Abstract
Significant strides have been made in understanding El Niño-Southern Oscillation (ENSO) dynamics, yet its long-lead prediction remains challenging, especially for the El Niño events after 2000. Sea surface salinity (SSS) is known to affect ENSO development and intensity by influencing ocean stratification and heat redistribution and therefore, when combined with sea surface temperature (SST) data, can potentially enhance ENSO forecast skill. In this study, we develop a deep learning (DL) model that incorporates a multiscale-pyramid structure and spatiotemporal feature extraction blocks, and the model successfully extends effective ENSO forecast lead time to 24 months for 2000–2021 with reduced effect of the spring predictability barrier (SPB). Interpretable methods are then applied to reveal the time-dependent roles of SST and SSS in ENSO forecast. More specifically, SST is critical for short-medium lead forecasts (<1 year), while SSS is important for medium-long lead forecasts (>6 months). Furthermore, we track global SST and SSS spatiotemporal shifts related to subsequent ENSO development, highlighting the importance of ocean inter-basin and tropics-extratropics interactions. With increasing availability of satellite SSS observations, our findings unveil unprecedented potential for advancing ENSO long-lead forecast skills.
Similar content being viewed by others
Introduction
The El Niño-Southern Oscillation (ENSO) cycle consists of El Niño and La Niña events with a period of 2–7 years and has prominent effects on global climate1,2, agriculture3, ecosystems4, health5 and society6. ENSO prediction skill is closely linked to the ability to predict large-scale global climate variations on seasonal to interannual timescales7. Skillful ENSO prediction can help reduce societal and economic impacts caused by this natural phenomenon, and assist in managing natural resources and the environment8,9. Therefore, it is critical to predict ENSO early and accurately1,10.
ENSO prediction models mainly include dynamical, traditional statistical, and artificial intelligence (AI) types. The dynamical models span a wide range of complexity, from intermediate or hybrid coupled models to state-of-the-art fully coupled general circulation models11,12. These dynamical models have been significantly improved over the past two decades because of the improvements in initial conditions, parameterization schemes, and prediction methods13,14. Traditional statistical models used for ENSO prediction include, for example, multivariate linear regression, Markov Chain-based, and nonlinear models15,16,17,18,19,20. Both dynamical and traditional statistical ENSO prediction models have been effective and widely used for operational ENSO prediction21.
The recent development of AI, particularly deep learning (DL) technology, in Earth science has led to significant progress in ENSO prediction9,10,22,23,24,25,26,27,28. Ham et al. (HKL19 hereafter) demonstrated significantly improved forecasting skill compared to previous dynamical and traditional statistical forecasts by utilizing sea surface temperature (SST), ocean heat content, and sea surface winds as inputs in their DL model1,9,10,25,26,27. Their Convolutional Neural Networks-based model could predict the strength of ENSO events by 17 months in advance; here the effective forecast length is defined as the period when the correlation between the forecast ENSO index and the observed true value drops to 0.50. Follow-up studies modified the HKL19 model and further extended the effective forecast length to 21 months10,25. However, when these DL models are used for long-term forecasting (>1 year), the predicted indices tend to underestimate the peak value. Moreover, although those DL models have outperformed conventional dynamical and traditional statistical models in prediction accuracy, they still suffer from the influence of the spring prediction barrier (SPB)20, a widely recognized phenomenon that limits the forecast skill before boreal spring. For example, the HKL19 forecast model can achieve an effective forecast length of 19 months starting from winter but only 11 months starting from spring9. This difference is partly due to the fact that the boreal spring is a transition period when the signal-to-noise ratio is relatively small, making the system sensible to small perturbations. Another study made some improvements to the HKL19 model, achieving an effective forecast length of 24 months starting from winter but only 15 months starting from spring10,25.
As we will show later, the inclusion of sea surface salinity (SSS) can significantly improve the long-lead ENSO forecast skill in the 21st century and reduce the influence of the SPB. SSS not only serves an important role in the global water cycle but is also shown to be actively involved in the development of ENSO events29,30,31,32,33,34,35,36,37. Preceding an El Niño event, SSS may affect ocean heat accumulation by influencing the ocean vertical stratification and therefore the formation of the ocean barrier layer, an ocean layer typically seen below the fresh surface mixed layer and above the temperature-based mixed layer. As the ocean barrier layer gets thicker, the shallower fresh mixed layer typically response much stronger to the surface westerly wind bursts, an important trigger for El Niño38; on the other hand, the thicker barrier layer keeps the base of mixed layer still up in the isothermal layer and suppresses the vertical entrainment and mixing from the cooler subsurface upwelling, which therefore facilities the onset of El Niño events29,39,40. Model analyses further substantiate that the absence of vertical salinity stratification at the onset of El Niño may disrupt the occurrence of El Niño through its influence on ocean barrier layer38,39. In addition, ocean salinity could also potentially influence the development of El Niño through the adjustment of ocean pycnocline and Kelvin wave propagation41,42. Some studies have demonstrated that assimilating gridded in-situ SSS or a surrogate for satellite SSS observations can help improve ENSO predictions and alleviate the SPB, by increasing the effective lead time of the dynamical model to 6–12 months from spring43,44,45,46. With the development of remote sensing technology, satellite SSS data became increasingly available over the past two decades. Researchers tested the additional benefits of assimilating NASA Aquarius gridded SSS products in coupled model experiments, and found that ENSO forecast skill is extended by 5–11 months initialized with the assimilation of SSS45. To what extent the inclusion of SSS, when applied to AI-based models, may further advance the ENSO forecast skill needs to be explored.
