GRRIEn Analysis: A Data Science Cheat Sheet for Earth Scientists Learning from Global Earth Observations

Elizabeth Carter aCivil and Environmental Engineering, Syracuse University, Syracuse, New York

Search for other papers by Elizabeth Carter in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0002-4920-0973
,
Carolynne Hultquist bSchool of Earth and Environment, University of Canterbury, Christchurch, New Zealand

Search for other papers by Carolynne Hultquist in
Current site
Google Scholar
PubMed
Close
, and
Tao Wen cEarth and Environmental Sciences, Syracuse University, Syracuse, New York

Search for other papers by Tao Wen in
Current site
Google Scholar
PubMed
Close
Open access

We are aware of a technical issue preventing figures and tables from showing in some newly published articles in the full-text HTML view.
While we are resolving the problem, please use the online PDF version of these articles to view figures and tables.

Abstract

Globally available environmental observations (EOs), specifically from satellites and coupled Earth system models, represent some of the largest datasets of the digital age. As the volume of global EOs continues to grow, so does the potential of these data to help Earth scientists discover trends and patterns in Earth systems at large spatial scales. To leverage global EOs for scientific insight, Earth scientists need targeted and accessible exposure to skills in reproducible scientific computing and spatiotemporal data science, and to be empowered to apply their domain understanding to interpret data-driven models for knowledge discovery. The Generalizable, Reproducible, Robust, and Interpreted Environmental (GRRIEn) analysis framework was developed to prepare Earth scientists with an introductory statistics background and limited/no understanding of programming and computational methods to use global EOs to successfully generalize insights from local/regional field measurements across unsampled times and locations. GRRIEn analysis is generalizable, meaning results from a sample are translated to landscape scales by combining direct environmental measurements with global EOs using supervised machine learning; robust, meaning that the model shows good performance on data with scale-dependent feature and observation dependence; reproducible, based on a standard repository structure so that other scientists can quickly and easily replicate the analysis with a few computational tools; and interpreted, meaning that Earth scientists apply domain expertise to ensure that model parameters reflect a physically plausible diagnosis of the environmental system. This tutorial presents standard steps for achieving GRRIEn analysis by combining conventions of rigor in traditional experimental design with the open-science movement.

Significance Statement

Earth science researchers in the digital age are often tasked with pioneering big data analyses, yet have limited formal training in statistics and computational methods such as databasing or computer programming. Earth science researchers often spend tremendous amounts of time learning core computational skills, and making core analytical mistakes, in the process of bridging this training gap, at risk to the reputability of observational geostatistical research. The GRRIEn analytical framework is a practical guide introducing community standards for each phase of the computational research pipeline (dataset engineering, model training, and model diagnostics) to promote rigorous, accessible use of global EOs in Earth systems research.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Elizabeth Carter, ekcarter@syr.edu

Abstract

Globally available environmental observations (EOs), specifically from satellites and coupled Earth system models, represent some of the largest datasets of the digital age. As the volume of global EOs continues to grow, so does the potential of these data to help Earth scientists discover trends and patterns in Earth systems at large spatial scales. To leverage global EOs for scientific insight, Earth scientists need targeted and accessible exposure to skills in reproducible scientific computing and spatiotemporal data science, and to be empowered to apply their domain understanding to interpret data-driven models for knowledge discovery. The Generalizable, Reproducible, Robust, and Interpreted Environmental (GRRIEn) analysis framework was developed to prepare Earth scientists with an introductory statistics background and limited/no understanding of programming and computational methods to use global EOs to successfully generalize insights from local/regional field measurements across unsampled times and locations. GRRIEn analysis is generalizable, meaning results from a sample are translated to landscape scales by combining direct environmental measurements with global EOs using supervised machine learning; robust, meaning that the model shows good performance on data with scale-dependent feature and observation dependence; reproducible, based on a standard repository structure so that other scientists can quickly and easily replicate the analysis with a few computational tools; and interpreted, meaning that Earth scientists apply domain expertise to ensure that model parameters reflect a physically plausible diagnosis of the environmental system. This tutorial presents standard steps for achieving GRRIEn analysis by combining conventions of rigor in traditional experimental design with the open-science movement.

Significance Statement

Earth science researchers in the digital age are often tasked with pioneering big data analyses, yet have limited formal training in statistics and computational methods such as databasing or computer programming. Earth science researchers often spend tremendous amounts of time learning core computational skills, and making core analytical mistakes, in the process of bridging this training gap, at risk to the reputability of observational geostatistical research. The GRRIEn analytical framework is a practical guide introducing community standards for each phase of the computational research pipeline (dataset engineering, model training, and model diagnostics) to promote rigorous, accessible use of global EOs in Earth systems research.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Elizabeth Carter, ekcarter@syr.edu

1. Introduction

The past 50 years have ushered in exponential growth in the volume of Earth observations. As of writing, there are over 300 Earth-observing satellites in operation by national space agencies globally, with another 79 platforms approved or in development for the next decade (Committee on Earth Observing Satellites 2022). Increased availability of telecommunications bandwidth and lowered costs of electronic components have led to a proliferation of in situ automatic Earth-monitoring networks (Stephens et al. 2020; Balsamo et al. 2018). Scaffolding the storage and processing of all these data, we have seen a more than a trillion-fold increase in global computer power in the last 50 years (Tredinnick and Laybats 2018). This shift in information content of environmental systems has led to a shift in research methods. To quote from the Stanford Earth Matters magazine: “The satellite and supercomputer are the rock, hammer, and compass of modern geoscientists.” Even though modern Earth scientists are often tasked with pioneering big-data analysis, of the top-10-ranked undergraduate programs in Earth science (U.S. News and World Report 2022), the majority require one or fewer semesters of coursework in probability and statistics, and none have required coursework in computational methods such as databasing or computer programming.

In the past several decades, the research community has proposed many guidelines to promote scientific rigor in observational computational research. One such framework that is widely accepted by the research community is Open Science by Design, as proposed by the National Academies of Sciences, Engineering, and Medicine (2018). Open Science by Design guides stakeholders to practice “open science,” a movement to make scientific research reproducible and accessible, throughout the entire research life cycle (ideation, knowledge generation, validation, dissemination, preservation). The same concept was later adopted by the Earth science community to form the Integrates, Coordinates, Openly, Networks (ICON) principles (Goldman et al. 2021) to promote open science approaches in Earth science with a particular emphasis on the need for growing open interdisciplinary collaboration. Both concepts emphasize that it is important for researchers to make research results follow the principles of Findability, Accessibility, Interoperability, Reusability (FAIR) (Wilkinson et al. 2016), which facilitates peer review of the entire computational research pipeline for improved research quality.

Generalizable, Robust, Reproducible, and Interpreted Environmental (GRRIEn, pronounced “grēn” like the color) supervised learning using global Earth observations (EOs) analysis builds on FAIR standards to empower Earth scientists to generate high-quality, easily reproducible geostatistical workflows utilizing openly available global geospatial data. We do this by outlining best management practices for each step in the computational analysis pipeline: dataset engineering, model training, and model diagnostics (Fig. 1).

Fig. 1.
Fig. 1.

GRRIEn analysis introduces best management practices to extract generalizable insight into environmental systems using supervised learning of global EOs, including standards for reproducible data engineering, robust model training, and domain specialist–interpreted model diagnostics. Figure adapted from National Academies of Sciences, Engineering, and Medicine (2018).

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

GRRIEn analysis applies to the context in which Earth scientists use global EOs to generalize insights from limited Earth system measurements to unsampled times and locations. We define major types of global EOs, classify the primary objectives when using global EOs in supervised learning, and briefly describe how global EOs can be reliably converted into analysis-ready data using geodatabase application programming interfaces (APIs) and open-source programming languages (section 2). We describe the two spatiotemporal data pitfalls, scale-dependent observation and feature dependency, detail how they impact the robustness of supervised modeling frameworks for the different modeling objectives, and present checklists for diagnosis and model-agnostic management of these pitfalls in dataset engineering and model training (appendix C, section 3, in the online supplemental material). Drawing from a suite of most-used tools in reproducible computational research, we propose a standard software repository structure to facilitate a highly adaptive, highly reproducible data sourcing, engineering, and modeling pipeline for replication of diverse supervised GRRIEn analysis workflows (section 4). Finally, we describe how experimental design principles translate into the era of global EOs, give an overview of explainable machine learning and artificial intelligence (AI), and explain the critical role of the modern Earth scientists in interpreting the physical plausibility of trained data-driven models (section 5). We conclude with how to incorporate GRRIEn into the experimental design and manuscript outline process and describe limitations of the method and future work (section 6).

Supplemental appendix A, “Required (Computational) Tools,” is a detailed outline of the assumed background knowledge for GRRIEn analysis. For each topic, we include resources for where to find more information and learn new skills. We strongly suggest reading this supplemental resource, and reviewing any background concepts as necessary, prior to learning the GRRIEn method.

2. Generalizable

Earth science disciplines have evolved in a historically data-poor environment. Early–mid-twentieth century geoscientists relied primarily on data from short-term site- or laboratory-based experiments (Clark and Gelfand 2006) and then used process models to infer modes of spatiotemporal variability across unsampled times and locations (Hilborn and Mangel 2013). A process model functions like an algorithmic narrative, merging theory and available data to quantitatively articulate and compare plausible hypotheses concerning drivers of spatiotemporal environmental variability (Hilborn and Mangel 2013). As one cannot properly quantify the importance of unparameterized, underparameterized, or inaccurately parameterized phenomena using process-based models, this method has a strong potential for perpetuating confirmation bias in Earth systems theories (Nickerson 1998; Bond et al. 2007; Shi et al. 2019). Global EOs, defined here as spatially continuous, temporally repeating data with continental to global coverage, are outcomes of an international movement to improve characterization of Earth systems (Nativi et al. 2015). As global EOs are approximately spatiotemporally continuous, they provide an observation-based framework to translate insight from limited field measurements across unsampled times and locations to validate theories of spatiotemporal drivers of environmental processes. For example, characterizing spatiotemporal variability in precipitation using point data from rain gauges is a highly socially relevant field of study that has been a major thread in hydrological study for decades (Kidd et al. 2017). In many gridded precipitation datasets, output from numerical weather models is used to spatially interpolate precipitation between gauges (e.g., Hersbach et al. 2020). Recent studies show that satellite observations, while containing independent sources of error, can help elucidate the nature of specific biases in precipitation estimation from numerical weather models (Xu et al. 2022). This is one of many examples of how global EOs have facilitated the evolution of theories of environmental systems that are described in process models. The task ahead is straightforward: how can we accurately use these global EOs as proxies of environmental processes to augment our understanding of how and why diverse Earth systems vary across space and time?

a. Getting started with global Earth observations

The diversity and scope of Earth observations are staggering and ever expanding (Balsamo et al. 2018), but to get started, we recommend a working knowledge of the following three data types that provide global coverage at a regular spatial and temporal scale: active satellite remote sensing data, passive satellite remote sensing data, and coupled Earth systems model (CESM) output (Fig. 2) More detail on these data types can be found in supplemental appendix A, section 1.

