. In contrast, during 2001–2022, stationary waves only accounted for 38%, with transient waves dominating at 62%. Notably, the amplitude of vertical wave activity post-2000 was significantly weaker, measuring only 13% of the pre-2000 level. These analyses demonstrate that January–February Northeast Pacific SST anomalies can significantly modulate stratosphere-entering planetary waves through combined stationary and transient wave effects prior to 2000, while the post-2000 period exhibits markedly weaker planetary wave responses to SST variability.
These results demonstrate distinct decadal shifts in the Northeast Pacific SST’s influence on stratospheric dynamics. During the pre-2000 period, positive SST anomalies in the northwestern Pacific effectively suppressed upward planetary wave propagation (Figure 6a), leading to a cooler and stronger polar vortex (Figure 8a,c). This dynamical response established atmospheric conditions that effectively suppressed the poleward transport of both ozone and heat from tropical regions. The chain of physical processes—from suppressed EP flux (Figure 6a) to vortex strengthening (Figure 8c) and ultimately to ozone loss—explains the robust negative ASO-SST relationships observed during this era. In contrast, the post-2000 period shows markedly different behavior. The weakened planetary wave response to SST variations (Figure 6b) corresponds with non-significant changes in stratospheric temperatures and circulation patterns (Figure 8b,d,f). This decoupling suggests fundamental changes in the ocean-atmosphere-stratosphere interaction pathways, where Northeast Pacific SST anomalies no longer effectively modulate the stratospheric polar vortex through wave-mean flow interactions. The absence of coherent dynamical responses helps explain the breakdown of ASO-SST relationships in recent decades. These findings highlight the importance of considering non-stationarity in stratosphere–troposphere coupling when assessing long-term ozone variability and its climate connections.
To quantitatively assess the predictive potential of northwestern Pacific SST anomalies for Arctic ozone variability, we developed a linear regression model with January–February SST_NEP as the predictor for February–March polar ozone intensity. The pre-2000 period demonstrates robust predictive skill, with the full-period (1980–2000) regression model yielding a significant correlation coefficient of 0.57 (p < 0.05) between fitted and observed ASO indices (Figure 9a). The model’s stability was further verified through a split-sample approach: when trained on 1980–1994 data, the model maintained strong skill with a correlation of 0.56 (p < 0.05) for the training period and even higher skill (r = 0.62, not significant due to small sample) during the independent test period (1995–2000, Figure 9c), demonstrating the temporal robustness of this relationship. In stark contrast, the post-2000 period shows markedly reduced predictive capability. The full-period (2001–2022) regression model achieves only marginal skill (r = 0.24, Figure 9b), while the split-period analysis reveals similarly limited performance (training period 2001–2016: r = 0.24; prediction period 2017–2022: r = 0.33). As shown in Figure 9e,f, the observed ASO index (blue line, MERRA-2) and predicted values based on the leave-one-out cross-validation method (green line) also demonstrate strong agreement during 1980–2000, with a statistically significant correlation coefficient of 0.47. However, this relationship becomes insignificant during 2001–2022. These findings are fully consistent with the results shown in Figure 9a–d. This substantial decline in predictive skill corresponds with the documented weakening of dynamical connections between Pacific SST and Arctic stratospheric variability, further confirming the decadal shift in stratosphere–troposphere coupling around 2000.

5. Conclusions and Discussion

This study analyzed the influence of Northeast Pacific SST anomalies from late winter (January–February) on the inter-annual variations in Arctic stratospheric ozone in late winter and early spring (February–March) using observations. It was found that there is a weakening in the connection between Northeast Pacific SST and ASO, marked by a distinct shift around 2000. Prior to 2000, the January–February SST are strong negative correlated with February–March ASO anomalies (r = −0.57), while this relationship became statistically insignificant in the subsequent two decades. The weakening of this linkage likely stems from significant alterations in atmospheric wave propagation, which modulate stratospheric ozone concentrations.
