Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-23T19:14:39.071Z Has data issue: false hasContentIssue false

Evolution of machine learning in environmental science—A perspective

Published online by Cambridge University Press:  13 April 2022

William W. Hsieh*
Affiliation:
Department of Earth, Ocean and Atmospheric Sciences, University of British Columbia, Vancouver, British Columbia V6T 1Z4, Canada
*
Corresponding author: E-mail: whsieh@eos.ubc.ca

Abstract

The growth of machine learning (ML) in environmental science can be divided into a slow phase lasting till the mid-2010s and a fast phase thereafter. The rapid transition was brought about by the emergence of powerful new ML methods, allowing ML to successfully tackle many problems where numerical models and statistical models have been hampered. Deep convolutional neural network models greatly advanced the use of ML on 2D or 3D data. Transfer learning has allowed ML to progress in climate science, where data records are generally short for ML. ML and physics are also merging in new areas, for example: (a) using ML for general circulation model parametrization, (b) adding physics constraints in ML models, and (c) using ML in data assimilation.

Type
Position Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

Impact Statement

This perspective paper reviews the evolution and growth of machine learning (ML) models in environmental science. The opaque nature of ML models led to decades of slow growth, but exponential growth commenced around the mid-2010s. Novel ML models which have contributed to this exponential growth (e.g., deep convolutional neural networks, encoder–decoder networks, and generative-adversarial networks) are reviewed, as well as approaches to merging ML models with physics-based models.

1. Introduction

Thirty years ago, a typical environmental scientist would know some statistics but would not have heard of “machine learning” (ML) and would know artificial intelligence (AI) only through science fiction. After World War II, the great popularity of AI in science fiction led to very unrealistic expectations on how fast AI research would progress. The inevitable disappointment led to negative reviews and two major “AI winters,” periods of poor funding around 1974–1980 and 1987–1993 (Crevier, Reference Crevier1993; Nilsson, Reference Nilsson2009). Partly to focus on a more specific aspect and partly to avoid the stigma associated with AI, many researchers started to refer to their work using other names, for example, ML, where the goal of ML is to have computers learn from data without being explicitly programmed.

While there is overlap between the data methods developed in statistics and in the younger field of ML (Figure 1), ML germinated mainly in computer science, psychology, engineering, and commerce, while statistics had largely been rooted in mathematics, leading to two fairly distinct cultures (Breiman, Reference Breiman2001b). When fitting a curve to a dataset, a statistician would ensure the number of adjustable model parameters is small compared to the sample size (i.e., number of observations) to avoid overfitting (i.e., the model fitting to the noise in the data). This prudent practice in statistics is not strictly followed in ML, as the number of parameters (called “weights” in ML) can be greater, sometimes much greater, than the sample size (Krizhevsky et al., Reference Krizhevsky, Sutskever and Hinton2012), as ML has developed ways to avoid overfitting while using a large number of parameters. The relatively large number of parameters renders ML models much more difficult to interpret than statistical models; hence, ML models are often regarded somewhat dismissively as “black boxes.” In ML, the artificial neural network (NN) model called the multilayer perceptron (MLP; Rumelhart et al., Reference Rumelhart, Hinton, Williams, Rumelhart and McClelland1986; Goodfellow et al., Reference Goodfellow, Bengio and Courville2016) has become widely used since the late 1980s.

Figure 1. Venn diagram illustrating the relation between artificial intelligence, statistics, machine learning, neural networks, and deep learning, as well as kernel methods, random forests, and boosting.

How readily an environmental science (ES) adopted NN or other ML models depended on whether successful physics-based models were available. Meteorology, where dynamical models have been routinely used for weather forecasting, has been slower to embrace NN models than hydrology, where by year 2000 there were already 43 hydrological papers using NN models (Maier and Dandy, Reference Maier and Dandy2000), as physics-based models were not very skillful in forecasting streamflow from precipitation data.

Relative to linear statistical models, the nonlinear ML models also need relatively large sample size to excel. Hence, oceanography, a field with far fewer in situ observations than hydrology or meteorology, and climate science, where the longtime scales preclude large effective sample size, are fields where the application of ML models has been hampered. Furthermore, averaging daily data to produce climate data linearizes the relation between the predictors and the response variables due to the central limit theorem, thereby reducing the nonlinear modelling advantage of ML models (Yuval and Hsieh, Reference Yuval and Hsieh2002). Another disadvantage of many ML models (e.g., NN) relative to linear statistical models is that they can extrapolate much worse when given new predictor data lying outside the original training domain (Hsieh, Reference Hsieh2020), as nonlinear extrapolation is an ill-posed problem. This is not ML-specific, since any nonlinear statistical model faces the same ill-posed problem when used for extrapolation.

