Volume 58, Issue 5 e2021WR031023
Research Article
Open Access

Physics-Informed Machine Learning Method for Large-Scale Data Assimilation Problems

Yu-Hong Yeung

Yu-Hong Yeung

Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA, USA

Contribution: Methodology, Software, Validation, Formal analysis, ​Investigation, Data curation, Writing - original draft, Visualization

Search for more papers by this author
David A. Barajas-Solano

David A. Barajas-Solano

Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA, USA

Contribution: Methodology, Software, Writing - original draft, Supervision

Search for more papers by this author
Alexandre M. Tartakovsky

Corresponding Author

Alexandre M. Tartakovsky

Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA, USA

Department of Civil and Environmental Engineering, University of Illinois Urbana-Champaign, Urbana, IL, USA

Correspondence to:

A. M. Tartakovsky,

[email protected]

Contribution: Conceptualization, Methodology, Formal analysis, ​Investigation, Resources, Writing - original draft, Supervision, Project administration, Funding acquisition

Search for more papers by this author
First published: 22 April 2022
Citations: 4

Abstract

We develop a physics-informed machine learning approach for large-scale data assimilation and parameter estimation and apply it for estimating transmissivity and hydraulic head in the two-dimensional steady-state subsurface flow model of the Hanford Site given synthetic measurements of said variables. In our approach, we extend the physics-informed conditional Karhunen-Loéve expansion (PICKLE) method to modeling subsurface flow with unknown flux (Neumann) and varying head (time-dependent Dirichlet) boundary conditions. We demonstrate that the PICKLE method is comparable in accuracy with the standard maximum a posteriori (MAP) method, but is significantly faster than MAP for large-scale problems. Both methods use a mesh to discretize the computational domain. In MAP, the parameters and states are discretized on the mesh; therefore, the size of the MAP parameter estimation problem directly depends on the mesh size. In PICKLE, the mesh is used to evaluate the residuals of the governing equation, while the parameters and states are approximated by the truncated conditional Karhunen-Loéve expansions with the number of parameters controlled by the smoothness of the parameter and state fields, and not by the mesh size. For a considered example, we demonstrate that the computational cost of PICKLE increases near linearly (as N1.15) with the number of grid nodes N, while that of MAP increases much faster (as N3.28). We also show that once trained for one set of Dirichlet boundary conditions (i.e., one river stage), the PICKLE method provides accurate estimates of the hydraulic head for any value of the Dirichlet boundary conditions (i.e., for any river stage).

Key Points

  • The modified physics-informed machine learning PICKLE method for large-scale data assimilation is proposed

  • PICKLE is orders of magnitude faster than traditional a posteriori probability method for the considered high-resolution Hanford model

  • Trained for one set of boundary conditions, the PICKLE method can model data for different values of the boundary conditions

Plain Language Summary

We present a novel physics-informed machine learning framework for parameter and state estimation in large-scale natural systems's models. Using the Hanford Site as an example, we demonstrate that the proposed framework is comparable in accuracy with the standard parameter estimation methods but is significantly faster than these methods.

1 Introduction

Improving the predictive ability of numerical models has been the main goal of computational sciences. However, when applied to natural systems such as subsurface flow and transport, predictive modeling is complicated by the inherent uncertainty in the distribution of subsurface properties, including hydraulic conductivity, that enter the subsurface models as parameters. Uniquely estimating subsurface parameters from measurements of parameters and system states (e.g., hydraulic head) without numerical regularization is not possible because of the ill-posedness of the arising inverse problems. To further complicate the matter, the multiple length scales of heterogeneity and multiple time scales of flow and transport processes create enough ambiguity such that the same data can be described with different models, including deterministic and stochastic partial differential equation (PDE) models, non-local (integro-differential equation) models, and, most recently, machine learning and artificial intelligence-based models.

In selecting the right modeling approach, one can rely on Occam's razor or the law of parsimony principle that the simplest explanation (model) is usually the right one. However, given the above mentioned uncertainty in model parameters, the ability of models to be conditioned on spatially varying data is another critical criterion in selecting a computational model (Neuman & Tartakovsky, 2009). In theory, any model that involves observable parameters and states can be conditioned on the measurements of these variables if such are available. However, the computational cost of conditioning models on data may vary significantly.

Conditioning models on direct measurements of space-varying parameters (e.g., transmissivity) is relatively straightforward and can be achieved using Kriging or Gaussian process regression (GPR; Neuman, 1993; Tipireddy et al., 2020). In Kriging, an estimate of the conductivity field is obtained as a linear combination of the measured values and basis functions given in terms of a covariance kernel (Matheron, 1963; Rasmussen, 2003). Then, the flow problem (conditioned on the measurements of transmissivity) can be found by solving the Darcy flow equation with the transmissivity field given by the Kriging estimate. The advantage of using Kriging for conditioning flow models on transmissivity measurements is that it provides an empirical Bayesian prediction in terms of the conditional mean (the most likely distribution of transmissivity given its measurements) and the transmissivity covariance (a measure of uncertainty). The conditional mean and covariance of transmissivity can be used to obtain probabilistic estimates of the hydraulic head and fluxes conditioned on the transmissivity measurements. However, conditioning deterministic or stochastic flow predictions on the measurements of both the hydraulic head and transmissivity is more challenging because it requires solving an inverse problem that typically involves computing forward solutions of the Darcy equation multiple times for different realizations of the transmissivity field.

The inverse problem of computing the deterministic conductivity and hydraulic head fields given sparse measurements of these fields can be solved via maximum a posteriori (MAP) estimation, a Bayesian point estimation approach that consists of computing the largest mode of the posterior distribution of conductivity conditioned on the observations (D. A. Barajas-Solano et al., 2014; Kitanidis, 1996). As we later demonstrate in this paper, the computational cost of the MAP method rapidly increases with the number of unknown parameters (as the power of three of the number of unknown parameters). Several methods have been proposed to address this curse of dimensionality by reducing the number of parameters in the representation of the space-varying conductivity field. The simplest approach involves a coarse representation of the conductivity field that disregards small scale heterogeneity (e.g., variations of the hydraulic conductivity within geological layers). However, numerous studies (e.g., Shuai et al., 2019; A. M. Tartakovsky, 2010) have demonstrated that small-scale variations in conductivity might have a significant effect on large scale flow and transport processes. The pilot point method (PPM; Certes & de Marsily, 1991) addresses this issue by modeling the small-scale heterogeneity as a function of a small number of parameters (pilot points) that are estimated through the inverse procedure. Some of the challenges in PPM include the dependence of parameter estimates on the number and locations of the pilot points, the functional approximation of small-scale heterogeneity, and the regularization schemes (Doherty et al., 2010). A singular value decomposition (SVD) analysis of the sensitivities of observations with respect to pilot points can be leveraged to reduce the effective dimension of the inverse problem (see, e.g., Tonkin & Doherty, 2005). Other approaches for reducing the effective dimension of the inverse problem include the principal component geostatistical approach (PCGA; Kitanidis & Lee, 2014; Lee & Kitanidis, 2014), and methods based on data assimilation over latent spaces constructed via machine learning approaches, such as variational autoencoder methods and conditional generative adversarial networks (Kadeethum et al., 2021; O’Malley et al., 2019), among others.