The underlying physical mechanism involves SST-induced modulation of stratospheric wave driving. In the pre-2000 period, warm SST anomalies in the northwestern Pacific effectively suppressed upward planetary wave propagation, weakening Brewer–Dobson circulation and reducing poleward transport of tropical ozone and heat. These dynamical processes resulted in robust negative correlations between January–February SSTs and subsequent February–March ASO anomalies, enabling skillful seasonal predictability in regression models. However, post-2000, this teleconnection weakened significantly due to altered atmospheric wave responses, with SST anomalies producing only localized circulation changes that failed to coherently modulate stratospheric dynamics. Consequently, the predictive power of SST precursors diminished, highlighting the non-stationary nature of ocean-stratosphere coupling. The observed shift in SST-ASO relationships, potentially linked to PDO or anthropogenic climate change, carries important implications for understanding Arctic ozone changes and improving seasonal forecasting frameworks in a changing climate. The breakdown of the previously robust relationship suggests that climate models capable of capturing such non-stationary behavior may achieve more accurate projections of future stratospheric ozone trends and their impacts on surface climate.

Author Contributions

Conceptualization, Q.L.; methodology, T.C.; software, T.C.; validation, T.C.; formal analysis, Q.L. and T.C.; investigation, Q.L. and T.C.; writing—original draft preparation, T.C.; writing—review and editing, Q.L. and T.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The datasets analyzed in this study are all publicly available from their sources: stratospheric ozone data were obtained from the SWOOSH dataset (available at https://csl.noaa.gov/groups/csl8/swoosh) and MERRA-2 reanalysis (accessible via https://climatedataguide.ucar.edu/climate-data/nasas-merra2-reanalysis), sea surface temperature data were acquired from the Hadley Centre HadISST dataset (https://www.metoffice.gov.uk/hadobs/hadisst), and meteorological fields were derived from the ERA5 reanalysis (https://www.ecmwf.int/en/forecasts/dataset/ecmwf-reanalysis-v5). All datasets were accessed on 9 May 2025.

Acknowledgments

We acknowledge the following data sources: the SWOOSH merged satellite ozone dataset, the MERRA-2 reanalysis, sea surface temperatures from Hadley Centre, and ERA5meteorological fields.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Madronich, S.; Lee-Taylor, J.M.; Wagner, M.; Kyle, J.; Hu, Z.; Landolfi, R. Estimation of Skin and Ocular Damage Avoided in the United States through Implementation of the Montreal Protocol on Substances That Deplete the Ozone Layer. ACS Earth Space Chem. 2021, 5, 1876–1888. [Google Scholar] [CrossRef] [PubMed]
  2. Parker, E.R. The Influence of Climate Change on Skin Cancer Incidence—A Review of the Evidence. Int. J. Dermatol. 2021, 7, 17–27. [Google Scholar] [CrossRef]
  3. Saurral, R.I.; Kucharski, F.; Raggio, G.A. Variations in ozone and greenhouse gases as drivers of Southern Hemisphere climate in a medium-complexity global climate model. Clim. Dyn. 2019, 53, 6645–6663. [Google Scholar] [CrossRef]
  4. Lee, S.; Feldstein, S.B. Detecting Ozone- and Greenhouse Gas–Driven Wind Trends with Observational Data. Science 2013, 339, 563–567. [Google Scholar] [CrossRef]