Overall, around year 2010, ML models were fairly well accepted in hydrology and remote sensing, but remained fringe in meteorology and even less developed in oceanography and climate science. Nevertheless, a number of books were written on the application of ML methods to ES in this earlier development phase (Abrahart et al., Reference Abrahart, Kneale and See2004; Blackwell and Chen, Reference Blackwell and Chen2009; Haupt et al., Reference Haupt, Pasini and Marzban2009; Hsieh, Reference Hsieh2009; Krasnopolsky, Reference Krasnopolsky2013). After this relatively flat phase, rapid growth of ML in ES commenced in the mid-2010s.

Section 2 looks at the emergence of new and more powerful ML methods in the last decade, whereas Section 3 reviews their applications to ES. The merging of ML and physical/dynamical models is examined in Section 4.

2. Evolution of ML Methods

The MLP NN model maps the input variables through layers of hidden neurons/nodes (i.e., intermediate variables) to the output variables. Given some training data for the input and output variables, the model weights can be solved by minimizing an objective function using a back-propagation algorithm. The traditional MLP NN is mostly limited to one or two hidden layers, because the gradients (error signals) in the back-propagation algorithm become vanishingly small after propagating through many layers. Without a solution for the vanishing gradient problem, NN research stalled while newer methods—kernel methods (e.g., support vector machines (Cortes and Vapnik, Reference Cortes and Vapnik1995)) emerging from the mid-1990s and random forests from 2001 (Breiman, Reference Breiman2001a)—seriously challenged NN’s dominant position in ML.

The traditional MLP NN has each neuron in one layer connected to all the neurons in the preceding layer. When working with image data, the MLP uses a huge number of weights—for example, mapping an input 100 × 100 image to just one neuron in the first hidden layer requires 10,000 weights! Since, in nature, biological neurons are connected only to neighboring neurons, to have every neuron in one layer of an NN model connected to all the neurons in the preceding layer is unnatural and very wasteful of computing resources. With inspiration from the animal visual cortex, the convolutional layer has been developed, where a neuron is only connected to a small patch of neurons in the preceding layer, thereby resulting in a drastic reduction of weights compared to the traditional fully connected layers and giving rise to convolutional neural networks (CNNs; LeCun et al., Reference LeCun, Boser, Denker, Henderson, Howard, Hubbard and Jackel1989).

Eventually, the vanishing gradient problem was overcome, and deep NN or deep learning (DL) models, that is, NN having $ \gtrsim 5 $ layers of mapping functions with adjustable weights (LeCun et al., Reference LeCun, Bengio and Hinton2015), emerged. In 2012, a deep NN model won the ImageNet Large Scale Visual Recognition Challenge (Krizhevsky et al., Reference Krizhevsky, Sutskever and Hinton2012). The huge reduction in the number of weights in convolutional layers made deep NN feasible.

With all the impressive breakthroughs in DL since 2012 (Goodfellow et al., Reference Goodfellow, Bengio and Courville2016), there is a popular misconception that DL models are superior to all other ML models. Actually, the best ML model is very problem-dependent. There are two main types of datasets, structured and unstructured. Structured datasets have a tabular format, for example, like an Excel spreadsheet, with the variables listed in columns. In contrast, unstructured datasets include images, videos, audio, text, and so forth. For unstructured datasets, DL has indeed been dominant. For structured datasets, however, ML models with shallow depth structure, for example, gradient boosting models such as XGBoost (Chen and Guestrin, Reference Chen and Guestrin2016), have often beaten deep NN in competitions, for example, those organized by Kaggle (www.kaggle.com).