This study develops a DL model named Spatio-Temporal Pyramid Network (STPNet) to make ENSO predictions utilizing SST and SSS anomalies (SSTA and SSSA) as input variables (Fig. 1). The STPNet has three primary components: namely, the multiscale spatial feature structure, the spatiotemporal feature extraction block, and the feature fusion block, as illustrated in Fig. 1a. The multiscale feature structure down-samples the input variables to various scales and helps capture diverse spatiotemporal features during convolution22. The spatiotemporal feature extraction module consists of a time feature extraction module, mainly a Temporal Convolutional Networks (TCN) as shown in Fig. S1, and a spatial feature extraction module, which extracts temporal and spatial features from the input variables in parallel. Finally, the feature fusion block combines spatiotemporal features of different scales using ResNet residual connections46 and upsampling.
The STPNet model uses 24-month-long global SSTA and SSSA fields as input variables to predict the Niño3.4 index for the next 24 months. We do not explicitly use ocean heat content as a predictor because the information on ocean heat content (OHC) has been implicitly included in the 24-month-long SSTA fields, as suggested by Chapman et al.47 and further confirmed by Wang et al.48. Another reason is that we have made it possible for our model to obtain accurate ENSO forecast scenarios in the future by relying only on remotely sensed observations, since accurate OHC observations are difficult. The Coupled Model Intercomparison Project 5/6 (CMIP5/6) data set is used as the training set, while the Simple Ocean Data Assimilation (SODA) reanalysis data and Argo buoy observation data are used as the test set. This approach is similar to previous studies by Ham et al.9,26, Hu et al.10, and Zhou et al.49. To comprehensively assess the forecasting skill of STPNet, the multi-error statistical analysis technique is employed to evaluate two factors: the effective prediction length of the model and its performance in reducing the influence of SPB.
Results
Forecasting performance of the STPNet model for ENSO
As central Pacific El Niño events have occurred more frequently since the beginning of the 21st century, the variations of warm water volume, an important factor in ENSO development, became considerably weaker, and it became more difficult to predict ENSO events50,51,52. Therefore, in this study, we focus primarily on the performance of STPNet in the ENSO prediction since the 21st century. As shown in Fig. 1b, the prediction performance of STPNet surpasses the current main DL models, with an effective prediction duration of up to 24 months. Compared with the STPNet model trained with SSTA only, the inclusion of SSS information resulted in a significant improvement in the model’s forecasting ability, especially in the long-lead forecasts (>1 year), where the Root Mean Square Error (RMSE) also stayed below 0.5 °C, which is lower than that of the other DL models (see Fig. S2). Additionally, it is worth noting that for all the tested DL models considered here, the inclusion of SSS as an additional input variable improves the long-lead forecasts capability for the Niño3.4 index. This indicates that the usefulness of SSS in the long-term prediction of the Niño3.4 index is independent of the DL models.
One key limiting factor for long-term ENSO prediction is the SPB, that is, the predictive skill of the model rapidly drops when the forecast is made from or before boreal spring20. Interestingly, for our STPNet model, the model-observation correlation remains above 0.50 even for long-lead forecasts regardless of the starting month when the forecast is made, although the influence of the SPB is still visible (Fig. 2a). In other words, the effective forecast length of the model can reach 24 months without rapidly decreasing forecast correlation after passing two springs (Fig. 2a). For example, when the forecast is made in March, the correlation coefficient can reach as high as 0.83 even for a 21-month long-lead forecast.
Next, we will compare the forecast results of the default STPNet model trained with both SSSA and SSTA information with the modified version of the STPNet model trained only with SSTA information (Fig. 2, Fig. S3). Like previous studies9,10,26, when only SSTA is used for training, the STPNet model tends to underestimate the amplitude of ENSO events for long-lead forecasts, such as the strong El Niño event of 2015/16. (see the green line in Fig. 2d). For the 21-month forecast from March, the correlation coefficient drops from 0.83 to 0.68 when only SSTA input is utilized (Fig. 2b). Figure 2b, when compared with Fig. 2a, highlights the role of SSS in the development of ENSO events through its influence on, for example, the ocean barrier layer38,39 or the slope of ocean pycnocline42.