Bayesian methods provide another approach for solving inverse problems. Compared to deterministic methods such as MAP, Bayesian methods aim to estimate the probabilistic distributions of the parameters and states conditioned on measurements (D. Barajas-Solano & Tartakovsky, 2019; Herckenrath et al., 2011; Li & Tartakovsky, 2020; Yoon et al., 2013), and by construction quantify the uncertainty in the inverse problem solution.

Physics-informed machine learning methods have recently emerged as promising tools for estimating parameters in differential equation models, including subsurface flow models. Many deterministic and probabilistic machine learning methods have been proposed where physics is enforced through optimization constraints or physics-model-generated data sets (Karpatne et al., 2017). Here, we focus on the physics-informed conditional Karhunen-Loéve expansion (PICKLE) method (A. Tartakovsky et al., 2020) where the state and parameter fields are represented using conditional Karhunen-Loéve expansions (CKLEs) (Tipireddy et al., 2020). The parameters in these expansions are found by minimizing the sum of square differences between the CKLE approximations and measurements of the fields, plus the sum of squared residuals of the governing equation. A similar approach is used in the physics-informed neural network (PINN) method (He et al., 2020; A. M. Tartakovsky et al., 2020) where deep neural networks are used to approximate parameter and state fields. The main difference between the physics-informed ML methods such as PICKLE and PINN and standard methods for solving inverse problems is that the ML methods only require computing derivatives with respect to time and space to evaluate the residuals and parameters as part of gradient-based minimization algorithms. There is no need to numerically solve the governing equation as in PPM, which is often the computational bottleneck of standard inverse methods. The key advantage of using CKLEs over neural networks for modeling the parameter field is that CKLEs enforce a spatial covariance structure on the modeled field and act as a geostatistics-based regularizer.

Like the PCGA method, PICKEL employs the eigendecomposition of the parameter covariance matrix. However, these methods differ in several key ways. In PCGA, the transmissivity field is modeled as a linear expansion with a polynomial basis plus a deviation component with a spatial distribution penalized via a weighted 2-norm, where the weight is the inverse of a covariance matrix that models the spatial variability of the deviation component; furthermore, the computational cost in PCGA is reduced by a matrix-free optimization algorithm that leverages the eigendecomposition of the covariance matrix. On the other hand, in PICKLE, the transmissivity field is fully modeled via a CKLE, and computational cost is controlled by the size of this CKLE.

In this work, we use the PICKLE method for obtaining deterministic estimates of transmissivity and hydraulic head fields conditioned on measurements of said fields. We apply this method for modeling steady-state two-dimensional groundwater flow at the Hanford Site given synthetic measurements of the transmissivity and hydraulic head. The synthetic measurements are generated using the hydraulic conductivity measurements and boundary conditions obtained in the Hanford Site calibration study of Cole et al. (2001). The PICKLE method was introduced in A. Tartakovsky et al. (2020) for parameter estimation in PDE models for a given set of deterministically known fixed boundary conditions. Here, we extend the PICKLE method for problems with uncertain flux boundary conditions and incrementally changing Dirichlet boundary conditions. Another significant contribution of this work is testing the PICKLE methods for problems with high-dimensional realistic parameter (transmissivity) fields (we find that more than 1000 terms in the KL expansion are needed to accurately approximate the log-transmissivity field obtained from the Hanford Site calibration study). We compare the performance of the PICKLE and MAP methods and show that the two methods have comparable accuracy, while the computational cost of MAP increases significantly faster with the problem size than that of PICKLE.

2 Groundwater Flow Model and Maximum a Posteriori Formulation of the Inverse Problem

We consider a two-dimensional model of groundwater flow. Our objective is to learn the spatial distribution of transmissivity urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0001 and hydraulic head urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0002 given the sparse measurements of T(x) and u(x), where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0003 is the simulation domain. We assume that the flow is governed by the boundary value problem (BVP)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0004(1)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0005(2)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0006(3)

where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0007 is the boundary of D and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0008 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0009 urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0010 are the portions of the boundary where the Neumann and Dirichlet boundary conditions are prescribed, respectively. In Equation 2, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0011 is the normal flux at the Neumann boundary urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0012, and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0013 is the unit vector normal to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0014. In Equation 3, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0015 is the prescribed hydraulic head on the Dirichlet boundary urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0016.

In this work, we assume that there are urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0017 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0018 observations of u and y = log  T, respectively, organized into the vectors us and ys. The locations of u and y observations are organized into the arrays Xu and Xy, respectively. In groundwater models, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0019 usually models the boundaries formed by rivers, in which case urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0020 is equal to the water level in these rivers, which is easy to measure. Therefore, we treat urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0021 as a known function. The homogeneous Neumann boundary condition urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0022 is imposed at the boundaries formed by the impermeable (e.g., basalt) layers, and we also treat this boundary condition as known. The non-homogeneous Neumann boundary conditions are commonly used to model groundwater inflow/outflow from/to rivers and lakes, groundwater discharge to the sea, and surface recharge. The measurements or estimates of groundwater inflow/outflow fluxes urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0023 are significantly less accurate than those of the Dirichlet boundary conditions uD(x; Jazayeri & Werner, 2019). Therefore, in this work, we consider two cases, one where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0024 is known, and another where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0025 is unknown and is estimated along with y(x) and u(x).

The transmissivity field T in Equation 1 is a function of x. Inverse problems for T(x) are typically solved in the context of numerical models where Equation 1 is discretized on a mesh and the values of T are estimated at a set of nodes associated with this mesh. In this work we employ a cell-centered finite volume (FV) discretization of Equations 1–3 using the two-point flux approximation (TPFA) of fluxes across cell faces. In this setting, the domain D is discretized into N FV cells, and the fields u and y are discretized into values located at the cell centers urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0026. The discrete field values are organized into the column vectors u and y, respectively, of length N. The coordinates of the cell centers are organized into the array Xc. The Neuman boundary condition function urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0027 is discretized into Nq values located at the centroids of the cell faces corresponding to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0028. These values are organized into the vector q. After discretizing the BVP using the TPFA-FV method, we obtain the set of algebraic equations
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0029(4)

with the stiffness matrix urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0030 and right-hand side urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0031 defined in Appendix A. In Equation 4, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0032 denotes the vector of discretized BVP residuals. The entries of the l vector correspond to the FV mass balance for each of the N cells. The set of cells can be split into three subsets: urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0033 (the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0034 cells adjacent to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0035), urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0036 (urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0037 cells adjacent to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0038), and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0039, where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0040 is of cardinality urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0041. Let urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0042, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0043 denote the vector of l entries corresponding to the indices in the subset urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0044. Only the mass balance for the cells in the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0045 set explicitly includes the q contributions; therefore, it follows that q enters into l only on the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0046 entries, and that urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0047 does not depend directly on q.