Fig. 2.
Fig. 2.

Three common sources of global EOs include active satellite remote sensing data, such as (left) synthetic aperture radar imagery; (center) passive satellite remote sensing data, including optical and passive microwave imagery; and (right) gridded outputs from global coupled Earth system models.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

b. Objectives of using global EOs in environmental systems analysis

In supervised learning, models are trained on coupled records of both input variables (predictors), and output variable(s) [label(s) or predictand(s)]. The GRRIEn framework deals specifically with supervised learning. It is meant to be applied to contexts where some measurements of an environmental process (predictand) are available, but the experimental questions cover a larger spatial extent or a different period of time and/or are at a different sampling frequency than the measured data. As global EOs are (approximately) spatiotemporally continuous observations, the overall objective of GRRIEn analysis is to train a supervised learning algorithm that can predict an environmental process using globally available EOs as input data. This algorithm can be used to generalize insights from direct observations of an environmental process sampled experimentally or in situ (hereafter called the label or output variable in a trained algorithm) across unlabeled times or locations where global EOs are available (see examples in Table 1). The prediction space of the generalizing algorithm will be bounded by an area of interest (AOI) and a period of interest (POI) that may or may not overlap with the labeled sample. Global EOs serving as predictors (i.e., input data to the predictive model) provide full coverage of the POI and AOI, whether or not it is possible to directly observe the environmental process at that scale (Fig. 3).

Table 1.

Examples of supervised learning for interpolation, extrapolation, and diagnostic modeling using globally available EOs as input.

Table 1.
Fig. 3.
Fig. 3.

(a) In supervised learning, a statistical model is trained that maps variability in input data to variability in a limited number of labels (in situ or experimental measurements) which are contemporaneous in space and/or time. (b) The trained model can then be used (top) for interpolation (filling in the gaps in spatiotemporal variability in the output for an AOI and POI that overlaps with the training data; (middle) for extrapolation (predicting in an AOI and/or POI that exceeds the bounds of the training data); and/or (bottom) to diagnose drivers and modes of spatial coverability between inputs and labels.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

There are three primary objectives when using global EOs in supervised learning: interpolation (prediction in an AOI or POI intersecting with sample space), extrapolation (prediction in an AOI or POI that does not intersect with the labeled sample space), or diagnostic modeling (using the trained algorithm as data to gain insight into the nature, causes, and associations of space/time variability in the environmental process) (Fig. 3). Table 1 provides detailed explanations and sample studies using global EOs as predictors in supervised analysis for interpolation, extrapolation, and diagnostic modeling.

3. Robust

In experimental research, the goal is to collect a representative sample to allow for the statistical evaluation of specific hypothesis(es) about independent drivers of variability in a system. Experimental researchers start with a highly controlled environment, such as a laboratory or field space, which will hold constant all physical variables deemed important by domain scientists, except the variables that will be experimentally varied (the predictors). Good experimental design ensures that data will be collected across the full range of possible values for each predictor variable. In addition, in multivariate experiments, independent representation of predictors is ensured by collecting data across all potential combinations of values of predictor variables. For example, in an experiment to determine how radiation and soil moisture (the input variables or predictors) impact plant growth (the output variable or predictand), low radiation–high soil moisture, low radiation–low soil moisture, high radiation–high soil moisture, and high radiation–high soil moisture treatments should be included. The number of samples to be collected across treatments is defined a priori based on the desired confidence in results (e.g., Campbell et al. 1969).

The field of statistics was developed to quantify the certainty of results from controlled experimental trials, and machine learning and deep learning are subfields of statistics (Runge et al. 2019; Blei and Smyth 2017). Because of this, all observational environmental data, especially spatial and temporal observational data collected at large scales, have characteristic divergences from data collected in controlled experimental trials, some of which must be addressed in order for data science algorithms to generalize well. Here, we discuss three sources of such divergences: inability to parameterize all drivers of variability in an environmental system, limitations to ensuring independence of observations sampled in time and space, and limitations to sampling for independence of multivariate predictors.

First, in observational experiments, we lack controlled environments. Many landscape phenomena will change at similar spatial and temporal scales as the environmental process of interest. If any feature driving variability in the environmental process is not parameterized (i.e., included as a predictor) in the model, that model is subject to omitted variable bias: something important is happening to the environmental process, and the impact is observable in our response variable, but it is not parameterized in the model.

Second, in observational data science, we often lack the ability to collect balanced, cross-replicated samples of multivariate predictors. This is because in environmental systems, multivariate predictors tend to be intrinsically related to, or dependent on, each other. For example, since rain comes from cloudy skies, there will be more observations of low radiation–high soil moisture conditions, and few to no observations of high radiation–high soil moisture conditions, in observational plant growth data (Carter et al. 2018b). This is called feature dependence, or a condition where we have some structure of covariability between individual predictors in our training data.

Third, especially when relying on global EOs which are collected at standard temporal and spatial intervals, we do not get to define a priori how many observations are collected per unit space or time to confirm or deny a predefined hypothesis. Since observations which are nearer to each other in space or in time tend to be related, our observations, which are sampled at regular intervals in space and time, will not be independent. Implicit observation dependence in spatiotemporal data means that our sampling frequency may either obscure or overrepresent the modes of spatiotemporal variability that we are trying to evaluate in our model.

In large-scale spatiotemporal systems, unaddressed scale-dependent feature and observation dependence can severely impact interpolation, extrapolation, and interpretation of supervised models. When working with global EOs, it is therefore important that we take certain steps to characterize feature and observation dependence in our datasets, and address it in dataset engineering and model training, to avoid misinterpretation of our results (Fig. 4) Please note that a checklist for robust spatiotemporal models can be viewed at the end of this section (Fig. 10), as well as a flowchart for diagnosing model-performance issues using the spatial distribution of model residuals (Fig. 12). These figures are referenced throughout the section.

Fig. 4.
Fig. 4.

The robust rules for GRRIEn analysis.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

a. Robust to dependent features

There are two main sources of feature dependence in observational datasets. Intrinsic feature dependence occurs when measurements are imperfect representations of an underlying latent process that cannot be measured directly. For example, we cannot directly measure soil texture, but we can measure percent of sand, silt, and clay in a soil sample. All these variables will be intrinsically correlated (a higher ratio of sand implies a lower ratio of silt/clay), yet each give different information on the unmeasurable quality of soil texture. Incidental feature dependence is caused by our inability to capture a representative sample. With incidental correlation, not all combinations of predictor variables exist in a study area (e.g., the lack of cross replication between radiation and soil moisture mentioned in the introduction to section 3).

Multicollinearity is a special case of feature dependence that occurs when one or more predictor variables (features) are linearly related. To understand how multicollinearity impacts inference and prediction in machine learning algorithms, we should start by reviewing how multicollinearity impacts ordinary least squares (OLS) linear regression. In OLS regression, the standard error around coefficient estimates on collinear predictors is increased. The stronger the linear relationship between the variables, the larger the inflation of standard error on their coefficient estimates. This means that any single realization of a coefficient estimate, derived from a sample, is likely to be further from the unknown “true” value that represents the linear dependence between the predictor and predictand across the whole population. Multicollinearity therefore leads to model instability, or a situation where small changes in training data can lead to large changes in model coefficient estimates.

Model instability is a problem for several reasons. To start, when two collinear predictor variables are included in a model, we cannot statistically identify which variable has a direct association or a causal impact with the predictand (Dormann et al. 2013). So, given two predictors, one of which has a direct association with the predictand, and one of which has no direct physical association with the predictand but is incidentally correlated with the other predictor, no statistical model can determine which is the causal driver. This issue impacts all data-driven models, not just linear regression (Dormann et al. 2013; Alin 2010; Feng et al. 2019; Kim 2019). The performance of the model no longer depends on the marginal distribution of individual independent predictor variables, but on the joint distribution of dependent predictor variables. If we enter into a prediction space where the two predictors are no longer incidentally correlated, the model may make inaccurate predictions, specifically if the model attributed substantial weight to the second, incidentally correlated predictor. In other words, with substantial feature dependency in input data, the model performance is contingent on the structure of multicollinearity that was present in the training dataset being conserved across the multivariate population as a whole. This means that even if a trained model shows good in-sample model fit statistics (i.e., our predictions are similar to our available data), if the model is being used to make predictions with a dataset where the correlation structure between predictors is different than it was during model training, such as we would expect with incidental feature dependence, these predictions are likely to be inaccurate. This is ultimately because parameters in an unstable model are substantially less likely to represent physically plausible characterizations of the relationship between individual predictor variables and the predictand (section 3c: checking your work; Fig. 12a).

This can be problematic in environmental systems analysis, as the structure of collinearity between predictor variables is often dynamic in both space and time (Fig. 5). For example, all meteorological variables will exhibit some degree of collinearity, and the strength and sign (positive or negative) of collinearity between meteorological variables will change over space (as a signature of local climate) as well as over time (as a signature of specific seasonal or weather patterns) (Dormann et al. 2013; Thornton et al. 1997; Carter et al. 2018a). Since we have a limited ability to directly sample environmental processes across landscape scales where we might capture this variability in feature dependence, in observational environmental systems analysis, incidental multicollinearity in training data is a common problem that cannot always be avoided. In Fig. 5a, this is demonstrated in the spatial variability in local correlation between summertime air temperature and monthly precipitation from 1950 to 2000 across the United States. This covariance has a physical cause: as it gets drier, incident solar radiation that would have been used to evaporate soil moisture is partitioned to sensible heat flux, increasing air temperatures. The correlation coefficient changes spatially because the specific covariance structure of meteorological variables is a signature of local climate (supplemental appendix 2, Fig. B1). Under climate change, we expect the exact structure of local covariability of meteorological variables to shift as well, as is demonstrated in Fig. 5b, which shows the local correlation coefficient between ensemble mean summertime and temperature and monthly precipitation projected for 2050–99 under a moderate emissions scenario in the CMIP5 multimodel ensemble.

Fig. 5.
Fig. 5.