How can “shallow” gradient boosting beat deep NN in structured data? Typically, in a structured dataset, the predictors are quite inhomogeneous (e.g., pressure, temperature, and humidity), whereas in an unstructured dataset, the predictors are more homogeneous (e.g., temperature at various pixels in a satellite image or at various grid points in a numerical model). Boosting is based on decision trees, where the effects of the predictors X are treated independently of each other, as the path through a decision tree is controlled by questions like: is “ $ {x}_1>a? $ ,” “ $ {x}_2>b? $ ,” and so forth (Breiman et al., Reference Breiman, Friedman, Olshen and Stone1984). In contrast, in NN, the predictors are combined by a linear combination ( $ {\sum}_i{w}_i{x}_i $ ) before being passed through an activation/transfer function onto the next layer. With inhomogeneous predictors, for example, temperature and pressure, treating the two separately as in decision trees intuitively makes more sense than trying to add the two together by a linear combination.

Another type of NN architecture is the encoder–decoder model, where the encoder part first maps from a high-dimensional input space to a low-dimensional space, then the decoder part maps back to a high-dimensional output space (Figure 2). If the target output data are the same as the input data, the model becomes an autoencoder, which has been used for nonlinear principal component analysis (PCA), as the low-dimensional space can be interpreted as nonlinear principal components (Kramer, Reference Kramer1991; Monahan, Reference Monahan2000; Hsieh, Reference Hsieh2001). The popular U-net (Ronneberger et al., Reference Ronneberger, Fischer and Brox2015) is a deep CNN model with an encoder–decoder architecture.

Figure 2. The encoder–decoder is an NN model with the first part (the encoder) mapping from the input x to u, the “code” or bottleneck, and the second part (the decoder) mapping from u to the output y. Dimensional compression is achieved by forcing the signal through the bottleneck. The encoder and the decoder are each illustrated with only one hidden layer for simplicity.

In many games, having two individuals playing against each other enhances the skill level of both, for example, a soccer goal scorer practicing against a goal keeper. The generative adversarial network (GAN) has two submodels, the generator and the discriminator, playing as adversaries, with the goal of producing realistic fake data (Goodfellow et al., Reference Goodfellow, Pouget-Abadie, Mirza, Xu, Warde-Farley, Ozair, Courville and Bengio2014). Given a random input vector, the generator outputs a set of fake data. The discriminator receives either real data or fake data as input and classifies them as either “real” or “fake” (Figure 3). If a fake is correctly classified, the generator’s model weights are updated, whereas if the fake is mistaken to be real, the discriminator’s model weights are updated. The skill levels of both players improve until at the end the discriminator can only identify fake data from the generator about 50% of the time. After training is done, the discriminator is discarded while the generator is retained to produce new fake data. The variational autoencoder model provides an alternative to GAN (Kingma and Welling, Reference Kingma and Welling2014).

Figure 3. Generative adversarial network with the generator creating a fake image (e.g., a fake Picasso painting) from random noise input, and the discriminator classifying images as either real or fake. Whether the discriminator classifies a fake image rightly or wrongly leads, respectively, to further training for the generator or for the discriminator.

Conditional generative adversarial network (CGAN), introduced by Mirza and Osindero (Reference Mirza and Osindero2014), supplies additional input x to both the generator and the discriminator. If X is an image, CGAN can be used for image-to-image translation tasks (Isola et al., Reference Isola, Zhu, Zhou and Efros2017), for example, translating a line drawing to a photo image, a map to a satellite map, or vice versa (Figure 4).

Figure 4. Conditional generative adversarial network where the generator G receives an image x and a random noise vector z as input. The discriminator D receives x plus either a fake image from G (left) or a real image y (right) as input. Here, a line drawing is converted to a photo image; similarly, a photo image can be converted to a line drawing.

Adapted from Figure 2 of Isola et al. (Reference Isola, Zhu, Zhou and Efros2017).

3. Applications in Environmental Science

Applications of ML to ES have come in roughly two groups. In the first group, ML methods are used largely as nonlinear generalizations of traditional statistical methods. For instance, MLP NN models are used for nonlinear regression, classification, PCA, and so forth (Hsieh, Reference Hsieh2009). In the second group, the ML methods do not have counterparts in statistics, for example, GAN or CGAN.

CNN models have appeared in ES in the last few years, for example, to estimate the posterior probability of three types of extreme events (tropical cyclones, atmospheric rivers, and weather fronts) from 2D images of atmospheric variables (Liu et al., Reference Liu, Racah, Prabhat, Correa, Khosrowshahi, Lavers, Kunkel, Wehner and Collins2016), to detect synoptic-scale weather fronts (cold front, warm front, and no front; Lagerquist et al., Reference Lagerquist, Mcgovern and Gagne2019) and for next-hour tornado prediction from radar images (Lagerquist et al., Reference Lagerquist, McGovern, Homeyer, Gagne and Smith2020).