In the MAP method, the vectors of unknown parameters y and q are estimated by minimizing the 2-norm of the discrepancy between observations and model predictions, that is,
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0048(5)
where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0049 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0050 are observation matrices that downsample the vectors u and y into the vectors of u and y values at the locations where us and ys are measured. Specifically, u(Xu) = Huu and y(Xy) = Hyy. The inverse problem (Equation 5) is ill-posed; therefore, it is necessary to introduce regularization penalties to the cost function. Here, we choose to penalize the 2-norm of the discrete gradient of the y field, and the 2-norm of q, with the regularization penalty coefficient γ. Our choice of regularization for q is selected for convenience; nevertheless, the difference in inversion errors between the cases of known and unknown Neumann boundary conditions is small (see Sections 5.1 and 5.2), which indicates that this choice is adequate for the problem being considered. In the first regularization term, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0051 is the TPFA approximation of the gradient operator. The cost function in Equation 5 is, up to an additive constant, equal to the negative joint posterior log-likelihood of y and q in a Bayesian interpretation of the inverse problem (Stuart, 2010), with the data misfit terms corresponding to a Gaussian log-likelihood, and the regularization penalties terms to a Gaussian log-prior; therefore, the estimates of y and q fields obtained from Equation 5 are equivalent to the mode of the posterior distributions of these fields. Here, we use MAP to benchmark the PICKLE method. The details of our MAP implementation are given in Appendix B.

3 PICKLE Method for Inverse Problems

3.1 Method Formulation

The PICKLE method was proposed in A. Tartakovsky et al. (2020) for solving inverse diffusion equations with unknown diffusion coefficients. In PICKLE, the hydraulic head u(x) and the unknown parameter field y(x) are represented via CKLEs as
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0052(6)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0053(7)
where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0054 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0055 are the vectors of unknown CKLE coefficients, and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0056 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0057 denote the eigendecompositions of the conditional covariance kernels urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0058 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0059, respectively, which model the spatial variability of the u(x) and y(x) fields conditioned on the available observations of y. These eigendecompositions are calculated by solving the eigenvalue problems
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0060(8)

Here, the superscript c denotes that uc and yc are conditioned on the measurements of y. Methods for computing urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0061, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0062, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0063, and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0064 are described in Section 3.2.

The number of KL terms Nu and Ny are selected to satisfy the conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0065(9)
for certain tolerances rtolu and rtoly. For the eigenproblems solved via the eigendecomposition of the covariance matrices urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0066 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0067 evaluated on the cell-centered FV scheme with N cells, these conditions can be approximated as
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0068(10)

In this work, we set rtolu = rtoly = 10−6, unless mentioned otherwise.

Let urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0069 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0070 denote the CKLE approximations of u and y, given by Sections 6 and 7, respectively. In PICKLE, the inverse problem is solved by minimizing the 2-norm of the residuals urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0071. Together with a regularization penalty urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0072 on urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0073 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0074, as well as a data misfit for u predictions, the corresponding minimization problem reads
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0075(11)
Here, we consider two forms of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0076. The first form,
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0077(12)
was proposed in A. Tartakovsky et al. (2020) and shown to perform well when the reference y field is generated as a realization of a Gaussian field. We also consider a form of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0078 that penalizes the gradients of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0079 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0080 (computed with the approximate gradient operator D) as in the MAP method:
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0081(13)
The PICKLE formulation (11) is different from that of A. Tartakovsky et al. (2020) because it allows for unknown flux boundary conditions. Below, we propose a cost-effective treatment for this problem. As noted in Section 2, the FV mass balances for the subset urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0082 of cells (which includes the vast majority of cells) do not directly involve the vector of normal fluxes q. This allows us to exclude q from the inverse problem by penalizing the 2-norm of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0083, that is, to only minimize residuals within urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0084 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0085 cells:
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0086(14)
A PICKLE estimate of q is then obtained by solving
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0087(15)

for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0088, where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0089 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0090 are the solutions of the minimization problem (Equation 14). We note that urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0091 is linear in q, so that the solution of Equation 15 is straightforward. Because we use a CKLE for y conditioned on y measurements, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0092 satisfies y observations by construction. On the other hand, the CKLE for u is not conditioned on u measurements; therefore, we have included a data misfit term with coefficient β in Equation 14 penalizing the deviation between u predictions and observations.

We use the Trust Region Reflective algorithm (Branch et al., 1999) to solve the minimization problems of both the PICKLE and MAP methods. To do this, we cast Equations 5 and 14 as nonlinear least-squares problems. The least-squares minimization algorithm requires the evaluations of the Jacobian matrix J of the objective vector with respect to the parameters to be inverted, which is the most computationally demanding part of the least-squares minimization. Jacobian evaluation in the PICKLE method only requires computing the derivatives of the PDE residuals with respect to the CKLE coefficients and does not require solving the BVP (Equations1–3). On the other hand, evaluating the Jacobian in the MAP method requires solving multiple BVPs, either for computing a finite difference approximation or for evaluating the Jacobian via the chain rule as in Appendix B. Therefore, Jacobian evaluation is significantly cheaper for PICKLE than for MAP. In addition, the cost of Jacobian evaluation depends on the Jacobian matrix size, which is urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0093 in MAP and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0094 in PICKLE. Depending on the field smoothness and the required resolution, Nu and Ny can be chosen so that Nu + Ny ≪ N, leading to much smaller Jacobian matrices for PICKLE than those for MAP. The details of the optimization algorithm are given in Appendix C.

3.2 Computing Covariance Functions

To construct the conditional covariance urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0095, we first select a parameterized unconditional (prior) covariance kernel Cy(x, y; θ) for the y field. In this work we employ the 5/2-Matérn kernel
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0096(16)
with hyperparameters θ = {σ, l}, where σ and l are the standard deviation and the correlation length of y. These hyperparameters are estimated by minimizing the marginal log-likelihood function (Rasmussen, 2003) of the observations ys. Once the unconditional kernel hyperparameters have been estimated, the conditional mean and covariance of y are then computed from the Kriging equations
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0097(17)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0098(18)
where Cs is the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0099 observation covariance matrix with elements Cs,ij = Cy(xi, xj) and C(x) is the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0100-dimensional vector with the components Ci(x) = Cy(x, xi), where xi, xj ∈ Xy. It should be noted that the proposed framework is not limited to isotropic transmissivity fields or isotropic unconditional kernels such as Equation 16. Specifically, parameterized anisotropic covariance kernels can be employed instead of Equation 16 to model directional dependence of the transmissivity field if such dependence is observed in data (Deutsch & Journel, 1992).