Pearson’s correlation coefficient [scaled between −1 and 1; color bar (Benesty et al. 2009)] between bias-corrected statistically downscaled Climate Model Intercomparison Project 5 ensemble mean monthly precipitation and daily max temperature. (left) Historical observations for 1950–99 and (right) moderate emissions forecast (RCP 4.5) for 2050–99 both indicate spatiotemporal variability collinearity between summertime max temperature and precipitation. Covariance of meteorological variables is a signature of local climate. As local climates shift due to global warming, so will the local covariability of meteorological variables (see right panel). This generates complexity for predicting environmental process response to meteorological variables under climate change (Taylor et al. 2012).

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Substantial changes in local correlation between summertime air temperature and precipitation under climate change also have a physical explanation, as we expect complex, nonlinear changes in the dependency structure between meteorological fields as air temperatures increase. For example, local summertime air temperatures will be increasing because of increased longwave absorptivity of the atmosphere (a different physical driver than meteorological drought), and precipitation will shift due to shifting atmospheric circulation and hydrologic intensification. We cannot collect a sample of meteorological data that is representative of the structure of covariability under climate change because these conditions do not yet exist. As the specific correlation structure between meteorological variables changes over space and time, and since the prediction skill of data-driven algorithms depends on the conservation of the correlation structure of predictor variables in the prediction space, we are likely to see spatiotemporal structure in the error of observational algorithms trained with meteorological variables as predictors, specifically in extrapolation (Dormann et al. 2013; Carter et al. 2018a). Since the structure of collinearity changes over space and time in many environmental processes, even within the training data, multicollinearity must be carefully assessed during routine exploratory data analysis and cannot safely be ignored during model training (Dormann et al. 2013). Because of this, we often talk about two metrics of model performance in a multicollinear system: model accuracy (how well predictions match labels) and model stability (how consistently variations in specific input variables map to changes in predictions between training samples and model parameterizations) (Graham 2003).

1) Diagnosing and data engineering with feature dependence

(i) Quantify multicollinearity globally and locally

With multivariate spatial datasets, multicollinearity must be evaluated both globally (assuming all observations represent a single population) and locally (treating different regions and/or time periods as unique populations). Global multicollinearity can be visualized by looking at a scatterplot matrix (pairwise correlation coefficients) or quantified by way of the dataset’s variance inflation factors (VIFs), condition numbers (CNs) (Alin 2010), or variance decomposition proportions (VDs) (Brauner and Shacham 1998). Multicollinearity can also be diagnosed locally by calculating geographically weighted VIF, CN, or VD at different spatial bandwidths (Kalogirou 2013; Wheeler 2009; Lu et al. 2014). Researchers should note substantial variability in locally calibrated metrics of multicollinearity and be transparent that these multicollinearities could affect model interpolation, and prediction, as well as interpretation/explanation.

(ii) Select representative training/testing data

To avoid overfitting when training algorithms, it is important to divide the entire pool of matched label/predictand observations into training data (data the model parameters are calibrated to), and testing data (data the model predictions are evaluated against). For iteratively trained algorithms or while hyperparameter tuning, validation data subsets are sampled from the training data to evaluate the skill of evolving model formulations (Berrar 2019). To avoid incidental multicollinearity, training/validation and/or testing data should contain a range of values between the minimum and maximum (max) expected value for each input and output variable, with both the marginal and joint distributions of the sample reflecting the population as a whole. When there is spatial or temporal variability in the structure of multicollinearity in the dataset, training data should also be strategically sampled, and when appropriate, training/validation/testing subsetting should be strategically designed to be representative of the range of covariability in input variables. This often means stratifying your sampling area to select diverse examples of different regional or temporal representations of input variable collinearity (positive or negative, weak or strong) (Tamura et al. 2017). This way, we can quantify the model’s mean performance across the training or training/validation sample, as well as on anomalous examples that may be present in the population as a whole.

In observational analysis, researchers often do not have control over training sample collection. When it is not feasible to collect representative training/validation data, testing data, which are data withheld from training and used to evaluate model stability, should always reflect anomalous (from global) spatiotemporal collinearity (Salazar et al. 2022). For example, the global Pearson’s correlation coefficient of two variables is 0.7 in a 1000-km2 study area, but the local correlation with a bandwidth of 5 km can range from −0.8 to 0.95. You would want to include regions with a strong negative, neutral, and strong positive correlation in your testing data.

When thinking about your intended extrapolation contexts, it is important to apply domain expertise to evaluate whether there might be shifts in the spatiotemporal collinearity of your predictors that is not represented in your training data. Are you planning to make predictions in regions or times where other factors, like climate change, might change the relationship between your input variables? Shifts in the correlation structure of input variables between your training data and your AOI or POI will likely impact the fidelity of extrapolations (Fig. 12b) (Dormann et al. 2013).

(iii) Reduce the number of input variables

Unnecessary intrinsic multicollinearity, where several collinear predictors are included in the model with no known relationship to the predictand, can undermine model accuracy, stability, and interpretability and should be avoided as a rule. In observational analysis, it is rare to have access to perfectly cross-replicated measurements of known or suspected drivers of variability in an environmental process. Instead, we have to compromise and use what observations we have available as “proxies” for the things we want to measure. When making use of proxy data, there is a difference between making scientifically informed choices about which proxies are important, and just throwing all available data into your model and letting it decide for you (a process commonly known as “data mining”). At this point in the big data revolution, a well-trained domain scientist will do a better job of picking important variables than a supercomputer. Step zero in addressing multicollinearity in your dataset is “feature engineering,” or using your domain expertise to select only the most physically important variables. The inclusion of any predictor variable in a data science model should be motivated by established theory and literature in your field. Think of each of your predictor variables as a hypothesis that you would like to test regarding what physical drivers or measurements are most relevant to your environmental system. If your meaningful predictors are intrinsically collinear, consider factor reduction [a review of methods can be found in Chan et al.(2022)], or use elastic net regularization (see below). Stepwise selection, where variables are chosen by evaluating change in prediction error when variables are either included or omitted from the suite of predictors, is strongly discouraged in collinear datasets, as it is likely that model instability may lead to rejection of important variables (Smith 2018).

2) Training and validating models with feature dependence

(i) Train for model stability as well as model fit

Prediction skill (and its opposite, error) can be decomposed into two components: accuracy (variance) and precision (bias). The sum of variance and bias in predictions is the total error of the model (Fig. 6). The default objective of most model-fitting protocols, like OLS regression, is to minimize total error in the model predictions of an unbiased model. When multicollinearity is present, models trained to minimize bias will tend toward being overfit. An overfit model has poor generalization, i.e., model parameters describe noise in the training data, not patterns in the population as a whole. Thus, overfit models yield inaccurate out-of-sample predictions (see Fig. 12a).

Fig. 6.
Fig. 6.

Total model skill (and its opposite, error) can be decomposed into accuracy (variance) and precision (bias). To avoid overfitting and to minimize total error in models with multicollinear data, we intentionally add bias to our model parameters using regularization parameters. Figure adapted from Bigbossfarin, CC0, via Wikimedia Commons.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Regularization parameters, like ridge (Hoerl and Kennard 1970), lasso (Tibshirani 1996), and elastic net [a linear combination of lasso and ridge parameters (Zou and Hastie 2005)], are integrated into statistical (Dormann et al. 2013), machine learning (Li et al. 2021; Carter et al. 2021; Mudereri et al. 2019), and deep learning models (Murugan and Durairaj 2017; Versloot 2020) to shrink model estimators. Regularization is therefore a systematic bias that we add to models to stabilize parameter estimates and reduce overfitting. We do this because we know that the lowest total error of a model will occur somewhere between the lowest-variance model and the lowest-bias model (Fig. 6, right). A biased model might miss some of the patterns in the data, but since it will not fit noise, it often generates a more physically plausible representation of the system and therefore yields more accurate interpolations and extrapolations (Dormann et al. 2013). The ridge and lasso regularization parameters contain coefficients which can be tuned. Increasing or decreasing the coefficient on your regularization parameters increases or decreases the amount of regularization, which increases or decreases the bias in your parameter estimates associated with collinearity, in order to increase model stability and minimize total error.

Different regularization techniques behave in different ways with multicollinearity. The ridge parameter achieves a “grouping effect”: if there is a group of highly collinear predictors, it will effectively partition the magnitude of response variance more equally between each collinear predictor variable, instead of arbitrarily all to one, improving model stability and reducing estimator bias. With a ridge penalty, unimportant coefficients will be reduced in magnitude, but no parameters will be shrunk to zero. As such, it is not directly suitable for feature selection. The lasso parameter, on the other hand, effectively eliminates uninfluential predictors by shrinking their associated coefficients to zero. Unlike ridge, however, it does not achieve the grouping effect, and therefore, the magnitude of coefficients on collinear predictors is likely to be subject to volatility. Elastic net regularization, which is a linear combination of ridge and lasso, achieves a “group selection effect,” producing physically plausible, stable coefficients that partition response variance within important groups of collinear predictors, while removing unimportant predictors from the model completely (achieving all-in-one factor reduction and model training). Elastic net shows consistent skill over other regularization and in situ factor reduction methods in model fit and stability in high-dimensional, collinear datasets when added to the loss functions of diverse machine learning and deep learning algorithms (Srisa-An 2021; Dormann et al. 2013). Supplemental appendix C provides an example of using regularization to manage multicollinearity.

(ii) Utilize cross-validation and ensemble learning

Models which are trained on independent, randomly permuted subsets of training/validation data in order to minimize out-of-sample prediction error are more stable and robust models under multicollinearity. Among these, models which use ensemble learning to integrate insights from random subsampling [e.g., bagged decision trees (Breiman 1996a), random forest (Breiman 2001), stacked models (Breiman 1996b), AdaBoost (Freund and Schapire 1999), and gradient boosting machines (Friedman 2001)] perform better than those that use iterative learning (Wang et al. 2021; Hembram et al. 2021; Adnan et al. 2020; Smith et al. 2013). For example, random forest regression, a bootstrapped ensemble learning method, is consistently shown to have higher stability and prediction accuracy than other machine learning/deep learning algorithms, such as naive bayes, boosted regression trees (Hembram et al. 2021), support vector machines (Adnan et al. 2020), and artificial neural networks (Smith et al. 2013) that train iteratively, when used on collinear observational datasets.