  5. Gonzalez, P.L.M.; Polvani, L.M.; Seager, R.; Correa, G.J.P. Stratospheric Ozone Depletion: A Key Driver of Recent Precipitation Trends in South Eastern South America. Clim. Dyn. 2014, 42, 1775–1792. [Google Scholar] [CrossRef]
  6. Previdi, M.; Polvani, L.M. Climate System Response to Stratospheric Ozone Depletion and Recovery. Q. J. R. Meteorol. Soc. 2014, 140, 2401–2419. [Google Scholar] [CrossRef]
  7. Kang, S.M.; Polvani, L.M.; Fyfe, J.C.; Son, S.-W.; Sigmond, M.; Correa, G.J.P. Modeling Evidence That Ozone Depletion Has Impacted Extreme Precipitation in the Austral Summer. Geophys. Res. Lett. 2013, 40, 4054–4059. [Google Scholar] [CrossRef]
  8. Bandoro, J.; Solomon, S.; Donohoe, A.; Thompson, D.W.J.; Santer, B.D. Influences of the Antarctic Ozone Hole on Southern Hemispheric Summer Climate Change. J. Clim. 2014, 27, 6245–6264. [Google Scholar] [CrossRef]
  9. Ma, X.; Xie, F.; Chen, X.; Wang, L.; Yang, G. Identifying a Leading Predictor of Arctic Stratospheric Ozone for April Precipitation in Eastern North America. Remote Sens. 2022, 14, 5040. [Google Scholar] [CrossRef]
  10. Ma, X.; Xie, F.; Li, J.; Zheng, X.; Tian, W.; Ding, R.; Sun, C.; Zhang, J. Effects of Arctic Stratospheric Ozone Changes on Spring Precipitation in the Northwestern United States. Atmos. Chem. Phys. 2019, 19, 861–875. [Google Scholar] [CrossRef]
  11. Ma, X.; Xie, F. Predicting April Precipitation in the Northwestern United States Based on Arctic Stratospheric Ozone and Local Circulation. Front. Earth Sci. 2020, 8, 56. [Google Scholar] [CrossRef]
  12. Xie, F.; Ma, X.; Li, J.; Huang, J.; Tian, W.; Zhang, J.; Hu, Y.; Sun, C.; Zhou, X.; Feng, J.; et al. An Advanced Impact of Arctic Stratospheric Ozone Changes on Spring Precipitation in China. Clim. Dyn. 2018, 51, 4029–4041. [Google Scholar] [CrossRef]
  13. Xie, F.; Ma, X.; Li, J.; Tian, W.; Ruan, C.; Sun, C. Using Observed Signals from the Arctic Stratosphere and Indian Ocean to Predict April–May Precipitation in Central China. J. Clim. 2020, 33, 131–143. [Google Scholar] [CrossRef]
  14. Friedel, M.; Chiodo, G.; Stenke, A.; Domeisen, D.I.V.; Fueglistaler, S.; Anet, J.G.; Peter, T. Springtime Arctic Ozone Depletion Forces Northern Hemisphere Climate Anomalies. Nat. Geosci. 2022, 15, 541–547. [Google Scholar] [CrossRef]
  15. Gillett, Z.E.; Arblaster, J.M.; Dittus, A.J.; Deushi, M.; Jöckel, P.; Kinnison, D.E.; Morgenstern, O.; Plummer, D.A.; Revell, L.E.; Rozanov, E.; et al. Evaluating the Relationship between Interannual Variations in the Antarctic Ozone Hole and Southern Hemisphere Surface Climate in Chemistry–Climate Models. J. Clim. 2019, 32, 3131–3151. [Google Scholar] [CrossRef]
  16. Bahramvash Shams, S.; Walden, V.P.; Hannigan, J.W.; Randel, W.J.; Petropavlovskikh, I.V.; Butler, A.H.; De La Cámara, A. Analyzing Ozone Variations and Uncertainties at High Latitudes during Sudden Stratospheric Warming Events Using MERRA-2. Atmos. Chem. Phys. 2022, 22, 5435–5458. [Google Scholar] [CrossRef]
  17. Lu, J.; Xie, F.; Tian, W.; Li, J.; Wuhu, F.; Chipperfield, M.; Zhang, J.; Ma, X. Interannual Variations in Lower Stratospheric Ozone During the Period 1984–2016. J. Geophys. Res. 2019, 124, 8225–8241. [Google Scholar] [CrossRef]
  18. Dietmüller, S.; Garny, H.; Eichinger, R.; Ball, W.T. Analysis of Recent Lower-Stratospheric Ozone Trends in Chemistry Climate Models. Atmos. Chem. Phys. 2021, 21, 6811–6837. [Google Scholar] [CrossRef]
  19. Randel, W.J.; Wu, F. A Stratospheric Ozone Profile Data Set for 1979–2005: Variability, Trends, and Comparisons with Column Ozone Data. J. Geophys. Res. 2007, 112, 2006JD007339. [Google Scholar] [CrossRef]