The evaluation of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0101 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0102 from u measurements using parameterized covariance models and marginal likelihood maximization is not adequate for two related reasons: (1) the u(x) field is not stationary (i.e., the covariance kernel of u depends on x and y and not just the distance between x and y as in, e.g., the Matérn kernel), which limits our choice of covariance kernels; and (2) this purely data-driven approach does not enforce the governing equations and boundary conditions on the mean and covariance function even approximately. Therefore, in PICKLE we employ a Monte Carlo (MC) simulation-based method for computing the conditional mean and covariance of u.

In the MC method, we treat the partially known u(x) and y(x) as random variables urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0103 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0104, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0105 (where Ω is the corresponding random outcome space) conditioned on observed measurements of y. We model urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0106 using the stochastic truncated CKLEs
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0107(19)

where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0108 is the vectors of independent and identically distributed Gaussian random variables. The eigenpairs urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0109 satisfy the same eigenvalue problem Equation 8 as those in the deterministic CKLE of Section 7. We note that, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0110 does not need to be the same as Ny in Section 6, for example, it can be chosen with smaller rtoly to obtain a more accurate MC solution. However, in this study we set urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0111. Next, we construct an ensemble of Nens realizations of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0112, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0113 by sampling ξ(i) from urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0114 and evaluating the CKLE model (Equation19) with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0115.

In cases with unknown urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0116, we model urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0117 as a Gaussian random field with the mean and variance estimated from field measurements. Next, we generate Nens random realizations of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0118, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0119. For each member of the ensemble yc,(i) and the corresponding Neumann boundary condition urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0120, we calculate u(i) by solving the PDE problem of Equations 1–3. The resulting ensemble urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0121 is used to compute the mean values of u at each FV cell centroid,
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0122(20)
and the N × N covariance matrix of u with elements
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0123(21)

In this work, we set Nens large enough to assure that the PICKLE estimates of y do not change with further increase of Nens. In general, for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0124 to have at least Nu non-zero eigenvalues urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0125, the ensemble size should be urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0126. When it is not feasible to perform urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0127 MC simulations, shrinkage estimators can be employed to regularize the covariance matrix estimation (Chen et al., 2009). The accuracy of the covariance estimation with a small number of MC simulations can be increased by performing additional less-expensive coarser-resolution simulations using the Multilevel MC approach (Giles, 2015; X. Yang et al., 20212019). Also, there are several computationally efficient alternatives to MC methods, including the moment equation method (e.g., ; Jarman & Tartakovsky, 2013; Neuman, 1993; D. M. Tartakovsky et al., 2003) and polynomial-chaos-based approaches (Li & Tartakovsky, 2020; Lin & Tartakovsky, 2010; Tipireddy et al., 2020), surrogate models (X. Yang et al., 2018), and generative physics-informed machine learning methods (L. Yang et al., 2019).

4 Hanford Site Case Study

We compare the performance of the PICKLE and MAP methods for parameter estimation in the steady-state two-dimensional groundwater model of the Hanford Site. The Hanford Site is a former nuclear production complex on the Columbia River in the U.S. state of Washington and is currently operated by the United States Department of Energy. It is the most contaminated nuclear site in the US and is the focus of many modeling (e.g., Cole et al., 2001) and remediation (e.g., Burger et al., 2020) studies. In this comparison study, we use two reference transmissivity fields y(x) = ln T(x) and boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0128 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0129 that are based on the three-dimensional Hanford Site calibration study (Cole et al., 2001). This calibration study was performed on the unstructured quadrilateral grid shown in Figure 1a with 4–17 horizontal layers depending on the Cartesian plane coordinates and produced an estimate of the three-dimensional conductivity field. We obtain the first reference transmissivity field by depth averaging the conductivity field over the unstructured mesh.

Details are in the caption following the image

(a) Mesh of the Hanford Site subsurface flow model in the calibration study of Cole et al. (2001). Cells that represent the Columbia River are highlighted in blue with the river head boundary conditions prescribed at the vertices of these cells. (b) A quasi-uniform coarse mesh with the N = 1475 cells used in this study.

The original lateral mesh contains three different mesh resolutions and includes both the western and eastern banks of the Columbia River (the “Columbia River” cells are highlighted in blue in Figure 1a). We simplify the mesh by removing the river cells and prescribing Dirichlet BC on the western side of the Columbia River and by coarsening the mesh to achieve a uniform resolution, as shown in Figure 1b. For mesh coarsening, we merge groups of fine cells into a single coarse cell while ensuring that the resulting coarse mesh remains boundary-conforming and that the coarse cells are quadrilateral. The resulting mesh has 1475 cells. The transmissivity of each coarse cell is computed as the geometric average of the transmissivities of the fine cells. The transmissivity field corresponding to the coarse mesh is shown in Figure 2. We refer to this field as reference field 1 (“RF1”). We note that the PICKLE method can employ the FV (as in this study) or finite elements discretization to evaluate the residuals and, therefore, can utilize a multiresolution mesh.

Details are in the caption following the image

The coarse-resolution (N = 1475) RF1 redy = ln T field, well locations, and the parts of boundaries where different types of boundary conditions are prescribed. The transmissivity T has units of m2/day.

Figure 2 also shows the locations of some of the wells at the Hanford Site. We note that the calibration study (Cole et al., 2001) gives coordinates of 558 wells at the Hanford Site but that some of these wells are located in the same coarse or fine-cells. Because the FV discretization exclusively uses cell centers to denote spatial locations, multiple wells are considered as one measurement if they are located within the same coarse cell. As a result, there are 323 wells in the FV model shown in Figure 2.

We hypothesize that the accuracy of the PICKLE method depends on the smoothness of the reference transmissivity field. To test this hypothesis, in Section 5.2 we generate the reference field 2 (“RF2”) transmissivity field by using GPR (Equation 17) with 50 measurements drawn from the RF1 field at the locations randomly picked from the locations of the wells. By construction, the RF2 field is smoother than the RF1 field. In Section 5.2 we compare the performance of the PICKLE and MAP for the RF2 field. We also study the performance of PICKLE relative to MAP as a function of the FV model resolution. For this, we generate a higher-resolution mesh by splitting each cell in the mesh in Figure 1b (1× resolution) into four (4× resolution) equal-area cells, resulting in 5,900 cells. We note that there are 408 wells at this resolution.