Performing classification on each pixel of an image is called semantic segmentation in ML. Most CNN models for semantic segmentation use an encoder–decoder architecture, including the popular U-net deep CNN model (Ronneberger et al., Reference Ronneberger, Fischer and Brox2015), giving classification (or regression) on individual output pixels. U-net has been applied to rain-type classification (no-rain, stratiform, convective, and others) using microwave satellite images (Choi and Kim, Reference Choi and Kim2020), to precipitation estimation using satellite infrared images (Sadeghi et al., Reference Sadeghi, Phu, Hsu and Sorooshian2020) and to cloud cover nowcasting using visible and infrared images (Berthomier et al., Reference Berthomier, Pradel and Perez2020).

As traditional MLP NN models require a very large number of weights when working with 2D images or 3D spatial fields, limited sample size often necessitates the compression of input variables to a modest number of principal components by PCA, a linear technique (Jolliffe, Reference Jolliffe2002). The introduction of the CNN model has drastically reduced the number of weights, so compression by PCA is not needed, and deep CNN models have noticeably improved the performance of ML methods when working with 2D or 3D spatial fields in ES since the mid-2010s.

PCA has been used to impute missing values in datasets (Jolliffe, Reference Jolliffe2002, Section 13.6) and U-net, with its encoder–decoder architecture, can replace PCA in this task. Using sea surface temperature (SST) data from two sources (NOAA Twentieth-Century Reanalysis [20CR] and Coupled Model Intercomparison Project Phase 5 [CMIP5]), Kadow et al. (Reference Kadow, Hall and Ulbrich2020) trained U-net models, which outperformed PCA and kriging methods in imputing missing SST values.

The CGAN (Figure 4) has also been used in ES: In atmospheric remote sensing, CGAN generated cloud structures in a 2D vertical plane in the satellite’s along-track direction (Leinonen et al., Reference Leinonen, Guillaume and Yuan2019). As an alternative to U-net in super-resolution applications, CGAN has been used to convert low-resolution unmanned aircraft system images to high-resolution images (Pashaei et al., Reference Pashaei, Starek, Kamangir and Berryhill2020).

The application of ML methods to climate problems has been impeded by the relatively small effective sample size in observational records. In transfer learning, ML models trained on a dataset with large sample size can transfer their learning to a different problem hampered by a relatively small sample size. Using a CNN to learn the El Niño-Southern Oscillation (ENSO) behavior from the dynamical models (CMIP5 climate model data for 2,961 months), then transferring the learning to observed data (by further training with 103 months of reanalysis data), Ham et al. (Reference Ham, Kim and Luo2019) developed a CNN model with better accuracy in ENSO prediction than the dynamical models.

4. Merging of Machine Learning and Physics

As some components of a physical model can be computationally expensive, ML methods have been developed to substitute for the physics: For atmospheric radiation in atmospheric general circulation models (GCMs), MLP NN models have been used to replace the equations of physics (Chevallier et al., Reference Chevallier, Morcrette, Cheruy and Scott2000; Krasnopolsky et al., Reference Krasnopolsky, Fox-Rabinovitz and Belochitski2008). For a simple coupled atmosphere–ocean model of the tropical Pacific, the atmospheric component has been replaced by an NN model (Tang and Hsieh, Reference Tang and Hsieh2002). Resolving clouds in a GCM would require high spatial resolution and prohibitive costs; hence, NN models, trained by a cloud-resolving model, have been used to supply convection parametrization in a GCM (Krasnopolsky et al., Reference Krasnopolsky, Fox-Rabinovitz and Belochitski2013; Brenowitz and Bretherton, Reference Brenowitz and Bretherton2018; Rasp et al., Reference Rasp, Pritchard and Gentine2018). Increasingly, ML methods are used to learn from high-resolution numerical models, then implemented as inexpensive parametrization schemes in GCMs.

In physics-informed machine learning, ML models can be solved satisfying the laws of physics, for example, conservation of energy, mass, and so forth. In the soft constraint approach, the physics constraints are satisfied approximately by adding an extra regularization term in the objective function of an NN model (Karniadakis et al., Reference Karniadakis, Kevrekidis, Lu, Perdikaris, Wang and Yang2021). Alternatively, in the hard constraint approach, the physics constraints are satisfied exactly by the NN architecture (Beucler et al., Reference Beucler, Pritchard, Rasp, Ott, Baldi and Gentine2021).