  20. Weber, M.; Arosio, C.; Coldewey-Egbers, M.; Fioletov, V.E.; Frith, S.M.; Wild, J.D.; Tourpali, K.; Burrows, J.P.; Loyola, D. Global Total Ozone Recovery Trends Attributed to Ozone-Depleting Substance (ODS) Changes Derived from Five Merged Ozone Datasets. Atmos. Chem. Phys. 2022, 22, 6843–6859. [Google Scholar] [CrossRef]
  21. Waugh, D.W.; Oman, L.; Kawa, S.R.; Stolarski, R.S.; Pawson, S.; Douglass, A.R.; Newman, P.A.; Nielsen, J.E. Impacts of Climate Change on Stratospheric Ozone Recovery. Geophys. Res. Lett. 2009, 36, 2008GL036223. [Google Scholar] [CrossRef]
  22. Chipperfield, M.P.; Bekki, S.; Dhomse, S.; Harris, N.R.P.; Hassler, B.; Hossaini, R.; Steinbrecht, W.; Thiéblemont, R.; Weber, M. Detecting Recovery of the Stratospheric Ozone Layer. Nature 2017, 549, 211–218. [Google Scholar] [CrossRef]
  23. Plummer, D.A.; Scinocca, J.F.; Shepherd, T.G.; Reader, M.C.; Jonsson, A.I. Quantifying the Contributions to Stratospheric Ozone Changes from Ozone Depleting Substances and Greenhouse Gases. Atmos. Chem. Phys. 2010, 10, 8803–8820. [Google Scholar] [CrossRef]
  24. Oman, L.D.; Douglass, A.R.; Ziemke, J.R.; Rodriguez, J.M.; Waugh, D.W.; Nielsen, J.E. The Ozone Response to ENSO in Aura Satellite Measurements and a Chemistry-climate Simulation. J. Geophys. Res. 2013, 118, 965–976. [Google Scholar] [CrossRef]
  25. Benito-Barca, S.; Calvo, N.; Abalos, M. Driving Mechanisms for the El Niño–Southern Oscillation Impact on Stratospheric Ozone. Atmos. Chem. Phys. 2022, 22, 15729–15745. [Google Scholar] [CrossRef]
  26. Olsen, M.A.; Manney, G.L.; Liu, J. The ENSO and QBO Impact on Ozone Variability and Stratosphere-Troposphere Exchange Relative to the Subtropical Jets. J. Geophys. Res. 2019, 124, 7379–7392. [Google Scholar] [CrossRef]
  27. Wang, W.; Hong, J.; Shangguan, M.; Wang, H.; Jiang, W.; Zhao, S. Zonally Asymmetric Influences of the Quasi-Biennial Oscillation on Stratospheric Ozone. Atmos. Chem. Phys. 2022, 22, 13695–13711. [Google Scholar] [CrossRef]
  28. Krzyścin, J.W. El Niño-Southern Oscillation and Indian Ocean Dipole Contribution to the Zonal Mean Total Ozone in the Northern Hemisphere. Int. J. Climatol. 2017, 37, 3517–3524. [Google Scholar] [CrossRef]
  29. Nath, O.; Singh, B.B.; Kunchala, R.K. El Niño Southern Oscillation (ENSO) and Indian Ocean Dipole (IOD) Signatures in Tropical Ozone in the Upper Troposphere Lower Stratosphere (UTLS). Meteorol. Atmos. Phys. 2024, 136, 10. [Google Scholar] [CrossRef]
  30. Gamelin, B.L.; Carvalho, L.M.V.; Kayano, M. The Combined Influence of ENSO and PDO on the Spring UTLS Ozone Variability in South America. Clim. Dyn. 2020, 55, 1539–1562. [Google Scholar] [CrossRef]
  31. Molina, M.J.; Rowland, F.S. Stratospheric Sink for Chlorofluoromethanes: Chlorine Atom-Catalyzed Destruction of Ozone. Nature 1974, 249, 810–812. [Google Scholar] [CrossRef]
  32. Solomon, P.; Barrett, J.; Mooney, T.; Connor, B.; Parrish, A.; Siskind, D.E. Rise and Decline of Active Chlorine in the Stratosphere. Geophys. Res. Lett. 2006, 33, 2006GL027029. [Google Scholar] [CrossRef]
  33. Mäder, J.A.; Staehelin, J.; Peter, T.; Brunner, D.; Rieder, H.E.; Stahel, W.A. Evidence for the Effectiveness of the Montreal Protocol to Protect the Ozone Layer. Atmos. Chem. Phys. 2010, 10, 12161–12171. [Google Scholar] [CrossRef]
  34. Newman, P.A.; McKenzie, R. UV Impacts Avoided by the Montreal Protocol. Photochem. Photobiol. Sci. 2011, 10, 1152–1160. [Google Scholar] [CrossRef]