The Dirichlet and Neumann boundaries urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0130 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0131 were defined in the Hanford Site calibration study and are shown in Figure 2. This calibration study also provides the estimates of the head urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0132 and fluxes urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0133 at these boundaries. In setting boundary conditions for our comparison study, we assume that urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0134 is known (in Sections 5.1 and 5.2, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0135 is given by the aforementioned calibration study, and in Section 5.4, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0136 is modified from the calibrated values to simulate the changing water levels in the Columbia and Yakima Rivers). In the case study with unknown urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0137, in the MC simulations we assume a normal distribution for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0138 with the mean and variance computed from the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0139 values in the calibration study. In the case study where we assume that urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0140 is known, we take the values of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0141 from the calibration study.

For each reference log-transmissivity field yref(x) = ln Tref(x), we generate the reference hydraulic head field uref(x) by solving the Darcy flow equation on the corresponding mesh with Dirichlet and (deterministic) Neumann boundary conditions set as described above. Then, we randomly pick urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0142 well locations and treat the values of yref at these locations as y measurements. We assume that u measurements are available at all well locations and the values of uref at these locations are taken as the u measurements. These synthetic data sets are used in the PICKLE and MAP methods to estimate the y(x) and u(x) fields.

All PICKLE and MAP simulations are performed using a 3.2 GHz 8-core Intel Xeon W CPU and 32 GB of 2666 MHz DDR4 RAM. The TPFA solver and the PICKLE and MAP methods are implemented in Python using the NumPy and SciPy packages. The weights in the PICKLE and MAP minimization problems are empirically found to minimize the error with respect to the reference y fields as β = 10, α = 10−4, and γ = 10−4. When a reference field is not known, these weights could be found using the standard cross-validation methods (Picard & Cook, 1984).

5 Numerical Experiments

5.1 RF1 Reference Field

First, we use PICKLE to estimate y on the coarse mesh using measurements of y and u generated for the RF1 reference field. We start with the unknown Neumann boundary condition case. The number of terms in the KL expansions of y and u are set to Ny = 1000 and Nu = 1000, respectively. The corresponding relative tolerances for these choices of Ny and Nu are urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0143 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0144, respectively. We assume that u measurements are available at all wells, that is, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0145. For the RF1 field, we find that the regularization in Equation 13 provides slightly more accurate results than the regularization in Equation 12. For example, for 10 different spatial distributions of 50 observations of y, the relative 2 errors in the estimated y field are in the ranges of 0.097–0.251 and 0.092–0.209 for the regularizers in Equations 12 and 13, respectively. The relative 2 errors are computed on the FV mesh as
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0146

Therefore, in all cases considered in this section we are using the regularization given by Equation 13.

Figure 3 shows the distribution of point errors in the PICKLE and MAP estimates of y relative to the RF1 y field obtained with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0147, 50, 100, and 200 y observations. For the considered measurement locations, the PICKLE and MAP methods have comparable accuracy for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0148, with MAP being more accurate for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0149.

Details are in the caption following the image

The coarse-resolution (N = 1475) RF1 reference y = ln T field and the PICKLE and MAP estimates of the y = ln T field and their point errors as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0150 given the unknown Neumann boundary conditions.

Because the inverse problem for y is ill-posed, the regularized PICKLE and MAP solutions depend not only on the number of measurements but also on the measurement locations. To study the effect of the measurement locations on the PICKLE and MAP estimation errors, for each value urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0151, we randomly generate 10 distributions of y measurement locations and estimate y for each of these locations distributions. Table 1a shows the ranges of relative 2 and absolute errors in the PICKLE and MAP y estimates as well as the number of iterations of the minimization algorithm and the execution time (in seconds) for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0152 ranging from 25 to 400. For comparison, we also show errors in y estimated via GPR (Equation 17). The error is defined as the maximum of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0153 (i = 1, …, N), where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0154 and yref(xi) are the values of the estimated and reference y fields at xi ∈ Xc.

Table 1. Performance of PICKLE and MAP in Estimating the Coarse-Resolution (N = 1475) y = ln T RF1 as Functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0155 With (a) Unknown and (b) Known Neumann Boundary Conditions
(a) Unknown Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0156
Solver 25 50 100 200 400
Least square iterations PICKLE 93–357 17–64 14–30 13–23 18–51
MAP 12–35 33–483 22–535 27–603 25–400
Execution time (s) PICKLE 180.43–495.34 159.94–592.47 142.71–300.60 144.73–253.47 143.26–378.35
MAP 402.08–1613.52 126.52–2134.91 70.40–2283.90 121.87–2737.72 116.74–1928.08
Relative 2 error GPR 0.175–0.244 0.147–0.180 0.119–0.156 0.095–0.114 0.069–0.083
PICKLE 0.130–0.285 0.100–0.152 0.083–0.119 0.076–0.090 0.056–0.069
MAP 0.095–0.130 0.085–0.100 0.076–0.576 0.067–0.296 0.052–0.062
Absolute error GPR 6.21–9.71 5.64–8.27 4.24–8.00 3.79–7.39 3.74–5.87
PICKLE 4.61–7.57 4.52–6.40 4.36–6.10 4.17–6.57 3.68–5.16
MAP 3.91–6.29 4.07–6.51 4.06–79.01 3.72–39.49 3.68–6.30
(b) Known Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0157
Solver 25 50 100 200 400
Least square iterations PICKLE 47–158 31–78 37–85 29–55 28–54
MAP 7662–15694 5343–13247 3497–19048 1488–4143 1081–2626
Execution time (s) PICKLE 155.06–510.83 120.67–305.00 137.37–324.25 119.43–216.47 119.43–216.47
MAP 1544.52–3133.03 1072.23–2582.03 726.96–3649.97 320.33–836.32 232.30–545.90
Relative 2 error PICKLE 0.104–0.347 0.092–0.209 0.081–0.109 0.071–0.083 0.057–0.064
MAP 0.092–0.107 0.083–0.100 0.074–0.085 0.064–0.071 0.050–0.069
Absolute error PICKLE 4.55–9.51 4.87–8.76 4.32–5.45 3.62–5.43 3.48–4.99
MAP 5.37–6.57 4.91–6.39 3.64–6.40 3.65–6.45 3.15–5.55

As expected, the accuracy of both methods increases with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0158. The PICKLE method is on average slightly less accurate than MAP in terms of both 2 and errors. However, MAP is more sensitive to the measurement locations. For example, for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0159 and 200, we observe that in MAP the maximum 2 errors are 0.576 and 0.296, respectively, versus 0.119 and 0.090 in PICKLE. We attribute the higher robustness of PICKLE relative to MAP with respect to measurement locations to the regularization effect of the CKLE representation of y. We also note that GPR has significantly larger errors than those in PICKLE and MAP for all considered examples.