Initially developed in numerical weather prediction (NWP), data assimilation (DA) aims to optimally merge theory (typically a numerical model based on physics) and observations. The most common DA method used in NWP is variational DA—4D-Var (three spatial dimensions plus time) and 3D-Var (spatial dimensions only; Kalnay, Reference Kalnay2003). Hsieh and Tang (Reference Hsieh and Tang1998) noted that back propagation used in finding the optimal NN solution is actually the same technique as the backward integration of the adjoint model used in variational DA; hence, one could combine numerical and NN models in DA by solving a single optimization problem. For the three-component dynamical system of Lorenz (Reference Lorenz1963), Tang and Hsieh (Reference Tang and Hsieh2001) replaced one of the dynamical equations with an NN equation, and tested using variational assimilation to estimate the parameters of the dynamical and NN equations and the initial conditions. In recent years, there has been increasing interest in merging ML and DA using 4D-Var in a Bayesian framework (Bocquet et al., Reference Bocquet, Brajard, Carrassi and Bertino2020; Geer, Reference Geer2021) or using the ensemble Kalman filter (Brajard et al., Reference Brajard, Carrassi, Bocquet and Bertino2020). NN models have the potential to greatly facilitate the building/maintaining of the tangent linear and adjoint models of the model physics in 4D-Var (Hatfield et al., Reference Hatfield, Chantry, Dueben, Lopez, Geer and Palmer2021).

5. Summary and Conclusion

The recent growth of ML in ES has been fueled mainly by deep NN models (Camps-Valls et al., Reference Camps-Valls, Tuia, Zhu and Reichstein2021), especially deep CNN models, which have greatly advanced the application of ML to 2D or 3D spatial data, gradually replacing many standard techniques like multiple linear regression, PCA, and so forth. Furthermore, some of the new ML models (e.g., GAN or CGAN) are no longer merely nonlinear generalization of a traditional statistical method. Progress has also been made in rendering ML methods less opaque and more interpretable (McGovern et al., Reference McGovern, Lagerquist, Gagne, Jergensen, Elmore, Homeyer and Smith2019; Ebert-Uphoff and Hilburn, Reference Ebert-Uphoff and Hilburn2020). In climate science, where observational records are usually short, transfer learning has been able to utilize long simulations by numerical models for pretraining ML models. ML and physics have also been merging, for example, in (a) the increasing use of ML for parametrization in GCMs, with the ML model trained using data from high-resolution numerical models, (b) the implementation of physics constraints in ML models, and (c) the increasing interest in using ML in DA.

My overall perspective is that the evolution of ML in ES has two distinct phases, slow initial acceptance followed by exponential growth starting in the mid-2010s. If an assessment were made on the progress of ML in ES in 2010, one would have come to the conclusion that, with the exception of hydrology and remote sensing, ML was not considered mainstream in ES, as numerical models (or even statistical models) were far more transparent and interpretable than “black box” models from ML. It is encouraging to see the resistance to ML models waning, as new ML approaches have been increasingly successful in tackling areas of ES where numerical models and statistical models have been hampered.

Data Availability Statement

No data were used in this perspective paper.

Author Contributions

Conceptualization: W.W.H.; Methodology: W.W.H.; Visualization: W.W.H.; Writing—original draft: W.W.H.; Writing—review & editing: W.W.H.

Funding Statement

This work received no specific grant from any funding agency, commercial, or not-for-profit sectors.

Competing Interests

The author declares no competing interests exist.

Footnotes

Current address: 4028 Hopesmore Drive, Victoria, British Columbia V8N 5S9, Canada.