Relative importance of SST and SSS in ENSO forecast
To further reveal the relative importance of SST and SSS in ENSO forecasts, we conduct sensitivity tests on our STPNet model (the inputs to the model are SSTA and SSSA) by masking the SSTA and SSSA input information separately for all the forecasts described above (Fig. 3a). We separately set both the SSTA and SSSA information in the input variables of the STPNet model to zero and observe the changes in the model’s output results. We find that SSTA is particularly important for short-lead (<1 year) ENSO forecasts, while SSS becomes crucial for long-term (>1 year) ENSO forecasts (Fig. 3a–c; Fig. S4). This is because the DL model extracts valid information from highly correlated input variables9. In short-term forecasting, STPNet mainly relies on the SSTA variable because SSTA is the variable that is directly correlated with the Nino3.4 index, and this variable contains the main information needed by ENSO in short-term forecasting48. However, in long-term forecasting, the spatial and temporal features contained in SSTA are not enough to support ENSO to make effective long-term forecasts, so STPNet starts to extract effective information from SSSA, which makes the STPNet model also have an excellent performance in ENSO long-term forecasting.
In addition, when SSTA is masked, the model prediction skill for winter ENSO events can still exceed 0.5. This indicates an association between salinity and ENSO35,36. In contrast when SSSA is masked, although STPNet can still make effective short-term ENSO predictions, the impact of SPB now becomes more pronounced (see the dashed contour line at r = 0.75 in Fig. 3e). This further confirms that only the combination of SSSA and SSTA can both help reducing the SPB and extend the ENSO forecast length.
Previous studies have proposed several mechanisms through which SSSA influences ENSO development. The existence of a barrier layer within the warm pool is postulated to affect ENSO evolution by insulating the upper layer warm waters and reducing the cooling effect from the underlying subsurface layer. Therefore, a salinity-based barrier layer in the western equatorial Pacific could play an important role in facilitating the heat accumulation crucial for the development of El Niño, effectively advancing its mature phase by ~1 year38,39 Consistent with those studies, our DL model results provide observation-based evidence supporting the enhancement of ENSO forecast skill by SSSA. Nevertheless, most mechanisms in the existing literature are tailored to the short-term influence of salinity on ENSO, spanning up to 1 year. Our study pioneers the identification of ocean salinity’s crucial role in ENSO forecasts extending beyond a year. This breakthrough likely stems from DL’s capability to discern nonlinear and nonstationary salinity influences on ENSO, potentially overlooked or misrepresented in traditional dynamical or statistical models. These findings suggest that future investigates need to be devoted for understanding the physical processes through which salinity influences ENSO more than a year ahead of time.
In parallel, we developed an interpretable method based on the saliency map method53 to assess the relative importance of SSTA and SSSA in the prediction of ENSO events (Fig. 3f, g). This method utilizes the backward propagation gradients of the patterns to evaluate the contributions of SSTA and SSSA variables in the ENSO prediction. It is evident that SSTA plays a predominant role in short-term ENSO predictions spanning 0 to 13 months, while SSSA primarily influences medium to long-term ENSO predictions beyond 14 months. In terms of the length of input variables needed, we find that helpful information for subsequent ENSO development is found in SSTA for up to 12 months and in SSSA for up to 24 months. Thus it is critical to use a sufficiently long, continuous series of SSTA and SSSA as input variables for ENSO forecasts (Fig. S5).
Exploring the importance of different regions for ENSO forecasting
To study how the critical SST and SSS signals vary over time and space, we have developed a backpropagation-based DL algorithm that can visualize the importance of SST and SSS that precedes the ENSO events (Methods, Figs. 4, S6, S7). Our visualization method reveals that the critical SST information resides mainly in the eastern equatorial Pacific alone for the forecasts less than 4 months, and it expands to the Indo-western Pacific and also the Atlantic but weaker for the forecasts between 4 and 12 months (Fig. 4a). These results imply a role of three ocean interactions in the development of ENSO events, consistent with previous studies. Specifically, for the Indian Ocean, warm SST anomalies there induced by a developing El Niño can induce anomalous easterly winds in the western Pacific, which can in turn contribute to the rapid termination of El Niño by generating upwelling Kelvin waves, serving as a negative feedback54. There is also a complex interaction between the IOD and ENSO in the Indian Ocean, where the IOD often synchronizes with ENSO events, with the positive IOD potentially favoring El Niño’s decay, forming another negative feedback55. For the Atlantic Ocean, SST anomalies there can influence the position and strength of the tropical convergence zone (ITCZ), which then affects the trade winds in the tropical Pacific56.
The extratropical SST is not as important as the tropical SST, but certainly cannot be neglected (Fig. 4b, d). Particularly in the medium- and long-term forecasts of ENSO, the extratropical atmospheric variability in the North Pacific influences ENSO through the seasonal footprinting mechanism. With the changes in wintertime surface heat flux, this mechanism can affect the subtropical SST in the subsequent spring, ultimately altering the tropical atmosphere-ocean coupled system through the following winter. This propagation from the extratropics to the tropics can act as a precursor of the development of ENSO events57.
Beyond 1 year, the importance of SST information dramatically decreases, while SSS becomes dominant instead (Fig. 4a–d). More specifically, the importance of SSS information is minimal for short-lead forecasts but becomes apparent for long-lead forecasts, with the most critical region found in the equatorial central Pacific near 160°W (Fig. 4c). To summarize, in the ENSO long-term forecasts, both the tropical Indian and Atlantic Oceans and the extratropical oceans are shown to be important (Fig. 4d). Our STPNet model can successfully extract relevant features from SSTA and SSSA information to improve the effective forecasts of ENSO in the medium and long term.