  35. Steinbrecht, W.; Froidevaux, L.; Fuller, R.; Wang, R.; Anderson, J.; Roth, C.; Bourassa, A.; Degenstein, D.; Damadeo, R.; Zawodny, J.; et al. An Update on Ozone Profile Trends for the Period 2000 to 2016. Atmos. Chem. Phys. 2017, 17, 10675–10690. [Google Scholar] [CrossRef]
  36. Tummon, F.; Hassler, B.; Harris, N.R.P.; Staehelin, J.; Steinbrecht, W.; Anderson, J.; Bodeker, G.E.; Bourassa, A.; Davis, S.M.; Degenstein, D.; et al. Intercomparison of Vertically Resolved Merged Satellite Ozone Data Sets: Interannual Variability and Long-Term Trends. Atmos. Chem. Phys. 2015, 15, 3021–3043. [Google Scholar] [CrossRef]
  37. Ball, W.T.; Alsing, J.; Mortlock, D.J.; Staehelin, J.; Haigh, J.D.; Peter, T.; Tummon, F.; Stübi, R.; Stenke, A.; Anderson, J.; et al. Evidence for a Continuous Decline in Lower Stratospheric Ozone Offsetting Ozone Layer Recovery. Atmos. Chem. Phys. 2018, 18, 1379–1394. [Google Scholar] [CrossRef]
  38. Wargan, K.; Orbe, C.; Pawson, S.; Ziemke, J.R.; Oman, L.D.; Olsen, M.A.; Coy, L.; Emma Knowland, K. Recent Decline in Extratropical Lower Stratospheric Ozone Attributed to Circulation Changes. Geophys. Res. Lett. 2018, 45, 5166–5176. [Google Scholar] [CrossRef] [PubMed]
  39. Tweedy, O.V.; Waugh, D.W.; Randel, W.J.; Abalos, M.; Oman, L.D.; Kinnison, D.E. The Impact of Boreal Summer ENSO Events on Tropical Lower Stratospheric Ozone. J. Geophys. Res. 2018, 123, 9843–9857. [Google Scholar] [CrossRef]
  40. Müller, R.; Grooß, J.-U.; Lemmen, C.; Heinze, D.; Dameris, M.; Bodeker, G. Simple Measures of Ozone Depletion in the Polar Stratosphere. Atmos. Chem. Phys. 2008, 8, 251–264. [Google Scholar] [CrossRef]
  41. Austin, J.; Wilson, R.J. Ensemble Simulations of the Decline and Recovery of Stratospheric Ozone. J. Geophys. Res. 2006, 111, 2005JD006907. [Google Scholar] [CrossRef]
  42. Harari, O.; Garfinkel, C.I.; Ziskin Ziv, S.; Morgenstern, O.; Zeng, G.; Tilmes, S.; Kinnison, D.; Deushi, M.; Jöckel, P.; Pozzer, A.; et al. Influence of Arctic Stratospheric Ozone on Surface Climate in CCMI Models. Atmos. Chem. Phys. 2019, 19, 9253–9268. [Google Scholar] [CrossRef]
  43. Cheung, J.C.H.; Haigh, J.D.; Jackson, D.R. Impact of EOS MLS Ozone Data on Medium-extended Range Ensemble Weather Forecasts. J. Geophys. Res. 2014, 119, 9253–9266. [Google Scholar] [CrossRef]
  44. Calvo, N.; Polvani, L.M.; Solomon, S. On the Surface Impact of Arctic Stratospheric Ozone Extremes. Environ. Res. Lett. 2015, 10, 094003. [Google Scholar] [CrossRef]
  45. Ivy, D.J.; Solomon, S.; Calvo, N.; Thompson, D.W.J. Observed Connections of Arctic Stratospheric Ozone Extremes to Northern Hemisphere Surface Climate. Environ. Res. Lett. 2017, 12, 024004. [Google Scholar] [CrossRef]
  46. Stone, K.A.; Solomon, S.; Kinnison, D.E.; Baggett, C.F.; Barnes, E.A. Prediction of Northern Hemisphere Regional Surface Temperatures Using Stratospheric Ozone Information. J. Geophys. Res. 2019, 124, 5922–5933. [Google Scholar] [CrossRef]
  47. Hurwitz, M.M.; Newman, P.A.; Garfinkel, C.I. The Arctic Vortex in March 2011: A Dynamical Perspective. Atmos. Chem. Phys. 2011, 11, 11447–11453. [Google Scholar] [CrossRef]
  48. Hu, D.; Guan, Z.; Tian, W.; Ren, R. Recent Strengthening of the Stratospheric Arctic Vortex Response to Warming in the Central North Pacific. Nat. Commun. 2018, 9, 1697. [Google Scholar] [CrossRef]
  49. Hu, D.; Guan, Z. Decadal Relationship between the Stratospheric Arctic Vortex and Pacific Decadal Oscillation. J. Clim. 2018, 31, 3371–3386. [Google Scholar] [CrossRef]