References

Abrahart, RJ, Kneale, PE and See, LM (eds.) (2004) Neural Networks for Hydrological Modelling. London: CRC Press.CrossRefGoogle Scholar
Berthomier, L, Pradel, B and Perez, L (2020) Cloud cover nowcasting with deep learning. In Tenth International Conference on Image Processing Theory, Tools and Applications (IPTA). Paris: IEEE, pp. 16. https://doi.org/10.1109/IPTA50016.2020.9286606Google Scholar
Beucler, T, Pritchard, M, Rasp, S, Ott, J, Baldi, P and Gentine, P (2021) Enforcing analytic constraints in neural networks emulating physical systems. Physical Review Letters 126, 098302.CrossRefGoogle ScholarPubMed
Blackwell, WJ and Chen, FW (2009) Neural Networks in Atmospheric Remote Sensing. Boston: Artech House.Google Scholar
Bocquet, M, Brajard, J, Carrassi, A and Bertino, L (2020) Bayesian inference of chaotic dynamics by merging data assimilation, machine learning and expectation-maximization. Foundations of Data Science 2, 5580.CrossRefGoogle Scholar
Brajard, J, Carrassi, A, Bocquet, M and Bertino, L (2020) Combining data assimilation and machine learning to infer unresolved scale parametrization. Philosophical Transactions of the Royal Society A: Mathematical, Physical, and Engineering Sciences 379, 20200086.CrossRefGoogle Scholar
Breiman, L (2001a) Random forests. Machine Learning 45, 532.CrossRefGoogle Scholar
Breiman, L (2001b) Statistical modeling: The two cultures. Statistical Science 16, 199215.CrossRefGoogle Scholar
Breiman, L, Friedman, J, Olshen, RA and Stone, C (1984) Classification and Regression Trees. New York: Chapman and Hall.Google Scholar
Brenowitz, ND and Bretherton, CS (2018) Prognostic validation of a neural network unified physics parameterization. Geophysical Research Letters 45, 62896298.CrossRefGoogle Scholar
Camps-Valls, G, Tuia, D, Zhu, XX and Reichstein, M (eds.) (2021) Deep Learning for the Earth Sciences: A Comprehensive Approach to Remote Sensing, Climate Science and Geosciences. Hoboken, NJ: Wiley.CrossRefGoogle Scholar
Chen, T and Guestrin, C (2016) XGBoost: A scalable tree boosting system. In KDD’16: Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pp. 785794. New York: Association for Computing Machinery.CrossRefGoogle Scholar
Chevallier, F, Morcrette, JJ, Cheruy, F and Scott, NA (2000) Use of a neural-network-based long-wave radiative-transfer scheme in the ECMWF atmospheric model. Quarterly Journal of the Royal Meteorological Society 126, 761776.CrossRefGoogle Scholar
Choi, Y and Kim, S (2020) Rain-type classification from microwave satellite observations using deep neural network segmentation. IEEE Geoscience and Remote Sensing Letters 18, 21372141.CrossRefGoogle Scholar
Cortes, C and Vapnik, V (1995) Support vector networks. Machine Learning 20, 273297.CrossRefGoogle Scholar
Crevier, D (1993) AI: The Tumultuous History of the Search for Artificial Intelligence. New York: Basic Books.Google Scholar
Ebert-Uphoff, I and Hilburn, K (2020) Evaluation, tuning, and interpretation of neural networks for working with images in meteorological applications. Bulletin of the American Meteorological Society 101, E2149E2170.CrossRefGoogle Scholar
Geer, AJ (2021) Learning earth system models from observations: Machine learning or data assimilation? Philosophical Transactions of the Royal Society A: Mathematical, Physical, and Engineering Sciences 379, 20200089.CrossRefGoogle ScholarPubMed
Goodfellow, I, Bengio, Y and Courville, A (2016) Deep Learning. Cambridge, MA: MIT Press.Google Scholar
Goodfellow, I, Pouget-Abadie, J, Mirza, M, Xu, B, Warde-Farley, D, Ozair, S, Courville, A and Bengio, Y (2014) Generative adversarial nets. In Advances in Neural Information Processing Systems, Vol. 27. Red Hook, NY: Curran Associates, Inc., pp. 26722680.Google Scholar
Ham, Y-G, Kim, J-H and Luo, J-J (2019) Deep learning for multi-year ENSO forecasts. Nature 573, 568572.CrossRefGoogle ScholarPubMed
Hatfield, S, Chantry, M, Dueben, P, Lopez, P, Geer, A and Palmer, T (2021) Building tangent-linear and adjoint models for data assimilation with neural networks. Journal of Advances in Modeling Earth Systems 13, e2021MS002521.CrossRefGoogle Scholar