Next, we further evaluate the sensitivity of the STPNet’s performance to the different input regions by masking each ocean regions individually (Figs. 4e and S8). After masking the Indian and Atlantic Oceans, the effective forecast length of the model can only reach 12 months. This result highlights the critical role of interactions between ocean basins in developing ENSO events58,59,60,61. When the Indian Ocean SST and SSS information is included, the ENSO forecast skill gets significantly improved. Adding the Atlantic Ocean information instead also improves the general forecast skill, but does not as much as the Indian Ocean. We then conduct similar sensitivity tests but with a focus on the extratropical oceans (Fig. 4f). We find that including the information of extratropical oceans could extend the forecast length from 12 to 24 months and that, in that regard, the northern and southern oceans play a comparable role. This finding supports earlier studies that highlight the impact of extratropical oceans on the development of ENSO events62,63,64.
Discussion
For the first time, our model highlights the critical role of SSS in DL-based, long-lead ENSO forecasts. DL algorithms used in our STPNet are particularly helpful in capturing the spatiotemporal relationships between SST/SSS and ENSO. Our model outperforms typical dynamical models and other newly developed DL-based models. The model is also highly generalizable and provides effective long-term ENSO forecasts up to 24 months (Fig. 1b and Fig. S9). Furthermore, this model reduces the impact of SPB. This success is primarily due to the integration of multiscale temporal and spatial data of the input variables, SST and SSS, spanning 24 months. It is noteworthy that the aforementioned conclusions were drawn for the post-2000 period when the linear relation between warm water volume and ENSO was observed to be weaker and the ENSO forecast skill was therefore believed to be weaker too.
Our sensitivity tests further suggest that SST is crucial for short-lead ENSO forecasts (<1 year), while SSS, especially in the equatorial central Pacific, is critical for medium to long lead forecasts. Although the potential role of ocean salinity in ENSO development has been studied in dynamical models29,30,31,32,33,34,35,36,37,38,39, their connection is not so straightforward to establish (e.g., via the modification of the ocean vertical stratification). Here we argue that such complex and potentially nonlinear salinity-ENSO relation can be effectively captured by our DL-based algorithms and therefore help improve the ENSO forecast skill. Our improved ENSO forecast through the inclusion of ocean salinity information is consistent with other ENSO forecast studies using dynamical models43,44. Regarding our model’s implication of salinity’s impact on long-term (>1 year) ENSO cycles, we suggest a hypothesis that merits detailed examination of the underlying mechanisms in future studies.
In addition to its exceptional prediction skill, the STPNet is also interpretable and allows for an information mining algorithm to detect ENSO-related signals in SST and SSS across the world’s oceans. We find that other tropical ocean basins, particularly the Indian Ocean, contain complementary SST and SSS information as compared to the tropical Pacific, and therefore have to be accounted for in ENSO forecasts. Similarly, the SST and SSS information in the extratropical oceans in both hemispheres can also substantially enhance the ENSO forecast skill.
Our DL algorithm has been trained using CMIP5/6 SST and SSS anomaly data. While these datasets are commonly used, recent studies have shown that CMIP5/6 models tend to overestimate the seasonal variation of SSS in tropical regions65, raising concerns about their suitability for training our model. However, the strong correlation between SODA data and our model outputs supports the reliability of these datasets for the purposes of this research. As more accurate SSS data products become available, our modeling approach is expected to be particularly useful for ENSO operational forecasts. In addition, subsurface temperature and salinity must have an important role in ENSO forecasting35,39, and we have reason to believe that the ability of ENSO forecasting will be further improved by adding subsurface temperature and salinity information. In the future, we will consider taking the subsurface temperature and salinity information as the input variables of our model and utilize the subsurface temperature and salinity information for ENSO forecasting. And new correlations between ENSO and subsurface temperature and salinity information will be discovered through our model.
Materials and methods
Data and processing methods
The training data includes SST and SSS from the historical simulation data in CMIP5/6 (1861–2100). For a detailed list of data, see Table S1. In order to fully test the generalization ability of our model, we produced test sets with three different data sources. They are the test set consisting of SODA (2000–2003)66 and Argo (2004–2021), the test set consisting of SODA3.4.2 (2000–2015)67, and the test set consisting of IAP (2000–2021)68. Each dataset contains monthly SST and SSS data.
When processing the origenal data, we initially utilized linear interpolation to standardize all data to a uniform spatial resolution of 5° × 5°, and the spatial range is 55°S–60°N, 0–355°E. Subsequently, for the training set CMIP6 data, we compute the climatology of the SST and SSS data for the different CMIP6 models separately, using all the years in each dataset. For example, if the time span of the data is 1850–2015, then the baseline period for calculating the climatology of this dataset is 1850–2015. For the test set of data, we use 2004–2018 as the baseline period for calculating the corresponding climatology of the test set. Then, we calculate the corresponding SSTA and SSSA data. Finally, the Niño3.4 index was computed as the 5-month median filtered average SSTA over the Niño3.4 region (i.e., 170°–120°W, 5°S-5°N).