(iii) Account for space–time variability in feature dependence

Models that calculate local (to space or time) parameters or weights are more stable and robust with dynamic patterns of multicollinearity, than are models which rely on global parameter estimates (Mahadi et al. 2022; Carter et al. 2018b; Wen et al. 2018). Some examples of locally calibrated models include geographically weighted and time-varying regression, in which both regression parameters and regularization parameters can be locally calibrated (Murakami et al. 2021; Li and Lam 2018; Wheeler 2009; Kalogirou 2013; Bárcena et al. 2014; Mahadi et al. 2022), and convolutional neural networks, which are common in computer vision and can learn complex spatial as well as spectral patterns in multidimensionally gridded datasets (such as remotely sensed imagery or multivariate meteorological data) as long as these complex patterns are consistent throughout the AOI (Chen et al. 2014; Audebert et al. 2019). For example, convolutional neural networks (CNNs) have been shown to produce more accurate predictions than models that rely only on one-dimensional (spectral) variability, such as generalized linear models, random forest regression, and artificial neural networks (Saha et al. 2022, 2021). For datasets where spatiotemporal collinearity is dynamic in time, CNNs that are locally calibrated, such as those that incorporate recurrent networks, show promising results (Guo et al. 2022; Zang et al. 2020; Chen et al. 2021).

b. Robust to dependent observations

Environmental processes change in space and in time, often because of different drivers. For example, average annual temperatures across a continent will vary from place to place as a function of latitude, topography, and global circulation (spatial), while hourly temperatures for a given location will change as a function of seasonality, weather, and the diurnal cycle (temporal) (Fig. 7). Since environmental processes change over space and time, and since space and time are both continuous fields, observations that are nearer to each other in space and time will be related (Anselin and Li 2020). Depending on what spatiotemporal driver you wish to characterize, your observations need to be sampled at some interval in space and/or time to capture important patterns, trends, and drivers of variability, without undersampling (sampling too far apart in space and/or time such that the information content of the signal is lost or changed) or oversampling (sampling too close together in space and/or time, which is akin to repeating a measurement) [section 3b(1)(i), Fig. 8]. In experimental analysis, defining the sampling interval (how much time will pass between subsequent samples being taken), sampling frequency (the inverse of sampling interval), and sample size (how many observations will be collected in total, the product of sampling interval and study duration) are important considerations for experimental design.

Fig. 7.
Fig. 7.

Schematic of the spatial and temporal scales (fprocess = 1/time or 1/length) of terrestrial water budget components modeled in hydrologic and hydrometeorological subroutines of coupled Earth systems models. Adapted from Cristiano et al. (2017).

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Fig. 8.
Fig. 8.

(top) Mismatched sampling frequency (fsample; points) and process frequency (line) can induce (bottom left) aliasing or (bottom right) autocorrelation in recreated signals. (center) The Nyquist frequency (fsample = 2 × fprocess) allows for recreation of a continuous signal from a discretized (sampled) signal with the fewest possible observations.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

1) Diagnosing and data engineering with observation dependence

(i) Define the process frequency

Most environmental systems will change over space and time due to different drivers, so the process frequency (fprocess) of interest in an analysis will have a temporal component and a spatial component. The temporal process frequency is the minimum frequency (roughly the difference between high and low values, such as the maximum rainfall intensity and zero rainfall intensity during a storm) of temporal variability of concern to the analysis (e.g., daily fprocess = 1 per day; or seasonal fprocess = 1 per 90 days). Similarly, the process spatial frequency is the inverse of the minimum linear distance of spatial variability of interest in the analysis. For example, if the goal of an analysis is to build a model that can interpolate convective precipitation between gauge stations, using Fig. 7 as a guide, the spatial fprocess is approximately 1/(100 m), and the temporal fprocess is approximately 1/(120 s). Defining the spatial and temporal fprocess is a critical part of observational experimental design, as variability in many environmental processes is associated with different physical forcings at different temporal/spatial scales. For example, a point-based precipitation time series integrates variability from mesoscale, synoptic-scale, and planetary-scale precipitation drivers, each representing a unique set of atmospheric forcings (Fig. 7).

(ii) Define the sampling frequency

The sampling frequency (fsample) is the spatial and temporal resolution of your dataset. For example, the spatial fsample of a satellite image is the inverse of the pixel size, and the temporal fsample is the inverse of the time between subsequent image captures. The magnitude of fsample relative to fprocess has stark implications for the datasets utility to accurately characterize an environmental process (de Knegt et al. 2010). Figure 8 shows a theoretical continuous environmental process as an f(t). We explore three different potential fsamples as points (top row). We then attempt to recreate our process signal by linearly interpolating between these sample points (bottom row). Undersampling (Fig. 8, left) changes the information content of the environmental process signal, a phenomenon commonly known as aliasing. Undersampling is associated with information loss, and we cannot resolve drivers of environmental process variability that act below our sampling frequency. Undersampling cannot be resolved computationally; it is a sampling problem which motivates the design of higher-resolution satellites and finer-scale coupled Earth system model simulations (Luvall et al 2017). We should consider any data that are sampled at a lower spatial or temporal frequency than the process frequency of interest to be aliased.

Figure 8 (middle column) shows a sampling frequency of exactly twice the signal frequency (fsample = 2 × fprocess), also known as the Nyquist frequency (Shannon 1949). The Nyquist frequency allows us to recreate a continuous signal from a discretized (sampled) signal with the fewest possible observations. Oversampling the signal (Fig. 8, right) retains the information content of the signal, but since we are now sampling too close together in time, our observations are no longer considered independent, which causes substantial problems in inference, prediction, and diagnostic modeling (Dubin 1998; Dormann et al. 2007).

In observational environmental analysis, just as we cannot specify a priori what we measure, we also typically cannot specify where and when the measurements are made. Global EOs discussed here are produced at standard spatial and temporal fsample, for example, a reanalysis dataset that contains hourly data on a 1-km grid, or a collection of satellite imagery with 7-day repeat coverage on a 10-m grid. In observational experimental design, decisions must be made about which spatiotemporal gridded EO datasets are most appropriate as inputs for an analysis. Likewise, we often need to combine observations at different sampling frequencies. In these cases, we will statistically upsample/upscale (use interpolation to model the dataset at higher temporal or spatial resolution) or downsample/downscale (use statistical aggregation to model the dataset at lower temporal or spatial resolution) one or more observations to represent the data as concomitant in space and/or time. When dealing with multiphase, multifrequency observational signals, we recommend use of the “engineer’s Nyquist” to estimate the ideal fsample, which is at least 2.5 times the process frequency (Srinivasan et al. 1998):
fsample>2.5×fprocess.
The engineer’s Nyquist provides some additional safeguard against aliasing environmental processes with heterogenous or nonstationary phase or frequency.

For example, using the above example analyzing convective precipitation with a spatial fprocess of 1/(100 m) and a temporal fprocess of 1/(120 s) (Fig. 7), we would to use precipitation data with a nominal spatial resolution of less than 40 m [fsample > 2.5 × 1/(100 m) or 1/(40 m)] and a nominal temporal resolution of less than 48 s [fsample > 2.5 × 1/(120 s) or 1/(48 s)]. The spatial scale of convective precipitation systems is often below the spatial sampling frequency of precipitation gauges and weather satellites. Because of this, gridded precipitation datasets, even those with spatial resolution below the size of convective storms, tend to have a negative bias in total precipitation estimates associated with the spatial aliasing of convective precipitation. In some places, like the midwestern United States, the majority of total precipitation is delivered by localized convective storms, leading to substantial negative bias in gridded precipitation estimates (Risser et al. 2019).

(iii) Do not use potentially aliased signals

In reality, an a priori definition of your spatial and temporal fprocess may not be a possibility. First, when dealing with complex, dynamic, unknown, or multiple process frequencies, it is unlikely that any single, heterogeneous engineer’s Nyquist exists that will uniformly protect against aliasing and inducing autocorrelation across the entire AOI and POI (Fig. 7) (Subba Rao and Terdik 2017). Second, because of historical limitations in environmental sampling, the spatiotemporal scales of variability for many environmental processes remain unknown. But even if you are unsure of your target fprocess, it is important to know the minimum fprocess that can be successfully evaluated from your data so that you do not incorrectly interpret aliased results. The minimum analyzable temporal fprocess for a dataset is minimum frequency, with a signal-to-noise ratio above 0 decibels (dB) of your experimental data time series (Subba Rao and Terdik 2017). The minimum analyzable spatial fprocess corresponds to the range of the spatial variogram of experimental data (Garrigues et al. 2006; Lark 2002; Bogaert and Russo 1999). A dataset cannot be used to assess the importance of, or make predictions using proxies of, physical drivers acting below the minimum analyzable fprocess.

(iv) Quantify autocorrelation

Spatiotemporal processes vary as a function of time [f(t)] and of two-dimensional space [f(x, y), where x and y may represent latitude and longitude]. Autocorrelation describes a situation where a variable is correlated with itself at some distance, either in time or in space, called a lag. At a lag of zero, we expect a signal to be perfectly correlated with itself. As we increase the lag, we expect the absolute value of the correlation to decrease. Positive autocorrelation occurs when subsequent observations have similar values, and negative autocorrelation occurs when subsequent observations have opposing values (Fig. 9).

Fig. 9.
Fig. 9.

(a) Time series showing positive, neutral, and negative autocorrelation, from left to right, respectively. (b) Spatial process showing positive autocorrelation or clustering, no autocorrelation, and negative autocorrelation (dispersal), from left to right, respectively. Adapted from Fortin and (Dale 2009).

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Fig. 10.
Fig. 10.

Checklist for robust data engineering and model development with (left) dependent features and (right) observations in spatiotemporal systems.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Autocorrelation can be a feature of the environmental process itself, or a feature of error in how the process has been observed or sampled. Examples of spatial processes that lead to autocorrelation include spatial diffusion (a process that spreads over space and time from an origin), spillover (a process that spreads across real or perceived boundaries, such as air pollution crossing national boundaries or diseases jumping between species), spatial interaction (a process characterized by movement in conscious response to certain characteristics of the environment, such as animal migration), and dispersal (a process that seeks distance from itself, like that observed in the population distribution of territorial mammals such as panthers). Examples of spatial error that lead to autocorrelation include oversampling (fprocessfsample), measurement error (e.g., antenna patterns in a radar image, a stream sensor that experiences calibration drift over time), and model misspecification (e.g., omitted variable bias or bias in a parameter estimate) (Dubin 1998; Dormann et al. 2007; McMillen 2003).