Haupt, SE, Pasini, A and Marzban, C (eds.) (2009) Artificial Intelligence Methods in the Environmental Sciences. Dordrecht: Springer.CrossRefGoogle Scholar
Hsieh, WW (2001) Nonlinear principal component analysis by neural networks. Tellus A: Dynamic Meteorology and Oceanography 53, 599615.CrossRefGoogle Scholar
Hsieh, WW (2009) Machine Learning Methods in the Environmental Sciences. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hsieh, WW (2020) Improving predictions by nonlinear regression models from outlying input data. Preprint, arXiv:2003.07926.Google Scholar
Hsieh, WW and Tang, B (1998) Applying neural network models to prediction and data analysis in meteorology and oceanography. Bulletin of the American Meteorological Society 79, 18551870.2.0.CO;2>CrossRefGoogle Scholar
Isola, P, Zhu, J-Y, Zhou, T and Efros, AA (2017) Image-to-image translation with conditional adversarial networks. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), pp. 11251134. Available at https://openaccess.thecvf.com/content_cvpr_2017/html/Isola_Image-To-Image_Translation_With_CVPR_2017_paper.html?ref=https://githubhelp.com.CrossRefGoogle Scholar
Jolliffe, IT (2002) Principal Component Analysis, 2nd Edn. New York: Springer.Google Scholar
Kadow, C, Hall, DM and Ulbrich, U (2020) Artificial intelligence reconstructs missing climate information. Nature Geoscience 13, 408413.CrossRefGoogle Scholar
Kalnay, E (2003) Atmospheric Modeling, Data Assimilation and Predictability. Cambridge: Cambridge University Press.Google Scholar
Karniadakis, GE, Kevrekidis, IG, Lu, L, Perdikaris, P, Wang, S and Yang, L (2021) Physics-informed machine learning. Nature Reviews Physics 3, 422440.CrossRefGoogle Scholar
Kingma, DP and Welling, M (2014) Auto-encoding variational Bayes. Preprint, arXiv:1312.6114.Google Scholar
Kramer, MA (1991) Nonlinear principal component analysis using autoassociative neural networks. AIChE Journal 37, 233243.CrossRefGoogle Scholar
Krasnopolsky, VM (2013) The Application of Neural Networks in the Earth System Sciences: Neural Network Emulations for Complex Multidimensional Mappings. Dordrecht: Springer.CrossRefGoogle Scholar
Krasnopolsky, VM, Fox-Rabinovitz, MS and Belochitski, AA (2008) Decadal climate simulations using accurate and fast neural network emulation of full, longwave and shortwave, radiation. Monthly Weather Review 136, 36833695.CrossRefGoogle Scholar
Krasnopolsky, VM, Fox-Rabinovitz, MS and Belochitski, AA (2013). Using ensemble of neural networks to learn stochastic convection parameterizations for climate and numerical weather prediction models from data simulated by a cloud resolving model. Advances in Artificial Neural Systems 2013, 485913.CrossRefGoogle Scholar
Krizhevsky, A, Sutskever, I and Hinton, GE (2012) ImageNet classification with deep convolutional neural networks. In Advances in Neural Information Processing Systems, Vol. 25, pp. 10901098. RedHook, NY: Curran Associates, Inc.Google Scholar
Lagerquist, R, Mcgovern, A and Gagne, DJ II (2019) Deep learning for spatially explicit prediction of synoptic-scale fronts. Weather and Forecasting 34, 11371160.CrossRefGoogle Scholar
Lagerquist, R, McGovern, A, Homeyer, CR, Gagne, DJ II and Smith, T (2020) Deep learning on three-dimensional multiscale data for next-hour tornado prediction. Monthly Weather Review 148, 28372861.CrossRefGoogle Scholar
LeCun, Y, Bengio, Y and Hinton, G (2015) Deep learning. Nature 521, 436444.CrossRefGoogle ScholarPubMed
LeCun, Y, Boser, B, Denker, JS, Henderson, D, Howard, RE, Hubbard, W and Jackel, LD (1989) Backpropagation applied to handwritten zip code recognition. Neural Computation 1, 541551.CrossRefGoogle Scholar
Leinonen, J, Guillaume, A and Yuan, T (2019) Reconstruction of cloud vertical structure with a generative adversarial network. Geophysical Research Letters 46, 70357044.CrossRefGoogle Scholar
Liu, Y, Racah, E, Prabhat, , Correa, J, Khosrowshahi, A, Lavers, D, Kunkel, K, Wehner, M and Collins, W (2016) Application of deep convolutional neural networks for detecting extreme weather in climate datasets. Preprint, arXiv:1605.01156.Google Scholar
Lorenz, EN (1963) Deterministic nonperiodic flow. Journal of the Atmospheric Sciences 20, 130141.2.0.CO;2>CrossRefGoogle Scholar