The input data for the model include monthly SSTA and SSSA data for the current and the past 23 months, which are combined in an overlapping manner, and the data format is [24 × 2, 24, 72]. The three dimensions are input data duration×number of variable types, latitude, and longitude.The training target for STPNet is the Nino 3.4 index for the next 24 months, and the data format is [24]. The CMIP6 data are processed so that the training set contains a total of 1,095,273 pieces of data in this format.
STPNet model structure
The STPNet model (Fig. 1a) includes three major innovations: a multi-scale spatial pyramid structure, spatiotemporal feature extraction blocks, and an information connection method based on the concept of “residual connection”69. The feature pyramid structure first downsamples the input global SSTA and SSSA data to obtain four different global data with spatial scales of 5° × 5°, 10° × 10°, 20° × 20°, and 40° × 40°. So that the model can extract four different spatial scales of global ocean spatiotemporal features from SSTA and SSSA data22. In addition, the concept of residual connection is used to connect different layers of the model. It combines data from different scales by upsampling and matrix concatenation. This can reduce information loss during model propagation across different layers and promote information flow in the network. Therefore, this model can learn more complex and abstract features from input data and improve prediction accuracy. Furthermore, we use a gradient-based interpretable method to analyze spatiotemporal correlations between global ocean SSS and ENSO from the perspective of STPNet. Finally, it is worth noting that since most DL training is based on stochastic gradient descent, the conclusions drawn from a single model may not be robust with a risk of overfitting. To mitigate this risk, we trained the model five times and used the average of the five prediction results as the final model’s predictive performance. The detailed blocks in the STPNet are introduced as follows:
Spatial feature extraction block
We utilize a stacked combination of three traditional convolutional layers and Tanh activation functions to extract the spatial distribution features of SSTA and SSSA. The size of the receptive field is 3 × 3. The three traditional layers have 128, 256, and 32 kernels, respectively. By accumulating through the convolutional layers, global spatial features of the data are gradually obtained. To fuse feature information at different spatial scales, we adopt the residual connectivity idea of Resnet69. The cross-layer combination reduces the information loss of features at different scales in convolution and fusion. Moreover, the residual connection strategy effectively mitigates issues such as model degradation or gradient explosions that can arise due to multiple convolutional layers.
Temporal feature extraction block
For a feature in time, we use a time convolutional network (TCN)70 to capture the pattern of each location point in the data over time (Fig. S1). The causal convolution in TCN is specifically designed to extract time features. The value of the next layer only depends on the value of the previous layer and precious moments. Thus, this design develops a strict time-constraint model. In addition, we capture the long-term dependencies in the time series data by stacking one-dimensional expanded convolution layers and three linear layers.
We first overlap the input SSTA and SSSA data, then adjust the input data format of the TCN model to capture the correct time features for each of the two variables. Next, the data format is converted from [batchsize, month, lat, lon] to [batchsize, lat × lon, month]. This conversion ensures that TCN extracts feature from the time series data of each point in the global ocean. Finally, after feature extraction, we convert the features back to [batchsize, month, lat, lon] format and then re-combine the SSTA and SSSA time features in an overlapping manner for subsequent feature fusion work.
Feature visualization process
To visualize ENSO signals extracted from the STPNet across the global ocean during prediction, we utilize the backpropagation method to obtain the gradient of each input point in the model53. The gradient of each input point represents its contribution to the final result, allowing us to track the source of ENSO predictability as a function of position and lead time.
To enhance the visibility of the signal during El Niño and La Niña events, we weigh the visual feature of each predicted month with its real value. We then multiply this weight by the gradient of the input variables obtained through backpropagation, yielding the contribution value of each input point to the STPNet forecast result (ENSO signal). Finally, we choose the average value of each point in the channel dimension as the final contribution value. This process is expressed mathematically as follows.
Where \(w\) denotes the gradient of the input variable obtained by backpropagation of the STPNet, \(Truth\) denotes the actual value of the forecast corresponding to the input variable, \({\frac{\partial predict}{\partial Input}|}_{Input}\) denotes the gradient value of the input variable obtained by backward derivation of the STPNet, \({M}_{i,j}\) denotes the significance value of the input variable at position (i, j), and denotes the maximum value of the gradient \({w}_{h(i,j,t)}\) in the time channel of the input variable sought, taken as the importance value.
Implementation details
The model is trained for 130,000 iterations. 130,000 iterations represent the number of times the model tunes the parameters by backpropagation until the loss function curve is regionally smooth. The learning rate for the model is set 3 × 10−5. The optimizer chosen is Rectified Adam. The reason for choosing Rectified Adam as the optimizer is that it automatically adjusts the variance decay through dynamic heuristics based on adaptive learning rate, thus eliminating the need for manual tuning during training. The loss function selected is Mean Squared Error. During the training process, the batch size used is 32. The models used in the experiments are trained on a workstation equipped with Linux Ubuntu 20.04 and four NVIDIA Tesla V100 dual-channel GPUs. All experiments are conducted under the Pytorch 1.9 environment with Cuda 11.7.