Like multicollinearity, autocorrelation in spatiotemporal data can cause substantial problems in statistical inference and must be quantified both in time and in space prior to modeling. Most methods for evaluating autocorrelation will calculate some metric of covariance (such as linear or rank correlation, covariance, or semicovariance) of a signal that has been offset from itself at increasing temporal or spatial distances, or lags. Our two primary goals in evaluating autocorrelation in exploratory analysis are to ascertain whether significant autocorrelation is present at a given fsample and at what lag this autocorrelation becomes insignificant (referred to as the range of the data). For the purpose of GRRIEn analysis, we recommend using Moran’s I statistic (Getis 2010) to calculate whether significant spatial autocorrelation is present globally in the data (Fig. 11 and supplemental appendix C), and the generation of spatial variograms (Garrigues et al. 2006; Oliver and Webster 1986) to estimate the range of this autocorrelation, or the distance at which autocorrelation becomes negligible (Fig. 12 and supplemental appendix C). For temporal autocorrelation, we recommend using an autocovariance function (ACF) to calculate both the significance and the range of temporal autocorrelation (Ma and Genton 2000; McLeod 1975). The diagnosis and management of autocorrelation in spatiotemporal data are the subject of extensive research, so these recommendations are necessarily simplistic. Excellent introductory information on the management of autocorrelation in data can be found in the literature (Dormann et al. 2007; Dubin 1998; Ramezan et al. 2019).

Fig. 11.
Fig. 11.

Analysis of spatial autocorrelation of model residuals for (a) OLS linear regression and (b) autocovariate regression predicting 2012–16 change in county-level voting patterns in U.S. presidential elections (McGovern et al. 2020). In (a) and (b), the top left panel shows the bootstrapped Moran’s I density plot and the top right shows a scatterplot generated by the splot package (Lumnitz et al. 2020) to visualize PySAL spatial analysis workflows (Rey and Anselin 2010). Figures in (a) show significant autocorrelation of model residuals, and figures in (b) show no significant autocorrelation of residuals. Including spatial autocorrelation as a covariate significantly decorrelated model residuals and is associated with an increase in the standard error on model coefficients [supplemental appendix C; code adapted from Wolf (2018)].

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Fig. 12.
Fig. 12.

Flowchart for interpretation of spatial autocorrelation of model residuals.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

2) Model training and validation with observation dependence

To understand how autocorrelation impacts statistical inference, we will start with the example of oversampling. Sampling too close together in time and/or space (fsample ≫ 3 × fprocess) is somewhat analogous to double counting ballots in an election. If we double count at random, it will inflate our sampling variance, and add instability to our results (Neville et al. 2004). If we double count systematically, it will bias our results (Jensen and Neville 2002). Whether we are double counting at random or using some sort of structure, we will think we have more votes than we do and will, therefore, have false confidence in our results (Ferraciolli et al. 2019). Like multicollinearity, autocorrelation is associated with model instability (Neville et al. 2004), but since it can artificially inflate confidence in predictor estimates (i.e., a negative bias in standard error of model parameters), it can make erroneous or biased estimators more difficult to identify during model validation (Fig. 12a and supplemental appendix C).

Because aliasing represents information loss, it is important to err on the side of oversampling your data. Failure to diagnose and model autocorrelation, however, has been associated with inflated estimates of model skill, bias in parameter estimates, and bias in feature selection in statistical, machine learning, and deep learning models (Kattenborn et al. 2022; Segurado et al. 2006; Sergeev et al. 2019; Brenning 2005; Ferraciolli et al. 2019), including CNNs (Kattenborn et al. 2022). Because of this, addressing autocorrelation is critical for models to generalize well in observational spatiotemporal systems.

The methods for addressing autocorrelation in supervised learning fall into two major categories: strategically downsampling/downscaling the input data (e.g., resampling to the range of the spatial variogram) prior to model training, and modeling autocorrelation during model training. The two main approaches to modeling autocorrelation include spatial lag models, in which autocorrelation is parameterized directly in the model, or spatial error models, which force autocorrelation to the structure of model residuals. Some autocorrelation-tolerant modeling methods include autocovariate regression, autoregressive models, spatial eigenvector mapping, generalized least squares regression (Dormann et al. 2007), and latent variable grouping (Neville and Jensen 2005; Carter et al. 2016). Spatial lag models are appropriate when modeling spatial processes, i.e., autocorrelation that is a generating feature of the spatial process serving as the predictand. Spatial error models are appropriate when autocorrelation is suspected to be an artifact of sampling, measurement, or model specification error. If autocovariance-associated spatial processes, measurement error (Fig. 12d), or dependence with spatiotemporal features (Fig. 12c) are inaccurately parameterized in the model, this will be apparent when looking at model residuals (section 3c: checking your work; Fig. 12).

c. Checking your work

In GRRIEn analysis, it is essential to example the residuals of your model (difference between predictions and observations of your predictand or error, Fig. 11) for spatial and temporal patterns. Spatial and temporal patterns in model residuals indicate that the model will not generalize well and offer clues as to why it is not robust (Dubin 1998; Dormann et al. 2007). Spatial or temporal autocorrelation of model residuals can be indicative of model instability induced by parameter and observation dependence or undiagnosed modes of intrinsic autocovariance in our environmental process. It can also indicate omitted variable bias, a condition where some important driver of spatiotemporal variability has not been parameterized in the model. Carefully examining the presence of spatial autocorrelation of model residuals (Moran’s I, supplemental appendix C; Fig. 11), the range of spatial autocorrelation of model residuals (spatial variogram), and the spatial distribution of model residuals (visual interpretation of spatial plots) can assist in diagnosing model misspecifications (Fig. 12).

If spatial autocorrelation of residuals is present globally, as diagnosed by Moran’s I, this needs to be clearly stated as a caveat to the results in the discussion, as it strongly indicates that estimates of model skill are inflated because of unparameterized autocorrelation in the data and increases the likelihood of biased estimators (Fig. 12a) or generalized model misspecification (Fig. 12d). When either global or local multicollinearity is present, it is also very important to plot the residuals of the trained model predictions over space and time. Spatiotemporal clustering in residuals is one sign that spatiotemporally variable multicollinearity has created model instability that has impacted the generalizability of your predictions. This is particularly true if your residuals cluster in regions that had locally divergent collinearity (e.g., if your residuals, or errors, are spatially clustered where local correlation coefficients, CN, VIF, and VD, are of a different sign or magnitude than the global correlation coefficients, CN, VIF, and VD) (Fig. 12b). Model residuals which are spatially concomitant with other unparameterized geographic covariates are indicative of omitted variable bias, which can be difficult to measure directly (Fig. 12c).

4. Reproducible

The term “reproducibility” is usually defined as obtaining the same results when others use the same datasets and methods of the original study. As outlined in the literature (National Academies of Sciences, Engineering, and Medicine 2018), reproducibility in research is often referred to as computational reproducibility: can another scientist understand your method sufficiently to replicate data processing, model building, and validation in their computational environment? The reproducible research community utilizes an ever-expanding collection of computational tools to facilitate easy sharing of code and data. We summarize the current best practices relating to geospatial analysis as standard elements of a GRRIEn repository (Fig. 13), which facilitates easy replication of data acquisition, engineering, model training, and model evaluation, all of which can be published on a platform like GitHub alongside peer-reviewed publication of your results. More information on the software and hardware tools required to create a GRRIEn repository can be found in supplemental appendix A [“Required (Computational) Tools”].

Fig. 13.
Fig. 13.

Standard elements of a reproducible repository for end-to-end data engineering, modeling, and interpretation, utilizing global EOs from public geospatial data repositories as predictors. Inputs (I) and outputs (O) of each code component are listed in README file and are either contained in the GitHub repository or, in the case of large datasets, sourced from the internet.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

a. GRRIEn repository elements

The GRRIEn repository contains standard elements used commonly in open geospatial research to ensure maximum portability and effortless reproducibility and adaptability of supervised learning workflows utilizing global Earth observations on public geodatabases. A GRRIEn repository should contain the following:

  • Raw data folder contains your in situ data, or labels, and standard delimiters of your AOI and POI (such as a start and end date, bounding box coordinates, or a shapefile) that can be used to collocate observations from global EOs.

  • Analysis-ready data folder contains the analysis-ready dataset merging your in situ data and global proxy variables; most often a data frame, matrix, or array.

  • Model outputs folder contains trained model objects that can be called to make predictions, surrogate models used for model explanation (see section 5: interpreted), as well as any data objects describing model fit and stability.

  • Figures and tables folder contains graphical and tabular representations of results, including graphs, figures, maps, and tables in open-source file formats.

  • Readme file (html or markdown file) is a document providing detailed instructions on how to use elements of the repository for end-to-end analysis, the hardware and software required, the process for submitting merge requests, and license.

  • Debian Linux–compatible software image (file) is a container image is a static file that includes executable code that will initiate an instance in the local compute environment containing all version-controlled software, system libraries, and system tools to execute all code in a repository.

  • Version history (.gitconfig file) is a detailed record of all edits and contributions made by individual team members to the repository over the life of the project.

  • Code folder is all scripts used to interact with the other files and elements in the repository.

b. GRRIEn code elements

The code contained in the GRRIEn repository should be sufficient to replicate a software and dataset version controlled end-to-end analysis. The basic steps of working through the GRRIEn code can be seen in Fig. 13 and include programmatic workflows to set up required software in a local computing environment, searching and accessing raw data from quality-controlled open data repositories, processing input data into the data structure required by the software algorithm, training and validating algorithms, explaining algorithms, and visualizing results. For portability (i.e., can be run on any computer) and traceability (i.e., each step of the research process is documented), the GRRIEn repository contains the code required to access data from the cloud, instead of storing data in the repository itself. As such, it builds on the movement to make all research data publicly available on quality-controlled databases with stable digital object identifiers (DOIs):

  • Setup.sh: bash script that initiates software image, installing all version-controlled software, libraries, and packages required by repository scripts, and creates a temporary data folder (temp data) in the user’s local environment. By providing required software, the setup.sh script ensures the portability of the workflow. Reads from “Debian–Linux compatible software image”; exports to user’s computer/compute platform.

  • Data source: utilizes online geodatabase API to programmatically access data for AOI and POI given database catalog search parameters. The data sourcing code promotes the use of published datasets in version-controlled repositories external to the GitHub repository. Reads from: “raw data” folder, internet; exports to “temp data.”

  • Process data code: collocates raw data (predictand) and downloaded global EO data (predictor) in space and time, including coordinate reference system conversions, resampling functions, and zonal statistics functions; converts collocated data into an analysis-ready format (i.e., stacked raster, numpy array, pandas dataframe, R DataFrame); preprocesses analysis-ready data for model training, including any feature reduction, variable transformations, and processing of missing data; calculates global and local multicollinearity and global and local autocorrelation of all variables. Reads from “raw data” folder, “temp data” folder; exports to: “temp data” folder, “analysis data” folder.

  • Model train and validate code: splits data into training and testing datasets; trains statistical model; derives statistics of model accuracy, stability, and model parameters; calculates Moran’s I, spatial variogram, and spatial plots of model residuals; and produces model explanations. Reads from “analysis data” folder; exports to “model data” folder.

  • Figure- and table-generating code: code required to generate figures and tables. Reads from “processed data” folder, “model data” folder; exports to “figures and tables” folder.