Table 1a also shows that the computational cost of PICKLE is significantly smaller than the cost of MAP, and the cost difference increases with increasing urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0160. Note that we give the total execution time of PICKLE that includes the cost of MC evaluation of the mean and covariance of u (approximately 28 s), GPR (approximately 0.4 s), and eigendecomposition (approximately 2.9 s). The computational cost of GPR for the considered problem is negligible relative to both PICKLE and MAP, and we do not show it in this table. As with the estimation errors, we observe that the computational cost of PICKLE is significantly less sensitive to the measurement locations than that of MAP. For example, the ratio between the PICKLE maximum and minimum execution times in 10 realizations for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0161 and 400 are 2.74 and 2.64, respectively. In MAP, for the same values of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0162, these ratios are 4.01 and 16.51. The larger variability in the MAP computational time corresponds to the larger variability in the number of iterations in the MAP's least-square minimization algorithm.

Next, we investigate the performance of the PICKLE and MAP methods as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0163 when the Neumann boundary conditions are known. We note that the GPR approach for estimating y is based solely on y measurements and, therefore, is independent of the boundary conditions. Therefore, we do not present GPR errors in this comparison study. Table 1b shows the errors and execution times of the PICKLE and MAP methods for the same sets of y measurements as in the unknown Neumann boundary condition cases. We find that the errors of both methods only slightly decrease (less than 5%) relative to the unknown Neumann boundary condition cases. The execution time of PICKLE is practically not affected by whether the Neumann boundary conditions are known deterministically or stochastically, while the MAP execution time is increased.

We hypothesize that increasing the number of KL terms in the CKLE of y should increase the accuracy of PICKLE because it allows capturing more accurately the spatial correlation structure of y(x). However, increasing the number of KL terms also increases the number of unknown parameters and, therefore, the computational cost of PICKLE. In Tables 2a and 2b, we compare the errors and execution time of PICKLE with 1000 and 1400 terms in the y CKLE for the cases with unknown and known boundary conditions, respectively. We observe that, contrary to our hypothesis, increasing the number of KL terms does not lead to a significant increase in the accuracy of PICKLE. The 2 error decreases slightly, with only significant (10%) improvement for the smallest considered number urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0164 of y measurements. This is because Ny = 1000 already corresponds to a very small value of rtoly = 6.4 × 10−6. The further increase in Ny does not significantly improve the approximation power of the CKLE but does render solving the minimization problem costlier. We observe a slight increase in the errors because larger number of KL terms might require a stronger regularization (i.e., larger values of β). On the other hand, the increase in Ny leads to a significant increase in the execution time of PICKLE, by approximately a factor of 4 for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0165 and a factor of 2 for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0166, 200, and 323.

Table 2. Performance of PICKLE as a Function of the Number of KL Terms for Estimating the Coarse-Resolution y = ln T RF1 Field With (a) Unknown and (b) Known Neumann Boundary Conditions
(a) Unknown Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0167
KL terms 50 100 200 323
Least square iterations 1000 36 19 13 9
1400 67 15 10 9
Execution time (s) 1000 317.18 166.63 115.46 81.59
1400 1374.94 306.18 202.58 179.37
Relative 2 error 1000 0.138 0.104 0.092 0.076
1400 0.114 0.100 0.088 0.076
Absolute error 1000 6.49 5.07 5.36 5.20
1400 6.57 5.49 5.50 5.32
(b) Known Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0168
KL terms 50 100 200 323
Least square iterations 1000 23 18 13 9
1400 42 12 11 9
Execution time (s) 1000 199.28 169.18 116.97 80.88
1400 887.05 238.54 226.61 179.82
Relative 2 error 1000 0.133 0.102 0.090 0.074
1400 0.116 0.102 0.088 0.074
Absolute error 1000 6.43 5.14 5.38 5.37
1400 6.51 5.51 5.42 5.39

5.2 RF2 Reference Field

Here, we estimate y using the synthetic measurements of y and u generated on the coarse and fine meshes for the RF2 reference field. We assume that u measurements are available at all wells, that is, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0169 and 408 on the coarse and fine meshes, respectively. As in Section 5.1, the number of KL terms in the y and u expansions is set to Ny = Nu = 1000. The corresponding relative tolerances for these choices of Ny and Nu are rtolu = 3.01 × 10−9 and rtoly = 7.9 × 10−6, respectively. Opposite to our results for the RF1 field, here we find that the regularization of Equation 12 provides more accurate results than the regularization of Equation 13. For 10 different spatial distributions of 50 observations of y, the relative 2 errors in the estimated y field are in the ranges of 0.008–0.028 and 0.041–0.078 for regularizers given by Equations 12 and 13, respectively. Therefore, in this section we are using the regularization (12).

Figure 4 shows the RF2 reference y field and the point errors in the PICKLE and MAP estimates of the y field on the coarse mesh obtained using urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0170, 25, 50, and 100 for the unknown Neumann boundary condition case. The locations of y measurements are randomly selected from the well locations. Table 3 lists the ranges of 2 and errors in the y estimates as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0171 obtained with the PICKLE, GPR, and MAP methods. For each urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0172, 10 different random spatial distributions of measurement locations are selected to compute these ranges. Subtables (a) and (b) give results for unknown and known Neumann boundary conditions, respectively. PICKLE's 2 errors are smaller than those of MAP for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0173 and 100. For urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0174 and 25, the lower bounds of 2 errors are smaller for PICKLE and the upper bounds are smaller for MAP. The absolute errors follow the same pattern as the 2 errors.

Details are in the caption following the image

The coarse-resolution (N = 1475) RF2 reference y = ln T field and the PICKLE and MAP estimates of the y = ln T field and their point errors as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0175 given the unknown Neumann boundary conditions. The dots in the reference field are the locations of the 50 observations of y for constructing the reference field.