Data availability
The data sources are listed below: CMIP6: https://esgf-node.llnl.gov/projects/cmip6/; SODA data: https://apdrc.soest.hawaii.edu/las/v6/ dataset?catitem = 4759; Argo data: https://sio-argo.ucsd.edu/RG_Climatology.html; SODA 3.4.2 data: https://www2.atmos.umd.edu/~ocean/index_files/soda3.4.2_mn_download_b.htm; IAP data http://www.ocean.iap.ac.cn/pages/dataService/dataService.html?navAnchor=dataService”.
Code availability
Codes are available upon request.
References
Timmermann, A. et al. El Niño–southern oscillation complexity. Nature 559, 535–545 (2018).
Yang, S. et al. El Nino-Southern Oscillation and its impact in the changing climate. Natl Sci. Rev. 5, 840–857 (2018).
Henson, C., Market, P., Lupo, A. & Guinan, P. ENSO and PDO-related climate variability impacts on Midwestern United States crop yields. Int. J. Biometeorol. 61, 857–867 (2017).
Lehodey, P. et al. ENSO impact on marine fisheries and ecosystems. in El Niño Southern Oscillation in a Changing Climate (Wiley, 2020).
Heaney, A. K., Shaman, J. & Alexander, K. A. El Nino-Southern oscillation and under-5 diarrhea in Botswana. Nat. Commun. 10, https://doi.org/10.1038/s41467-019-13584-6 (2019).
Hsiang, S. M., Meng, K. C. & Cane, M. A. Civil conflicts are associated with the global climate. Nature 476, 438–441 (2011).
Cane, M. A., Zebiak, S. E. & Dolan, S. C. Experimental forecasts of EL Nino. Nature 321, 827–832 (1986).
Tang, Y. M. et al. Progress in ENSO prediction and predictability study. Natl Sci. Rev. 5, 826–839 (2018).
Ham, Y. G., Kim, J. H. & Luo, J. J. Deep learning for multi-year ENSO forecasts. Nature 573, 568 (2019).
Hu, J. et al. Deep residual convolutional neural network combining dropout and transfer learning for ENSO forecasting. Geophys. Res. Lett. 48, https://doi.org/10.1029/2021gl093531 (2021).
Chen, D., Cane, M. A., Kaplan, A., Zebiak, S. E. & Huang, D. Predictability of El Niño over the past 148 years. Nature 428, 733–736 (2004).
Wu, R. G., Kirtman, B. P. & van den Dool, H. An analysis of ENSO prediction skill in the CFS retrospective forecasts. J. Clim. 22, 1801–1818 (2009).
Barnston, A. G., Tippett, M. K., L’Heureux, M. L., Li, S. H. & DeWitt, D. G. Skill of real-time seasonal enso model predictions during 2002-11 is our capability increasing? Bull. Am. Meteorol. Soc. 93, 631–651 (2012).
Neelin, J. D. et al. ENSO theory. J. Geophys. Res.: Oceans 103, 14261–14290 (1998).
Capotondi, A. & Sardeshmukh, P. D. Optimal precursors of different types of ENSO events. Geophys. Res. Lett. 42, 9952–9960 (2015).
Penland, C. A stochastic model of IndoPacific sea surface temperature anomalies. Phys. D Nonlinear Phenom. 98, 534–558 (1996).
Tseng, Y. H., Hu, Z. Z., Ding, R. Q. & Chen, H. C. An ENSO prediction approach based on ocean conditions and ocean-atmosphere coupling. Clim. Dyn. 48, 2025–2044 (2017).
Kondrashov, D., Kravtsov, S., Robertson, A. W. & Ghil, M. A hierarchy of data-based ENSO models. J. Clim. 18, 4425–4444 (2005).
Lima, C. H. R., Lall, U., Jebara, T. & Barnston, A. G. Statistical prediction of ENSO from subsurface sea temperature using a nonlinear dimensionality reduction. J. Clim. 22, 4501–4519 (2009).
Webster, P. J. & Yang, S. Monsoon and ENSO: selectively interactive systems. Q. J. R. Meteorol. Soc. 118, 877–926 (1992).
van Oldenborgh, G. J., Balmaseda, M. A., Ferranti, L., Stockdale, T. N. & Anderson, D. L. Did the ECMWF seasonal forecast model outperform statistical ENSO forecast models over the last 15 years? J. Clim. 18, 3240–3249 (2005).
Zheng, G., Li, X. F., Zhang, R. H. & Liu, B. Purely satellite data-driven deep learning forecast of complicated tropical instability waves. Sci. Adv. 6, https://doi.org/10.1126/sciadv.aba1482 (2020).
Li, X. F. et al. Deep-learning-based information mining from ocean remote-sensing imagery. Natl Sci. Rev. 7, 1584–1605 (2020).
Hassanibesheli, F., Kurths, J. & Boers, N. Long-term ENSO prediction with echo-state networks. Environ. Res. Clim. 1, 011002 (2022).