Having these elements in your repository will enable end-to-end replication and portability of the research workflow, ensuring that other users can implement it without having to invest time in compiling software, managing dependency chain issues, sourcing input data, or learning new computational techniques.

5. Interpreted

When trying to build models that are generalizable at landscape scales, it is not enough to rely purely on traditional metrics of model fit. One of the most awe-inspiring things about life on planet Earth is its propensity toward uniqueness. It is very difficult to collect representative samples of landscape-scale phenomena (Meyer and Pebesma 2022). Even in interpolation, patterns of interdependence between both parameters and observations in large-scale space/time systems are frequently both complex and dynamic in scale (Dormann et al. 2013). Because of global change, both natural and anthropogenic, forecasting and hindcasting in time will implicitly introduce new patterns in the scale dependence of and dependence between environmental variables (Refsgaard et al. 2014). Plus, in observational analysis, we can never capture every factor that drives variability in our environmental process, either by proxy or direct measurement, so omitted variable bias is functionally endemic. All this means that even after selecting the most important variables, training for model stability as well as model fit, ensuring that your data are not aliased, accounting for autocorrelation, and presenting your analytical pipeline in a format that can be vetted by your peers, your model may still produce erroneous interpolations and extrapolations in the wild.

To this end, the most scientifically relevant data emerging from your GRRIEn analysis pipeline are those about the trained model itself. The most critical tool you need to successfully implement GRRIEn analysis is your domain expertise as an Earth scientist. In environmental analysis, you cannot reasonably conclude that your model will make the right predictions in different contexts unless you make sure those predictions are happening for the right reason: do your model weights and parameters reflect a physically plausible diagnosis of the environmental system? In other words, has your published model been interpreted within the context of your theoretical understanding of the system it purports to represent? It is the process of expert interpretation of the machine learning algorithm that leads to knowledge discovery in environmental data science.

a. Interpreted modeling step 1: Form an interpretable hypothesis

Prior to modeling, use your domain expertise to write a series of hypothesis statements for every predictor variable used in model training, using very specific verbiage reflecting data and relationship type. This hypothesis should be motivated by examples from the experimental literature in your field. Things to consider in your hypothesis statement include the following:

  • What kind of data are you using? Are your variables continuous or categorical? If continuous, are the values bounded, and what are the data distributions? Have the data been log transformed, normalized, or standardized prior to analysis? If categorical, are categories nominal or ordinal? Are there any categories that are arbitrary? Do you have class imbalances, or any categories that are not representatively sampled in the training data?

  • What types of relationships do you expect to find? For continuous or ordinal data, is the relationship between the predictand and individual predictor linear or nonlinear? If nonlinear, is the relationship monotonic (predictand consistently either increases or decreases as predictor increases) or nonmonotonic [predictand increases with increase in predictor for some range(s) of predictor value and decreases for other range(s) of predictor value]? If nonmonotonic, is the relationship global (e.g., polynomial relationship) or local (e.g., changepoints or peak over threshold response)? In multivariate datasets, are there interactions between predictors (e.g., maize yield response to air temperature is positive when precipitation is high, negative when precipitation is low)? If interactions are suspected, are these interactions isotropic (monotonic across both variables) or anisotropic (nonmonotonic across either variable)? When evaluating interactions between continuous and categorical data, what moment(s) of the distribution of the continuous variable are likely to modify, or be modified by, the categorical variable (e.g., does storm type change the mean, variance, skew, or kurtosis of precipitation rate)?

b. Interpreted modeling step 2: Identify a model interpretation method

Model explanation methods fall into two main categories: instance (local) explanation methods and model (global) explanation methods (Fig. 14).

Fig. 14.
Fig. 14.

The model explanation pipeline for black-box and white-box algorithms. For white-box models, global explanations can be extracted from the model itself. For black-box algorithms, global explanations must be inferred by way of a surrogate model, trained from both input data and predictions. Local explanations are derived from interpreting predictions across all possible values of predictors. Figure adapted from Burkart and Huber (2021).

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

Instance (local) explanation methods allow the user to understand how the predictand responds to the predictor by producing local (to values of predictor) predictions for specific ranges of input values [e.g., like one example (or sample) from the training dataset]. Local explanation methods are useful for exploring and qualitatively characterizing nonlinear (including nonmonotonic and anisotropic) relationships between input and output variables. Local feature importance metrics are often interpreted both quantitatively (e.g., comparing the magnitude of variance in the predictand associated with individual predictor variables to determine the relative importance of a variable among multivariate predictors) and visually (e.g., a graph of predictions along the range of possible values of a predictor). Methods for instance explanation include locally interpretable model-agnostic explanations (LIMEs), Shapely values, and local sensitivity analysis (Ryo et al. 2021).

Model (global) explanation methods quantify or summarize the importance of a predictor across all possible values of predictands. Generally, this is accomplished by providing an easy-to-understand function that can generalize how a predictand will respond to all values of predictors. For the purpose of generating global explanations, we can divide supervised algorithms into two main classes: white-box models and black-box models (Fig. 14). With white-box models, which include linear models (multivariate regression, logistic regression), decision trees, rule-based models, interactive models, and Bayesian networks (Burkart and Huber 2021), global feature importance can be deciphered from the model object itself. For example, the sign and magnitude of regression coefficients yield insight into the nature and strength of the relationship between predictor and predictand. Black-box models, such as neural networks, map relationships between predictors and predictands through a series of weights applied to different transformations of data, and as such, they produce model objects that are difficult for humans to quantitatively interpret. For black-box models, we must generate surrogate models that approximate the trained model’s function, often using coefficients associated with linear or nonlinear (i.e., polynomial, interactive terms) representations of input variables, to arrive at a human-interpretable global explanation. Methods for surrogate model imputation include the sum of Shapely values (Lundberg and Lee 2017; Aas et al. 2021), decision paths (Van Assche and Blockeel 2007; Sagi and Rokach 2020), and counterfactual explanations (Verma et al. 2020).

Both local and global feature importance offer utility in helping scientists interpret black-box machine learning algorithms for knowledge discovery. Local feature importance metrics can assist with providing a physical explanation for feature selection, comparing the relative importance of predictors, confirming predictand values associated with peak-over-threshold responses, and confirming qualitative theories of variability in systems. Global model explanations are useful if your goal is to learn a complex, unknown function describing your environmental system (such as to parameterize a process model). They are strongly encouraged any time you will be using your machine learning method to extrapolate beyond the spatial/temporal domain of your training dataset or interpolate to a different spatial/temporal resolution that may be associated with unique physical drivers.

c. Interpreted modeling step 3: Use local and global explanation methods to confirm or reject original hypotheses

In model interpretation, we use local and global feature importance metrics to evaluate our original hypothesis. Do local and global feature importance metrics confirm or reject your original data hypothesis, as it related to your data types and variable types? If not, does this represent a plausible new discovery of spatiotemporal variability in your system? If a model lacks a physical explanation, the model is not robust.

6. Conclusions: Using GRRIEn for experimental design

Most students are trained to write a research manuscript containing five core sections: introduction, methods, results, discussion, conclusion. This standard format has evolved alongside experimental science. Much like our GRRIEn repository structure (section 4) ensures reproducibility of computational research, the format of a standard research manuscript is a well-worn, highly effective roadmap for reproducibility in experimental research. Each section corresponds to a stage of the analysis, and the guidelines for what must be included in these sections mirror details which must be attended to for a rigorous experimental analysis to occur. Figure 15 translates this foundational manuscript structure for a Generalizable, Reproducible, Robust, and Interpreted Environmental (GRRIEn) analysis.

Fig. 15.
Fig. 15.

Structuring the research manuscript for GRRIEn analysis.

Citation: Artificial Intelligence for the Earth Systems 2, 2; 10.1175/AIES-D-22-0065.1

The core sections are as follows:

  • Introduction: The introduction theoretically motivates the analysis, presents research objectives (interpolation, extrapolation, or diagnostic modeling), and decomposes this objective into a set of physically motivated hypotheses relating the predictand (environmental process observations) to predictors (globally available Earth observations).

  • Methods: The methods section introduces the predictand and predictors; defines the area of interest and period of interest; defines and theoretically motivates process and sampling frequencies; demonstrates that training data reflect environmental process variability across the AOI and POI; describes how predictand and predictors were coassociated into an analysis-ready dataset, including any spatial/temporal resampling and coordinate reference system conversions; describes how subsetting of testing and training data is representative of any scale-dependent feature or observation dependence in the data; and motivates model selection, such as regularization, local calibration, or autocorrelation functions, using any feature and observation dependence in your data.

  • Results: The results section presents results of interpolation or extrapolation, quantifies model fit and stability, decomposes model error into bias and variance, provides spatial analysis of model residuals, and presents local/global model interpretation.

  • Discussion: Using original hypotheses as guides, the discussion section contextualizes model interpretation in theoretical background discussed in introduction; describes caveats in results of interpolation or extrapolation related to autocorrelation, multicollinearity, omitted variable bias, or nonsensical model diagnostics; and discusses the limitations of the current suite of global EOs in resolving research objectives that relate to sensor design, or data spatial or temporal resolution.

  • Conclusions: The conclusions section summarizes consistent findings and knowledge discovery related to the environmental process.

Global Earth observations provide an opportunity to analyze spatiotemporal variability in environmental systems at unprecedented levels of detail, but to unlock this information, Earth scientists must work with a suite of computational tools that are evolving alongside the data and require technical training. For Earth science disciplines to maintain a legacy of rigor, access to these computational tools for research must not interfere with focused training in, or application of, skills in theoretical and applied sciences in a researcher’s area of expertise. The goal of GRRIEn analysis is to anchor the role of the Earth scientists embarking in uncharted territories of computational research in the traditions of rigor, consistency, transparency, and theoretical history that have long formed the foundation for scientific understanding, by simplifying expectations for open spatiotemporal data science. The GRRIEn framework is not meant to be complete or static. The computer hardware, software, data, and algorithms that are used in environmental data science are all evolving at breakneck speed. Instead, by outlining universal components of computational experimental pipelines in the Earth sciences, including dataset engineering, model training, and model interpretation, the GRRIEn framework seeks to set a baseline for community standards that formally integrates the methods of the scientific process within computational research. Earth scientists embarking on research in the digital age may find themselves in uncharted territory. The GRRIEn analysis framework is intended to serve as a packing list of critical computational tools and spatiotemporal statistics to enable traceable paths of true knowledge discovery on this journey.

Acknowledgments.