  50. Kren, A.C.; Marsh, D.R.; Smith, A.K.; Pilewskie, P. Wintertime Northern Hemisphere Response in the Stratosphere to the Pacific Decadal Oscillation Using the Whole Atmosphere Community Climate Model. J. Clim. 2016, 29, 1031–1049. [Google Scholar] [CrossRef]
  51. Woo, S.-H.; Sung, M.-K.; Son, S.-W.; Kug, J.-S. Connection between Weak Stratospheric Vortex Events and the Pacific Decadal Oscillation. Clim. Dyn. 2015, 45, 3481–3492. [Google Scholar] [CrossRef]
  52. Hurwitz, M.M.; Newman, P.A.; Garfinkel, C.I. On the Influence of North Pacific Sea Surface Temperature on the Arctic Winter Climate. J. Geophys. Res. 2012, 117, 2012JD017819. [Google Scholar] [CrossRef]
  53. Garfinkel, C.I.; Hurwitz, M.M.; Oman, L.D. Effect of Recent Sea Surface Temperature Trends on the Arctic Stratospheric Vortex. J. Geophys. Res. 2015, 120, 5404–5416. [Google Scholar] [CrossRef]
  54. Liu, M.; Hu, D.; Zhang, F. Connections Between Stratospheric Ozone Concentrations Over the Arctic and Sea Surface Temperatures in the North Pacific. J. Geophys. Res. 2020, 125, e2019JD031690. [Google Scholar] [CrossRef]
  55. Wang, T.; Tian, W.; Lin, Y.; Gou, X.; Liu, H.; Wang, X.; Xie, F. Decadal Changes in the Relationship between Arctic Stratospheric Ozone and Sea Surface Temperatures in the North Pacific. Atmos. Res. 2023, 292, 106870. [Google Scholar] [CrossRef]
  56. Davis, S.M.; Rosenlof, K.H.; Hassler, B.; Hurst, D.F.; Read, W.G.; Vömel, H.; Selkirk, H.; Fujiwara, M.; Damadeo, R. The Stratospheric Water and Ozone Satellite Homogenized (SWOOSH) Database:A Long-Term Database for Climate Studies. Earth Syst. Sci. Data 2016, 8, 461–490. [Google Scholar] [CrossRef] [PubMed]
  57. Wargan, K.; Labow, G.; Frith, S.; Pawson, S.; Livesey, N.; Partyka, G. Evaluation of the Ozone Fields in NASA’s MERRA-2 Reanalysis. J. Clim. 2017, 30, 2961–2988. [Google Scholar] [CrossRef] [PubMed]
  58. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.; Peubey, C.; Radu, R.; Schepers, D.; et al. The ERA5 Global Reanalysis. Q. J. R. Meteorol. Soc. 2020, 146, 1999–2049. [Google Scholar] [CrossRef]
  59. Rayner, N.A.; Parker, D.E.; Horton, E.B.; Folland, C.K.; Alexander, L.V.; Rowell, D.P.; Kent, E.C.; Kaplan, A. Global Analyses of Sea Surface Temperature, Sea Ice, and Night Marine Air Temperature since the Late Nineteenth Century. J. Geophys. Res. 2003, 108, 4407. [Google Scholar] [CrossRef]
  60. Andrews, D.G.; Holton, J.R.; Leovy, C.B. Middle Atmosphere Dynamics; Academic Press: Cambridge, MA, USA, 1987; 489p. [Google Scholar]
  61. Wong, T.-T. Performance evaluation of classification algorithms by k-fold and leave-one-out cross validation. Pattern Recognit. 2015, 48, 2839–2846. [Google Scholar] [CrossRef]
  62. Efron, B.; Tibshirani, R. Bootstrap Methods for Standard Errors, Confidence Intervals, and Other Measures of Statistical Accuracy. Stat. Sci. 1986, 1, 54–57. [Google Scholar] [CrossRef]
  63. Weber, M.; Dhomse, S.; Wittrock, F.; Richter, A.; Sinnhuber, B.; Burrows, J.P. Dynamical Control of NH and SH Winter/Spring Total Ozone from GOME Observations in 1995–2002. Geophys. Res. Lett. 2003, 30, 389–401. [Google Scholar] [CrossRef]
Figure 1. (a) Climatology of annual-mean zonal-mean ozone mixing ratios (ppmv), averaged from 1984 to 2022, derived from the SWOOSH combined dataset. (b) as in (a), but in Dobson Units (DUs). (c) Standard deviation of detrended zonal-mean ozone anomalies (ppmv), with piecewise linear detrending applied separately to 1984–1997 and 1998–2022. (d) as in (c) but in DUs.