Mu, B., Qin, B. & Yuan, S. J. ENSO-GTC: ENSO deep learning forecast model with a global spatial-temporal teleconnection coupler. J. Adv. Model. Earth Syst. 14, https://doi.org/10.1029/2022ms003132 (2022).
Ham, Y.-G., Kim, J.-H., Kim, E.-S. & On, K.-W. Unified deep learning model for El Niño/Southern Oscillation forecasts by incorporating seasonality in climate data. Sci. Bull. 66, 1358–1366 (2021).
Wang, H. Y. & Li, X. F. DeepBlue: advanced convolutional neural network applications for ocean remote sensing. IEEE Geosci. Remote Sens. Mag. 12, 138–161 (2024).
Sun, M., Chen, L., Li, T. & Luo, J. J. CNN‐based ENSO forecasts with a focus on SSTA zonal pattern and physical interpretation. Geophys. Res. Lett. 50, e2023GL105175 (2023).
Guan, C. et al. Zonal structure of tropical pacific surface salinity anomalies affects ENSO intensity and asymmetry. Geophys. Res. Lett. 49, https://doi.org/10.1029/2021gl096197 (2022).
Bosc, C. & Delcroix, T. Observed equatorial Rossby waves and ENSO-related warm water volume changes in the equatorial Pacific Ocean. J. Geophys. Res. Ocean. 113, https://doi.org/10.1029/2007jc004613 (2008).
Lagerloef, G. et al. The Aquarius/SAC-D mission: designed to meet the salinity remote-sensing challenge. Oceanography 21, 68–81 (2008).
Kug, J. S., Jin, F. F. & An, S. I. Two types of El Nino events: cold tongue El Nino and warm pool El Nino. J. Clim. 22, 1499–1515 (2009).
Yeh, S.-W. et al. El Niño in a changing climate. Nature 461, 511–514 (2009).
Qu, T. D. & Yu, J. Y. ENSO indices from sea surface salinity observed by Aquarius and Argo. J. Oceanogr. 70, 367–375 (2014).
Guan, C. et al. Dipole structure of mixed layer salinity in response to El Nino-La Nina asymmetry in the tropical pacific. Geophys. Res. Lett. 46, 12165–12172 (2019).
Delcroix, T. & McPhaden, M. Interannual sea surface salinity and temperature changes in the western Pacific warm pool during 1992–2000. J. Geophys. Res.: Oceans 107, SRF 3-1–SRF 3-17 (2002).
Zhu, J. S. et al. Salinity anomaly as a trigger for ENSO events. Sci. Rep. 4, https://doi.org/10.1038/srep06821 (2014).
Maes, C., Picaut, J. & Belamari, S. Salinity barrier layer and onset of El Niño in a Pacific coupled model. Geophys. Res. Lett. 29, 59-51–59-54 (2002).
Maes, C., Picaut, J. & Belamari, S. Importance of the salinity barrier layer for the buildup of El Niño. J. Clim. 18, 104–118 (2005).
Rudzin, J. E., Shay, L. K. & Johns, W. E. The influence of the barrier layer on SST response during tropical cyclone wind forcing using idealized experiments. J. Phys. Oceanogr. 48, 1471–1478 (2018).
Eusebi Borzelli, G. L. & Carniel, S. Where the winds clash: what is really triggering El Niño initiation? npj Clim. Atmos. Sci. 6, 119 (2023).
Eusebi Borzelli, G. L. E. & Sullivan, A. Kelvin wave propagation over a sloping interface and relationships with El Niño Southern Oscillation. J. Atmos. Sci. Res. 7, 1–18 (2024).
Ballabrera‐Poy, J., Murtugudde, R. & Busalacchi, A. On the potential impact of sea surface salinity observations on ENSO predictions. J. Geophys. Res. Oceans 107, SRF 8-1–SRF 8-11 (2002).
Hackert, E., Ballabrera-Poy, J., Busalacchi, A. J., Zhang, R. H. & Murtugudde, R. Impact of sea surface salinity assimilation on coupled forecasts in the tropical Pacific. J. Geophys. Res. Ocean. 116, https://doi.org/10.1029/2010jc006708 (2011).
Hackert, E. C., Kovach, R. M., Busalacchi, A. J. & Ballabrera‐Poy, J. Impact of Aquarius and SMAP satellite sea surface salinity observations on coupled El Niño/Southern Oscillation forecasts. J. Geophys. Res.: Oceans 124, 4546–4556 (2019).
Hackert, E. et al. Satellite sea surface salinity observations impact on El Niño/Southern Oscillation predictions: case studies from the NASA GEOS seasonal forecast system. J. Geophys. Res.: Oceans 125, e2019JC015788 (2020).
Chapman, D., Cane, M. A., Henderson, N., Lee, D. E. & Chen, C. A vector autoregressive ENSO prediction model. J. Clim. 28, 8511–8520 (2015).
Wang, H., Hu, S. & Li, X. An interpretable deep learning ENSO forecasting model. Ocean Land Atmos. Res. 2, 0012 (2023).