Maier, HR and Dandy, GC (2000) Neural networks for the prediction and forecasting of water resources variables: A review of modelling issues and applications. Environmental Modelling and Software 15, 101124.CrossRefGoogle Scholar
McGovern, A, Lagerquist, R, Gagne, DJ II, Jergensen, GE, Elmore, KL, Homeyer, CR and Smith, T (2019) Making the black box more transparent: Understanding the physical implications of machine learning. Bulletin of the American Meteorological Society 100, 21752199.CrossRefGoogle Scholar
Mirza, M and Osindero, S (2014) Conditional generative adversarial nets. Preprint, arXiv:1411.1784.Google Scholar
Monahan, AH (2000) Nonlinear principal component analysis by neural networks: Theory and application to the Lorenz system. Journal of Climate 13, 821835.2.0.CO;2>CrossRefGoogle Scholar
Nilsson, NJ (2009) The Quest for Artificial Intelligence. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Pashaei, M, Starek, MJ, Kamangir, H and Berryhill, J (2020) Deep learning-based single image super-resolution: An investigation for dense scene reconstruction with UAS photogrammetry. Remote Sensing 12, 1757.CrossRefGoogle Scholar
Rasp, S, Pritchard, MS and Gentine, P (2018) Deep learning to represent subgrid processes in climate models. Proceedings of the National Academy of Sciences of the United States of America 115, 96849689.CrossRefGoogle ScholarPubMed
Ronneberger, O, Fischer, P and Brox, T (2015) U-net: Convolutional networks for biomedical image segmentation. In Medical Image Computing and Computer-Assisted Intervention (MICCAI 2015), Lecture Notes in Computer Science, Vol. 9351. Cham, Switzerland: Springer, pp. 234241.CrossRefGoogle Scholar
Rumelhart, DE, Hinton, GE and Williams, RJ (1986) Learning internal representations by error propagation. In Rumelhart, D, McClelland, J and PDP Research Group (eds.), Parallel Distributed Processing, Vol. 1. Cambridge, MA: MIT Press, pp. 318362.CrossRefGoogle Scholar
Sadeghi, M, Phu, N, Hsu, K and Sorooshian, S (2020) Improving near real-time precipitation estimation using a U-net convolutional neural network and geographical information. Environmental Modelling and Software 134, 104856.CrossRefGoogle Scholar
Tang, Y and Hsieh, WW (2001) Coupling neural networks to incomplete dynamical systems via variational data assimilation. Monthly Weather Review 129, 818834.2.0.CO;2>CrossRefGoogle Scholar
Tang, Y and Hsieh, WW (2002) Hybrid coupled models of the tropical Pacific: II ENSO prediction. Climate Dynamics 19, 343353.Google Scholar
Yuval, and Hsieh, WW (2002) The impact of time-averaging on the detectability of nonlinear empirical relations. Quarterly Journal of the Royal Meteorological Society 128, 16091622.CrossRefGoogle Scholar
Figure 0

Figure 1. Venn diagram illustrating the relation between artificial intelligence, statistics, machine learning, neural networks, and deep learning, as well as kernel methods, random forests, and boosting.

Figure 1

Figure 2. The encoder–decoder is an NN model with the first part (the encoder) mapping from the input x to u, the “code” or bottleneck, and the second part (the decoder) mapping from u to the output y. Dimensional compression is achieved by forcing the signal through the bottleneck. The encoder and the decoder are each illustrated with only one hidden layer for simplicity.

Figure 2

Figure 3. Generative adversarial network with the generator creating a fake image (e.g., a fake Picasso painting) from random noise input, and the discriminator classifying images as either real or fake. Whether the discriminator classifies a fake image rightly or wrongly leads, respectively, to further training for the generator or for the discriminator.

Figure 3

Figure 4. Conditional generative adversarial network where the generator G receives an image x and a random noise vector z as input. The discriminator D receives x plus either a fake image from G (left) or a real image y (right) as input. Here, a line drawing is converted to a photo image; similarly, a photo image can be converted to a line drawing.Adapted from Figure 2 of Isola et al. (2017).