Figure 1. (a) Climatology of annual-mean zonal-mean ozone mixing ratios (ppmv), averaged from 1984 to 2022, derived from the SWOOSH combined dataset. (b) as in (a), but in Dobson Units (DUs). (c) Standard deviation of detrended zonal-mean ozone anomalies (ppmv), with piecewise linear detrending applied separately to 1984–1997 and 1998–2022. (d) as in (c) but in DUs.
Remotesensing 17 02777 g001
Figure 2. (a) Time series of Arctic Stratospheric Ozone (ASO) index anomalies (averaged over 60–90°N and 1–150 hPa; units: DU) from ERA5 (green line), MERRA2 (orange line), and SWOOSH (black line) datasets, showing separate linear trends for pre- and post-1997 periods. (b) Time series of detrended ASO index anomalies with piecewise linear detrending applied to 1980–1997 and 1998–2022. (c) Monthly standard deviation of the detrended ASO index anomalies.
Figure 2. (a) Time series of Arctic Stratospheric Ozone (ASO) index anomalies (averaged over 60–90°N and 1–150 hPa; units: DU) from ERA5 (green line), MERRA2 (orange line), and SWOOSH (black line) datasets, showing separate linear trends for pre- and post-1997 periods. (b) Time series of detrended ASO index anomalies with piecewise linear detrending applied to 1980–1997 and 1998–2022. (c) Monthly standard deviation of the detrended ASO index anomalies.
Remotesensing 17 02777 g002
Figure 3. (a) Time series of February–March ASO index anomalies (solid line) from MERRA2 with segmented linear trends (dashed line; pre-1997 vs. post-1997; DU). (b) Time series of detrended February–March ASO index anomalies after piecewise linear detrending (1980–1997 and 1998–2022 periods; DU). (c,d) Correlation coefficient between detrended February–March ASO index anomalies and January–February SST anomalies. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. (e) The 21-year running correlation between January–February SST and February–March ASO. The ozone and SST data are from MERRA2 and Hadley Centre, respectively.
Figure 3. (a) Time series of February–March ASO index anomalies (solid line) from MERRA2 with segmented linear trends (dashed line; pre-1997 vs. post-1997; DU). (b) Time series of detrended February–March ASO index anomalies after piecewise linear detrending (1980–1997 and 1998–2022 periods; DU). (c,d) Correlation coefficient between detrended February–March ASO index anomalies and January–February SST anomalies. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. (e) The 21-year running correlation between January–February SST and February–March ASO. The ozone and SST data are from MERRA2 and Hadley Centre, respectively.
Remotesensing 17 02777 g003
Figure 4. (a) Standardized time series of detrended February–March ASO index anomalies (blue line) and January–February SST anomalies (red line) averaged over the Northeast Pacific (20–40°N, 105–127°W, SST_NEP) for 1980–2000. The number in the top-right corner denotes the correlation coefficient, with the asterisk (*) signifying statistical significance at the 95% confidence level. (b) As in (a), but for the period of 2001–2022. (c) Scatter plot of standardized detrended ASO versus SST_NEP anomalies with regression line (blue line; 1980–2000). (d) As in (c) but for 2001–2022. The ozone and SST data are from MERRA2 and Hadley Centre, respectively.
Figure 4. (a) Standardized time series of detrended February–March ASO index anomalies (blue line) and January–February SST anomalies (red line) averaged over the Northeast Pacific (20–40°N, 105–127°W, SST_NEP) for 1980–2000. The number in the top-right corner denotes the correlation coefficient, with the asterisk (*) signifying statistical significance at the 95% confidence level. (b) As in (a), but for the period of 2001–2022. (c) Scatter plot of standardized detrended ASO versus SST_NEP anomalies with regression line (blue line; 1980–2000). (d) As in (c) but for 2001–2022. The ozone and SST data are from MERRA2 and Hadley Centre, respectively.
Remotesensing 17 02777 g004
Figure 5. Correlation coefficient between the detrended January–February SST_NEP and zonal wind anomalies at (a) 500 hPa and (c) 200 hPa for 1980–2000. (b,d) Same as (a,c), but for 2001–2022. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. The SST and wind data are from Hadley Centre and ERA5, respectively.