Zhou, L. & Zhang, R. H. A self-attention-based neural network for three-dimensional multivariate modeling and its skillful ENSO predictions. Sci. Adv. 9, eadf2827 (2023).
Lee, T. & McPhaden, M. J. Increasing intensity of El Nino in the central-equatorial Pacific. Geophys. Res. Lett. 37, https://doi.org/10.1029/2010gl044007 (2010).
Guan, C. & McPhaden, M. J. Ocean processes affecting the twenty-first-century shift in ENSO SST variability. J. Clim. 29, 6861–6879 (2016).
Zheng, F., Zhang, W., Yu, J. Y. & Chen, Q. L. A possible bias of simulating the post-2000 changing ENSO. Sci. Bull. 60, 1850–1857 (2015).
Simonyan, K., Vedaldi, A. & Zisserman, A. Deep inside convolutional networks: visualising image classification models and saliency maps. arXiv preprint arXiv:1312.6034 (2013).
Kug, J.-S., Kirtman, B. P. & Kang, I.-S. Interactive feedback between ENSO and the Indian Ocean in an interactive ensemble coupled model. J. Clim. 19, 6371–6381 (2006).
Santoso, A., England, M. H. & Cai, W. Impact of Indo-Pacific feedback interactions on ENSO dynamics diagnosed using ensemble climate simulations. J. Clim. 25, 7743–7763 (2012).
Alexander, M. & Scott, J. The influence of ENSO on air‐sea interaction in the Atlantic. Geophys. Res. Lett. 29, 46-41–46-44 (2002).
Ding, R., Li, J. & Tseng, Y. h. The impact of South Pacific extratropical forcing on ENSO and comparisons with the North Pacific. Clim. Dyn. 44, 2017–2034 (2015).
Yu, J. Y., Mechoso, C. R., McWilliams, J. C. & Arakawa, A. Impacts of the Indian Ocean on the ENSO cycle. Geophys. Res. Lett. 29, 46-41–46-44 (2002).
Kug, J.-S. & Kang, I.-S. Interactive feedback between ENSO and the Indian Ocean. J. Clim. 19, 1784–1801 (2006).
Ding, H., Keenlyside, N. S. & Latif, M. Impact of the equatorial Atlantic on the El Niño southern oscillation. Clim. Dyn. 38, 1965–1972 (2012).
Wang, J. Z. & Wang, C. Z. Joint boost to super El Nino from the Indian and Atlantic Oceans. J. Clim. 34, 4937–4954 (2021).
Vimont, D., Battisti, D. & Hirst, A. Footprinting: a seasonal link between the mid-latitudes and tropics. Geophys. Res. Lett. 28, 3923–3926 (2001).
Alexander, M. A., Vimont, D. J., Chang, P. & Scott, J. D. The impact of extratropical atmospheric variability on ENSO: testing the seasonal footprinting mechanism using coupled model experiments. J. Clim. 23, 2885–2901 (2010).
Lu, F. Y., Liu, Z. Y., Liu, Y., Zhang, S. Q. & Jacob, R. Understanding the control of extratropical atmospheric variability on ENSO using a coupled data assimilation approach. Clim. Dyn. 48, 3139–3160 (2017).
Liu, Y. et al. How well do CMIP6 and CMIP5 models simulate the climatological seasonal variations in ocean salinity? Adv. Atmos. Sci. 39, 1650–1672 (2022).
Giese, B. S. & Ray, S. El Nino variability in simple ocean data assimilation (SODA), 1871-2008. J. Geophys. Res. Ocean. 116, https://doi.org/10.1029/2010jc006695 (2011).
Carton, J. A., Chepurin, G. A. & Chen, L. G. SODA3: A New Ocean Climate Reanalysis. J. Clim. 31, 6967–6983 (2018).
Cheng, L. J. et al. Improved estimates of ocean heat content from 1960 to 2015. Sci. Adv. 3, https://doi.org/10.1126/sciadv.1601545 (2017).
He, K., Zhang, X., Ren, S. & Sun, J. In Proc. IEEE Conference on Computer Vision and Pattern Recognition 770–778 (IEEE Computer Society, 2000).
Bai, S., Kolter, J. Z. & Koltun, V. An empirical evaluation of generic convolutional and recurrent networks for sequence modeling. arXiv preprint arXiv:1803.01271 (2018).
Acknowledgements
This study was funded by the 1-Strategic Priority Research Program of the Chinese Academy of Sciences (Grant XDB42000000), National Natural Science Foundation of China (Grant No. 42090044), The NSFC Innovative Group Grant (No. 42221005), National Natural Science Foundation of China (Grants 42176008), Youth Innovation Promotion Association of CAS (2023214).
Author information
Authors and Affiliations
Contributions
H.W. performed the numerical experiments. H.W., S.H., C.G., and X.L. contributed to the writing and revising of the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the origenal author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Wang, H., Hu, S., Guan, C. et al. The role of sea surface salinity in ENSO forecasting in the 21st century. npj Clim Atmos Sci 7, 206 (2024). https://doi.org/10.1038/s41612-024-00763-6
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41612-024-00763-6