We want to acknowledge the anonymous reviewers of this manuscript, whose substantial contributions of time, expertise, and insight can be seen throughout the finished work. In addition, we acknowledge the World Climate Research Programme’s Working Group on Coupled Modelling, which is responsible for the Coupled Model Intercomparison Project (CMIP), and we thank the climate modeling groups (listed in Fig. 5 of this paper) for producing and making available their model output. For CMIP, the U.S. Department of Energy’s Program for Climate Model Diagnosis and Intercomparison provides coordinating support and led development of software infrastructure in partnership with the Global Organization for Earth System Science Portals.

Data availability statement.

Figure 12 utilizes U.S. elections data (McGovern et al. 2020), prepared by the Center for Spatial Data Science at the University of Chicago.

REFERENCES

  • Aas, K., M. Jullum, and A. Løland, 2021: Explaining individual predictions when features are dependent: More accurate approximations to Shapley values. Artif. Intell., 298, 103502, https://doi.org/10.1016/j.artint.2021.103502.

    • Search Google Scholar
    • Export Citation
  • Adnan, M. S. G., M. S. Rahman, N. Ahmed, B. Ahmed, M. F. Rabbi, and R. M. Rahman, 2020: Improving spatial agreement in machine learning-based landslide susceptibility mapping. Remote Sens., 12, 3347, https://doi.org/10.3390/rs12203347.

    • Search Google Scholar
    • Export Citation
  • Alin, A., 2010: Multicollinearity. Wiley Interdiscip. Rev.: Comput. Stat., 2, 370374, https://doi.org/10.1002/wics.84.

  • Anselin, L., and X. Li, 2020: Tobler’s law in a multivariate world. Geogr. Anal., 52, 494510, https://doi.org/10.1111/gean.12237.

  • Audebert, N., B. Le Saux, and S. Lefevre, 2019: Deep learning for classification of hyperspectral data: A comparative review. IEEE Geosci. Remote Sens. Mag., 7, 159173, https://doi.org/10.1109/MGRS.2019.2912563.

    • Search Google Scholar
    • Export Citation
  • Balsamo, G., and Coauthors, 2018: Satellite and in situ observations for advancing global Earth surface modelling: A review. Remote Sens., 10, 2038, https://doi.org/10.3390/rs10122038.

    • Search Google Scholar
    • Export Citation
  • Bárcena, M. J., P. Menéndez, M. B. Palacios, and F. Tusell, 2014: Alleviating the effect of collinearity in geographically weighted regression. J. Geogr. Syst., 16, 441466, https://doi.org/10.1007/s10109-014-0199-6.

    • Search Google Scholar
    • Export Citation
  • Benesty, J., J. Chen, Y. Huang, and I. Cohen, 2009: Pearson correlation coefficient. Noise Reduction in Speech Processing, J. Benesty et al., Eds., Springer Topics in Signal Processing, Vol. 2, Springer, 1–4.

  • Berrar, D., 2019: Cross-validation. Encyclopedia of Bioinformatcis and Computational Biology, C. Schonbach, K. Nakai, and S. Ranganathan, Eds., Academic Press, 542–545.

  • Blei, D. M., and P. Smyth, 2017: Science and data science. Proc. Natl. Acad. Sci. USA, 114, 86898692, https://doi.org/10.1073/pnas.1702076114.

    • Search Google Scholar
    • Export Citation
  • Bogaert, P., and D. Russo, 1999: Optimal spatial sampling design for the estimation of the variogram based on a least squares approach. Water Resour. Res., 35, 12751289, https://doi.org/10.1029/1998WR900078.

    • Search Google Scholar
    • Export Citation
  • Bond, C. E., A. D. Gibbs, Z. K. Shipton, and S. Jones, 2007: What do you think this is? “Conceptual uncertainty” in geoscience interpretation. GSA Today, 17, 4, https://doi.org/10.1130/GSAT01711A.1.

    • Search Google Scholar
    • Export Citation
  • Brauner, N., and M. Shacham, 1998: Role of range and precision of the independent variable in regression of data. AIChE J., 44, 603611, https://doi.org/10.1002/aic.690440311.

    • Search Google Scholar
    • Export Citation
  • Breiman, L., 1996a: Bagging predictors. Mach. Learn., 24, 123140, https://doi.org/10.1007/BF00058655.

  • Breiman, L., 1996b: Stacked regressions. Mach. Learn., 24, 4964, https://doi.org/10.1007/BF00117832.

  • Breiman, L., 2001: Random forests. Mach. Learn., 45, 532, https://doi.org/10.1023/A:1010933404324.

  • Brenning, A., 2005: Spatial prediction models for landslide hazards: Review, comparison and evaluation. Nat. Hazards Earth Syst. Sci., 5, 853862, https://doi.org/10.5194/nhess-5-853-2005.

    • Search Google Scholar
    • Export Citation
  • Burkart, N., and M. F. Huber, 2021: A survey on the explainability of supervised machine learning. J. Artif. Intell. Res., 70, 245317, https://doi.org/10.1613/jair.1.12228.

    • Search Google Scholar
    • Export Citation
  • Campbell, C. A., W. L. Pelton, and K. F. Nielsen, 1969: Influence of solar radiation and soil moisture on growth and yield of Chinook wheat. Can. J. Plant Sci., 49, 685699, https://doi.org/10.4141/cjps69-120.

    • Search Google Scholar
    • Export Citation
  • Carter, E., D. A. Herrera, and S. Steinschneider, 2021: Feature engineering for subseasonal-to-seasonal warm-season precipitation forecasts in the Midwestern United States: Toward a unifying hypothesis of anomalous warm-season hydroclimatic circulation. J. Climate, 34, 82918318, https://doi.org/10.1175/JCLI-D-20-0264.1.

    • Search Google Scholar
    • Export Citation
  • Carter, E. K., J. Melkonian, S. J. Riha, and S. B. Shaw, 2016: Separating heat stress from moisture stress: Analyzing yield response to high temperature in irrigated maize. Environ. Res. Lett., 11, 094012, https://doi.org/10.1088/1748-9326/11/9/094012.

    • Search Google Scholar
    • Export Citation
  • Carter, E. K., J. Melkonian, S. Steinschneider, and S. J. Riha, 2018a: Rainfed maize yield response to management and climate covariability at large spatial scales. Agric. For. Meteor., 256257, 242252, https://doi.org/10.1016/j.agrformet.2018.02.029.

    • Search Google Scholar
    • Export Citation
  • Carter, E. K., S. J. Riha, J. Melkonian, and S. Steinschneider, 2018b: Yield response to climate, management, and genotype: A large-scale observational analysis to identify climate-adaptive crop management practices in high-input maize systems. Environ. Res. Lett., 13, 114006, https://doi.org/10.1088/1748-9326/aae7a8.

    • Search Google Scholar
    • Export Citation
  • Challinor, A. J., J. Watson, D. B. Lobell, S. M. Howden, D. R. Smith, and N. Chhetri, 2014: A meta-analysis of crop yield under climate change and adaptation. Nat. Climate Change, 4, 287291, https://doi.org/10.1038/nclimate2153.

    • Search Google Scholar
    • Export Citation
  • Chan, J. Y.-L., S. M. H. Leow, K. T. Bea, W. K. Cheng, S. W. Phoong, Z.-W. Hong, and Y.-L. Chen, 2022: Mitigating the multicollinearity problem and its machine learning approach: A review. Mathematics, 10, 1283, https://doi.org/10.3390/math10081283.

    • Search Google Scholar
    • Export Citation
  • Chen, Y., Z. Lin, X. Zhao, G. Wang, and Y. Gu, 2014: Deep learning-based classification of hyperspectral data. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens., 7, 20942107, https://doi.org/10.1109/JSTARS.2014.2329330.

    • Search Google Scholar
    • Export Citation
  • Chen, Y., Y. Wang, Z. Dong, J. Su, Z. Han, D. Zhou, Y. Zhao, and Y. Bao, 2021: 2-D regional short-term wind speed forecast based on CNN-LSTM deep learning model. Energy Convers. Manage., 244, 114451, https://doi.org/10.1016/j.enconman.2021.114451.

    • Search Google Scholar
    • Export Citation
  • Clark, J. S., and A. E. Gelfand, 2006: A future for models and data in environmental science. Trends Ecol. Evol., 21, 375380, https://doi.org/10.1016/j.tree.2006.03.016.

    • Search Google Scholar
    • Export Citation
  • Committee on Earth Observing Satellites, 2022: Instruments table. CEOS database, accessed 23 May 2023, http://database.eohandbook.com/index.aspx.

  • Cristiano, E., M.-C. ten Veldhuis, and N. van de Giesen, 2017: Spatial and temporal variability of rainfall and their effects on hydrological response in urban areas—A review. Hydrol. Earth Syst. Sci., 21, 38593878, https://doi.org/10.5194/hess-21-3859-2017.

    • Search Google Scholar
    • Export Citation
  • de Knegt, H. J., and Coauthors, 2010: Spatial autocorrelation and the scaling of species–environment relationships. Ecology, 91, 24552465, https://doi.org/10.1890/09-1359.1.

    • Search Google Scholar
    • Export Citation
  • Dormann, C. F., and Coauthors, 2007: Methods to account for spatial autocorrelation in the analysis of species distributional data: A review. Ecography, 30, 609628, https://doi.org/10.1111/j.2007.0906-7590.05171.x.

    • Search Google Scholar
    • Export Citation
  • Dormann, C. F., and Coauthors, 2013: Collinearity: A review of methods to deal with it and a simulation study evaluating their performance. Ecography, 36, 2746, https://doi.org/10.1111/j.1600-0587.2012.07348.x.

    • Search Google Scholar
    • Export Citation
  • Dubin, R. A., 1998: Spatial autocorrelation: A primer. J. Hous. Econ., 7, 304327, https://doi.org/10.1006/jhec.1998.0236.

  • Feng, X., D. S. Park, Y. Liang, R. Pandey, and M. Papeş, 2019: Collinearity in ecological niche modeling: Confusions and challenges. Ecol. Evol., 9, 10 36510 376, https://doi.org/10.1002/ece3.5555.

    • Search Google Scholar
    • Export Citation
  • Ferraciolli, M. A., F. F. Bocca, and L. H. A. Rodrigues, 2019: Neglecting spatial autocorrelation causes underestimation of the error of sugarcane yield models. Comput. Electron. Agric., 161, 233240, https://doi.org/10.1016/j.compag.2018.09.003.

    • Search Google Scholar
    • Export Citation
  • Fortin, J. S., and M. R. T. Dale, 2009: Spatial autocorrelation. The SAGE Handbook of Spatial Analysis, A. S. Fotheringham and P. A Rogerson, Eds., SAGE Publications, 89–103.