Figure 5. Correlation coefficient between the detrended January–February SST_NEP and zonal wind anomalies at (a) 500 hPa and (c) 200 hPa for 1980–2000. (b,d) Same as (a,c), but for 2001–2022. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. The SST and wind data are from Hadley Centre and ERA5, respectively.
Remotesensing 17 02777 g005
Figure 6. (a) Composite differences in detrended January–February EP flux anomalies between positive and negative SST_NEP events during 1980–2000. Arrows denote the horizontal and vertical components of the Eliassen–Palm flux (units: kg/s2), while shading represents the corresponding EP flux divergence (units: m/s/day). The vertical flux was multiplied by 70. (b) Same as (a), but for 2001–2022.
Figure 6. (a) Composite differences in detrended January–February EP flux anomalies between positive and negative SST_NEP events during 1980–2000. Arrows denote the horizontal and vertical components of the Eliassen–Palm flux (units: kg/s2), while shading represents the corresponding EP flux divergence (units: m/s/day). The vertical flux was multiplied by 70. (b) Same as (a), but for 2001–2022.
Remotesensing 17 02777 g006
Figure 7. Same as Figure 6, but for (a,b) stationary and (c,d) transient waves.
Figure 7. Same as Figure 6, but for (a,b) stationary and (c,d) transient waves.
Remotesensing 17 02777 g007
Figure 8. Correlation coefficient between the detrended January–February SST_NEP and February–March anomalies of (a) temperature, (c) zonal wind, and (e) geopotential height for 1980–2000. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. (b,d,f) Same as (a,c,e), but for 2001–2022.
Figure 8. Correlation coefficient between the detrended January–February SST_NEP and February–March anomalies of (a) temperature, (c) zonal wind, and (e) geopotential height for 1980–2000. Diagonal indicates statistical significance at the 95% confidence level based on bootstrap test. (b,d,f) Same as (a,c,e), but for 2001–2022.
Remotesensing 17 02777 g008
Figure 9. (a) Observed (blue line, MERRA-2) and fitted (green line) detrended February–March ASO index anomalies from a linear regression model trained on 1980–2000. (c) Same as (a), but with the regression model trained on 1980–1994 data, showing out-of-sample predictions (orange line) for 1995–2000. The number in the top center denotes the correlation coefficient, with the asterisk (*) signifying statistical significance at the 95% confidence level. (b) Same as (a), but for 2001–2022. (d) Similar to (c) but with the regression model trained on 2001–2016 data and used to predict 2017–2022 anomalies. (e,f) Same as (a,b), but display the predicted ASO index values obtained from the leave-one-out cross-validation method.
Figure 9. (a) Observed (blue line, MERRA-2) and fitted (green line) detrended February–March ASO index anomalies from a linear regression model trained on 1980–2000. (c) Same as (a), but with the regression model trained on 1980–1994 data, showing out-of-sample predictions (orange line) for 1995–2000. The number in the top center denotes the correlation coefficient, with the asterisk (*) signifying statistical significance at the 95% confidence level. (b) Same as (a), but for 2001–2022. (d) Similar to (c) but with the regression model trained on 2001–2016 data and used to predict 2017–2022 anomalies. (e,f) Same as (a,b), but display the predicted ASO index values obtained from the leave-one-out cross-validation method.
Remotesensing 17 02777 g009
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

Chen, T.; Liao, Q. Decadal Breakdown of Northeast Pacific SST–Arctic Stratospheric Ozone Coupling. Remote Sens. 2025, 17, 2777. https://doi.org/10.3390/rs17162777

AMA Style

Chen T, Liao Q. Decadal Breakdown of Northeast Pacific SST–Arctic Stratospheric Ozone Coupling. Remote Sensing. 2025; 17(16):2777. https://doi.org/10.3390/rs17162777

Chicago/Turabian Style

Chen, Tailong, and Qixiang Liao. 2025. "Decadal Breakdown of Northeast Pacific SST–Arctic Stratospheric Ozone Coupling" Remote Sensing 17, no. 16: 2777. https://doi.org/10.3390/rs17162777

APA Style

Chen, T., & Liao, Q. (2025). Decadal Breakdown of Northeast Pacific SST–Arctic Stratospheric Ozone Coupling. Remote Sensing, 17(16), 2777. https://doi.org/10.3390/rs17162777

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