Table 3. Performance of PICKLE and MAP for Estimating the Coarse-Resolution (N = 1475) y = ln T RF2 as Functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0176 With (a) Unknown and (b) Known Neumann Boundary Conditions
(a) Unknown Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0177
Solvers 10 25 50 100
Least square iterations PICKLE 15–34 11–18 11–14 9–11
MAP 51–156 14–87 14–62 14–144
Execution time (s) PICKLE 63.5–110.2 85.3–138 81.5–105 83.4–114
MAP 66.2–202.4 28.4–191 29.5–138 38.1–306
Relative 2 error GPR 0.089–0.165 0.069–0.100 0.040–0.078 0.023–0.034
PICKLE 0.036–0.161 0.017–0.052 0.009–0.028 0.006–0.006
MAP 0.038–0.070 0.029–0.044 0.021–0.032 0.017–0.021
Absolute error GPR 3.34–4.45 2.05–4.29 2.11–4.36 1.04–2.37
PICKLE 1.16–3.61 0.831–1.51 0.792–1.20 0.781–0.854
MAP 1.18–1.53 1.08–1.44 0.858–1.49 0.790–1.09
(b) Known Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0178
Solvers 10 25 50 100
Least square iterations PICKLE 14–30 11–17 11–14 9–11
MAP 61–109 17–193 14–53 14–47
Execution time (s) PICKLE 58.2–107 53.3–71.8 53.5–68.2 72.8–94
MAP 66.6–126 72.1–122 66.4–121 39.5–326
Relative 2 error PICKLE 0.030–0.075 0.015–0.048 0.008–0.028 0.006–0.008
MAP 0.035–0.056 0.028–0.039 0.020–0.030 0.017–0.020
Absolute error PICKLE 1.14–2.16 0.834–1.46 0.794–1.28 0.734–0.985
MAP 1.02–1.48 1.02–1.45 0.873–1.47 0.800–1.08

For this coarse resolution, the execution time of PICKLE is larger than that of MAP. Both PICKLE and MAP perform well for unknown Neumann boundary conditions with estimation errors being only slightly larger than those in the case of known Neumann boundary conditions. The 2 and errors in estimating the RF2 field are significantly smaller than those in estimating the RF1 field, which is not surprising given the relative smoothness of the RF2 field. For the same reason, the execution times of both PICKLE and MAP methods are significantly smaller for modeling measurements from RF2 than RF1.

Next, we study the relative 2 error in the PICKLE solution for y(x) as a function of Ny and Nu, the number of terms in the CKLE of y and u, respectively, for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0179. For simplicity, we set Ny = Nu. Figure 5 shows that the 2 error decreases as Ny increases and, for the considered RF2 field, reaches the asymptotic value of less than 0.07 at Ny ≈ 800. Therefore, rtoly of the order of 10−6 (which is used in this work) and the corresponding Ny = 1000 are sufficient to obtain an accurate approximating of the RF2 y field and the corresponding reference u field. We note that for the (diffusion-type) Darcy equation, the solution u(x) is always smoother than the parameter field y(x). Therefore, the computational cost of PICKLE can be reduced by setting rtolu = rtoly, which for diffusion equations would result in Nu < Ny.

Details are in the caption following the image

Relative 2 errors versus the number of KL terms.

Finally, we test the relative performance of the PICKLE and MAP methods as a function of the resolution of the flow model by estimating y and u using the finer mesh with N = 5900. Table 4 lists the ranges of 2 and errors in the PICKLE, GPR, and MAP estimates of y as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0180, as well as execution times obtained from 10 different random distributions of measurements for each value of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0181. At this resolution, PICKLE is more accurate than MAP for most considered configurations and numbers of measurements. Tables 4a and 4b show results for unknown and known boundary conditions, respectively. Figure 6 shows the RF2 y field with the resolution N = 5900 and the point errors of the PICKLE and MAP estimates of this y field obtained with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0182, 25, 50, and 100, and unknown flux boundary conditions. It follows from Table 4 that the PICKLE 2 errors are smaller than those of MAP except the upper ranges of the errors for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0183 and 25. The lower bound of errors is lower for the PICKLE method (except for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0184) while the upper bound is larger (except for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0185). The errors for unknown Neumann boundary conditions are slightly larger in both methods than those in the case with known Neumann boundary conditions.

Table 4. Performance of PICKLE and MAP for Estimating the Fine-Resolution (N = 5900) y = ln T RF2 as Functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0186 With (a) Unknown and (b) Known Neumann Boundary Conditions
(a) Unknown Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0187
Solvers 10 25 50 100
Least square iterations PICKLE 14–35 11–16 11–15 9–11
MAP 196–288 102–234 78–199 69–205
Execution time (s) PICKLE 208–290 188–203 196–217 206–265
MAP 12,459–16944 6031–12374 3977–10121 4072–8186
Relative 2 error GPR 0.089–0.165 0.068–0.098 0.043–0.077 0.020–0.032
PICKLE 0.040–0.099 0.013–0.063 0.011–0.029 0.004–0.013
MAP 0.041–0.065 0.035–0.048 0.030–0.037 0.023–0.027
Absolute error GPR 2.95–4.40 2.06–4.19 2.18–3.80 0.80–2.10
PICKLE 1.70–2.99 0.715–2.32 0.476–2.04 0.348–0.834
MAP 1.33–1.51 1.18–1.55 0.961–1.48 0.877–1.30
(b) Known Neumann boundary conditions
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0188
Solvers 10 25 50 100
Least square iterations PICKLE 14–37 11–15 11–15 8–11
MAP 65–109 69–128 84–136 86–129
Execution time (s) PICKLE 193–200 181–200 192–217 192–223
MAP 6096–9333 6277–10348 4228–7124 4520–6330
Relative 2 error PICKLE 0.030–0.080 0.012–0.063 0.010–0.028 0.004–0.010
MAP 0.036–0.057 0.029–0.041 0.024–0.033 0.020–0.025
Absolute error PICKLE 1.65–2.80 0.709–2.29 0.476–1.93 0.302–0.823
MAP 1.33–1.60 1.06–1.69 0.900–1.62 0.922–1.37
Details are in the caption following the image

The fine-resolution (N = 5900) RF2 reference y = ln T field and the PICKLE and MAP estimates of the y = ln T field and their point errors as functions of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0189 given the unknown Neumann boundary conditions. The dots in the reference field are the locations of the 50 observations of y for constructing the reference field.

5.3 Scaling of the Execution Time With the Problem Size

Comparing Tables 3 and 4 it can be seen that the execution times of both PICKLE and MAP increase with increasing mesh resolution; however, the execution time of PICKLE increases slower than that of MAP. To further study the dependence of the computational cost of PICKLE and MAP on resolution, in Figure 7 we plot the execution times of these methods as functions of N for both the RF1 and RF2 reference fields. An additional mesh with N = 23,600 FV cells is generated by dividing each cell in the mesh with N = 5900 into four cells. The number of y measurements in all simulations is set to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0190. Figure 7 also shows that a power-law model fits the execution time for both methods. We note that for N = 23,600, the MAP method did not converge after running for two days. Therefore, the power law relationships for the MAP method are obtained based on the execution times for N = 1475 and 5900 and used to estimate MAP's execution times for the highest resolution.

Details are in the caption following the image

Execution times of PICKLE and MAP versus the number of FV cells for the y = ln T RF1 and RF2 reference fields. The execution times of MAP for the mesh with 23600 FV cells are estimated by extrapolation.

From Figure 7, we see that the PICKLE and MAP execution times increase as N1.1 and N3.2, respectively, for both the RF1 and RF2 fields. The close to linear dependence of PICKLE's execution time on the problem size gives it a computational advantage over the MAP method.