  • Freund, Y., and R. E. Schapire, 1999: A short introduction to boosting. J. Japan Soc. Artif. Intell., 14, 771780.

  • Friedman, J. H., 2001: Greedy function approximation: A gradient boosting machine. Ann. Stat., 29, 11891232, https://doi.org/10.1214/aos/1013203451.

    • Search Google Scholar
    • Export Citation
  • Garrigues, S., D. Allard, F. Baret, and M. Weiss, 2006: Quantifying spatial heterogeneity at the landscape scale using variogram models. Remote Sens. Environ., 103, 8196, https://doi.org/10.1016/j.rse.2006.03.013.

    • Search Google Scholar
    • Export Citation
  • Getis, A., 2010: Spatial autocorrelation. Handbook of Applied Spatial Analysis: Software Tools, Methods and Applications, M. M. Fischer and A. Getis, Eds., Springer, 255–278.

  • Goldman, A. E., S. R. Emani, L. C. Pérez-Angel, J. A. Rodríguez-Ramos, J. C. Stegen, and P. Fox, 2021: Special collection on open collaboration across geosciences. Eos, 102, https://doi.org/10.1029/2021EO153180.

    • Search Google Scholar
    • Export Citation
  • Graham, M. H., 2003: Confronting multicollinearity in ecological multiple regression. Ecology, 84, 28092815, https://doi.org/10.1890/02-3114.

    • Search Google Scholar
    • Export Citation
  • Guo, J., Q. Xiong, J. Chen, E. Miao, C. Wu, Q. Zhu, Z. Yang, and J. Chen, 2022: Study of static thermal deformation modeling based on a hybrid CNN-LSTM model with spatiotemporal correlation. Int. J. Adv. Manuf. Technol., 119, 26012613, https://doi.org/10.1007/s00170-021-08462-9.

    • Search Google Scholar
    • Export Citation
  • Hembram, T. K., S. Saha, B. Pradhan, K. N. Abdul Maulud, and A. M. Alamri, 2021: Robustness analysis of machine learning classifiers in predicting spatial gully erosion susceptibility with altered training samples. Geomatics Nat. Hazards Risk, 12, 794828, https://doi.org/10.1080/19475705.2021.1890644.

    • Search Google Scholar
    • Export Citation
  • Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 19992049, https://doi.org/10.1002/qj.3803.

    • Search Google Scholar
    • Export Citation
  • Hilborn, R., and M. Mangel, 2013: The Ecological Detective. Princeton University Press, 336 pp.

  • Hoerl, A. E., and R. W. Kennard, 1970: Ridge regression: Biased estimation for nonorthogonal problems. Technometrics, 12, 5567, https://doi.org/10.1080/00401706.1970.10488634.

    • Search Google Scholar
    • Export Citation
  • Jensen, D., and J. Neville, 2002: Linkage and autocorrelation cause feature selection bias in relational learning. Proc. 19th Int. Conf. Machine Learning (ICML2002), Sydney, Australia, ICML, 259–266, https://groups.cs.umass.edu/jensen/wp-content/uploads/sites/17/2022/03//jensen-neville-icml2002.pdf.

  • Kalogirou, S., 2013: Testing geographically weighted multicollinearity diagnostics. GISRUK 2013, Liverpool, United Kingdom, GISRUK, 1–10, https://www.geos.ed.ac.uk/∼gisteac/proceedingsonline/GISRUK2013/gisruk2013_submission_2.pdf.

  • Kattenborn, T., F. Schiefer, J. Frey, H. Feilhauer, M. D. Mahecha, and C. F. Dormann, 2022: Spatially autocorrelated training and validation samples inflate performance assessment of convolutional neural networks. ISPRS Open J. Photogramm. Remote Sens., 5, 100018, https://doi.org/10.1016/j.ophoto.2022.100018.

    • Search Google Scholar
    • Export Citation
  • Kidd, C., A. Becker, G. J. Huffman, C. L. Muller, P. Joe, G. Skofronick-Jackson, and D. B. Kirschbaum, 2017: So, how much of the Earth’s surface is covered by rain gauges? Bull. Amer. Meteor. Soc., 98, 6978, https://doi.org/10.1175/BAMS-D-14-00283.1.

    • Search Google Scholar
    • Export Citation
  • Kim, J. H., 2019: Multicollinearity and misleading statistical results. Korean J. Anesthesiol., 72, 558569, https://doi.org/10.4097/kja.19087.

    • Search Google Scholar
    • Export Citation
  • Lark, R. M., 2002: Optimized spatial sampling of soil for estimation of the variogram by maximum likelihood. Geoderma, 105, 4980, https://doi.org/10.1016/S0016-7061(01)00092-1.

    • Search Google Scholar
    • Export Citation
  • Li, K., and N. S. N. Lam, 2018: Geographically weighted elastic net: A variable-selection and modeling method under the spatially nonstationary condition. Ann. Amer. Assoc. Geogr., 108, 15821600, https://doi.org/10.1080/24694452.2018.1425129.

    • Search Google Scholar
    • Export Citation
  • Li, M., D. Niu, Z. Ji, X. Cui, and L. Sun, 2021: Forecast research on multidimensional influencing factors of global offshore wind power investment based on random forest and elastic net. Sustainability, 13, 12262, https://doi.org/10.3390/su132112262.

    • Search Google Scholar
    • Export Citation
  • Lu, B., P. Harris, M. Charlton, and C. Brunsdon, 2014: The GWmodel R package: Further topics for exploring spatial heterogeneity using geographically weighted models. Geo Spat. Inf. Sci., 17, 85101, https://doi.org/10.1080/10095020.2014.917453.

    • Search Google Scholar
    • Export Citation
  • Lumnitz, S., and Coauthors, 2020: Splot–Visual analytics for spatial statistics. J. Open Source Software, 5, 1882, https://doi.org/10.21105/joss.01882.

    • Search Google Scholar
    • Export Citation
  • Lundberg, S. M., and S. I. Lee, 2017: A unified approach to interpreting model predictions. Advances in Neural Information Processing Systems, I. Guyon et al., Eds., Vol. 30, Curran Associates, https://papers.nips.cc/paper_files/paper/2017/file/8a20a8621978632d76c43dfd28b67767-Paper.pdf.

  • Luvall, J. C., C. Lee, N. Stavros, and N. Glenn, 2017: Surface biology and geology designated observables. Surface Biology Geology Community Workshop, Washington, DC, NASA, https://ntrs.nasa.gov/api/citations/20190025933/downloads/20190025933.pdf.

  • Ma, Y., and M. G. Genton, 2000: Highly robust estimation of the autocovariance function. J. Time Ser. Anal., 21, 663684, https://doi.org/10.1111/1467-9892.00203.

    • Search Google Scholar
    • Export Citation
  • Mahadi, M., T. Ballal, M. Moinuddin, and U. M. Al-Saggaf, 2022: A recursive least-squares with a time-varying regularization parameter. Appl. Sci., 12, 2077, https://doi.org/10.3390/app12042077.

    • Search Google Scholar
    • Export Citation
  • McGovern, T., S. Larson, B. Morris, and M. Hodges, 2020: County-level presidential election results for 2008, 2012, and 2016. Zenodo, accessed 17 February 2023, https://doi.org/10.5281/zenodo.3975765.

  • McLeod, I., 1975: Derivation of the theoretical autocovariance function of autoregressive-moving average time series. J. Roy. Stat. Soc., 24C, 255256, https://doi.org/10.2307/2346573.

    • Search Google Scholar
    • Export Citation
  • McMillen, D. P., 2003: Spatial autocorrelation or model misspecification? Int. Reg. Sci. Rev., 26, 208217, https://doi.org/10.1177/0160017602250977.

    • Search Google Scholar
    • Export Citation
  • Meyer, H., and E. Pebesma, 2022: Machine learning-based global maps of ecological variables and the challenge of assessing them. Nat. Commun., 13, 2208, https://doi.org/10.1038/s41467-022-29838-9.

    • Search Google Scholar
    • Export Citation
  • Mudereri, B., T. Dube, E. M. Adel-Rahman, and S. Niassy, 2019: A comparative analysis of PlanetScope and Sentinel-2 space-borne sensors in mapping Striga weed using guided regularised random forest classification ensemble. Int. Arch. Photogramm. Remote Sens. Spatial Inf. Sci., 42, 701708, https://doi.org/10.5194/isprs-archives-xlii-2-w13-701-2019.

    • Search Google Scholar
    • Export Citation
  • Murakami, D., N. Tsutsumida, T. Yoshida, T. Nakaya, and B. Lu, 2021: Scalable GWR: A linear-time algorithm for large-scale geographically weighted regression with polynomial kernels. Ann. Amer. Assoc. Geogr., 111, 459480, https://doi.org/10.1080/24694452.2020.1774350.

    • Search Google Scholar
    • Export Citation
  • Murugan, P., and S. Durairaj, 2017: Regularization and optimization strategies in deep convolutional neural network. arXiv, 1712.04711v1, https://doi.org/10.48550/arXiv.1712.04711.

  • National Academies of Sciences, Engineering, and Medicine, 2018: Open Science by Design: Realizing a Vision for 21st Century Research. National Academies Press, 232 pp.

  • Nativi, S., P. Mazzetti, M. Santoro, F. Papeschi, M. Craglia, and O. Ochiai, 2015: Big data challenges in building the global Earth observation system of systems. Environ. Modell. Software, 68, 126, https://doi.org/10.1016/j.envsoft.2015.01.017.

    • Search Google Scholar
    • Export Citation
  • Neville, J., and D. Jensen, 2005: Leveraging relational autocorrelation with latent group models. Fifth IEEE Int. Conf. Data Mining (ICDM’05), Houston, TX, IEEE, 49–55, https://doi.org/10.1109/ICDM.2005.89.

  • Neville, J., O. Simsek, and D. Jensen, 2004: Autocorrelation and relational learning: Challenges and opportunities. DTIC Rep., 9 pp., https://apps.dtic.mil/sti/pdfs/ADA472226.pdf.

  • Nickerson, R. S., 1998: Confirmation bias: A ubiquitous phenomenon in many guises. Rev. Gen. Psychol., 2, 175220, https://doi.org/10.1037/1089-2680.2.2.175.

    • Search Google Scholar
    • Export Citation
  • Oliver, M. A., and R. Webster, 1986: Semi-variograms for modelling the spatial pattern of landform and soil properties. Earth Surf. Processes Landforms, 11, 491504, https://doi.org/10.1002/esp.3290110504.

    • Search Google Scholar
    • Export Citation
  • Ramezan, C. A., T. A. Warner, and A. E. Maxwell, 2019: Evaluation of sampling and cross-validation tuning strategies for regional-scale machine learning classification. R