5.4 Modeling u and y Measurements Corresponding to Varying Boundary Conditions

In many natural systems such as the Hanford Site, boundary conditions can change with time. Once “trained” for one set of boundary conditions, PICKLE can be used without additional retraining to model data corresponding to different boundary conditions. Specifically, the covariance kernel Cu(x′, x″) that is calculated from MC simulations for certain boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0191 can be used to estimate y and u using measurements that correspond to boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0192 different than urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0193. In Appendix D we demonstrate that Cu(x′, x″) depends on the deterministic boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0194 only through the gradient of the mean hydraulic head field. We also demonstrate that an urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0195 change in Dirichlet boundary conditions leads to a urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0196 change in the covariance, where L is the size of the domain in the direction of mean flow. Here, we assume that this change in the covariance can be disregarded and we treat Cu(x′, x″) as independent of the Dirichlet boundary condition values.

As an example, we consider a case where the Dirichlet boundary condition urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0197 incrementally changes with time over the range (urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0198, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0199) in response to changes in the water level in the Columbia and Yakima Rivers. We denote urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0200 (i = 1, …, Nt) as the Dirichlet BC at each of Nt discrete times. At the ith time, the measurements urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0201 are collected at urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0202 spatial locations. The Neumann boundary conditions are assumed to be unknown. The measurements ys are available at urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0203 locations (ys do not change in time). To model the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0204 data, the covariance function Cu(x′, x″) can be found using the MC method described in Section 3.2 and setting urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0205, where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0206 is any one BC from urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0207. We emphasize that Cu(x′, x″) should be computed for only one member of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0208. This covariance is then employed to solve inverse problems for any member of urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0209.

To apply PICKLE to the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0210 measurements, in addition to Cu(x′, x″) we need to estimate urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0211 that, unlike the covariance, depends on urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0212. Section 3.2 describes the MC method for computing urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0213 that could be expensive to perform for each urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0214. Here, we propose to approximately compute urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0215 as the solution to the mean Darcy flow equation.
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0216(22)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0217(23)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0218(24)

where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0219, and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0220 is given by Equation 17. Equation 22 is a mean field equation, which disregards the term urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0221, where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0222 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0223.

To test this approach, we assume that the reference y field is given by the RF1 field from which we draw urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0224 measurements of y at random locations. Furthermore, we assume that u(x) is sampled at urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0225 locations, and that 3 measurements of u(x) gathered at 3 different times are available at each location forming three vectors of u measurements urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0226 with the locations urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0227 (i = 1, 2, 3). In general, the locations of u measurements can change over time, but in this work we assume that the locations of u measurements are the same for all considered boundary condition values. These boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0228 (i = 1, 2, 3) are constructed as follows: urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0229 is given by the calibration study (Cole et al., 2001), urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0230m, and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0231m. Three reference fields u(i) (x) are computed by solving Equations 1–3 with y(x) given by the RF1 y field and subject to the Dirichlet BCs urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0232 (i = 1, 2, 3). The vector urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0233 is drawn from each reference field u(i)(x).

We compute Cu(x′, x″) and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0234 from MCS for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0235. The mean u fields urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0236 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0237 are approximated by solving Equations 22–24. Figure 8 shows the PICKLE estimates urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0238 and the corresponding point errors with respect to the reference fields u(i) (i = 1, 2, 3). For all three fields, the errors in the estimated u fields are similar, with the average relative errors less than 0.5% and maximum point errors less than 4%. These results show that the PICKLE model trained for one boundary condition (urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0239 in this case) can be used to accurately predict the u field for the other boundary conditions. Note that in general, the PICKLE estimate of y from the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0240 and ys measurements could depend on i because the parameter estimation is an ill-posed problem. However, for this setting we find that the PICKLE estimates of y are within 0.01% of each other.

Details are in the caption following the image

The PICKLE estimates urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0241 (i = 1, 2, 3) (top row) and the point errors with respect to the reference fields u(i) (bottom row). The dots are locations of the wells where u measurements are available.

Finally, we note that the covariance Cu(x′, x″) and the mean hydraulic head urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0242 field can be used to model measurements urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0243 by using GPR. Such a “physics-informed” GPR approach will result in the conditional mean head prediction
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0244(25)
and the conditional variance of u (that can be considered as the measure of uncertainty in the GPR model of u)
urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0245(26)
where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0246 is the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0247 observation covariance matrix with elements urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0248, Cu(x) is the urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0249-dimensional vector with components Cu,k(x) = Cu(x, xk), and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0250 is the vector with components urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0251, where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0252. We refer to the GPR model of Equations 25 and 26 as “physics-informed” GPR because the (prior) mean and covariance of u, Cu(x′, x″), and the mean hydraulic head urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0253 are computed from a physics model (i.e., Equations 1 and 3) rather than from u measurements. Further discussion on physics-informed GPR methods can be found in X. Yang et al. (20212019).

6 Conclusions

We proposed the PICKLE method for assimilating data in models with unknown and time-changing boundary conditions, and used it to estimate the transmissivity and hydraulic head fields for the two-dimensional steady-state groundwater model of the Hanford Site with incrementally varying-in-time Dirichlet boundary conditions and uncertain Neumann boundary conditions. The PICKLE method is based on the approximation of unknown parameters and state variables via CKLEs. The CKLE approximation of a field enforces (exactly matches) the field measurements and the covariance structure, that is, it models the field as a realization of a conditional Gaussian field with a prescribed covariance function. To test the applicability of CKLE-based approximations for natural systems such as the Hanford Site, we considered two reference log-transmissivity fields, both representing the complexity of the Hanford Site. The first transmissivity field, referred to as the RF1 field, is constructed by depth-averaging the conductivity field obtained in a previous calibration study that did not make any regularizing Gaussianity assumptions. The RF2 reference log-transmissivity field was constructed using GPR and 50 values of the RF1 log-transmissivity field at locations randomly selected from 323 Hanford Site well locations. By construction, the RF2 field is smoother than the RF1 field. The comparison against the MAP method, a standard method for solving inverse problems, for the RF1 and RF2 reference fields reveals the following relative advantages and disadvantages of the PICKLE method:
  1. For the synthetic data generated from the RF1 and RF2 y fields, we demonstrated that the MAP and PICKLE execution times scale with mesh resolution as N1.15 and N3.27, respectively, where N is the number of FV cells. The close to linear dependence of PICKLE's execution time on the problem size gives PICKLE a computational advantage over the MAP method for large-scale problems. We consider this to be the main advantage of the PICKLE method

  2. For the same number of measurements, the accuracy of PICKLE and MAP depends on the measurement locations. The MAP method is more accurate for the RF1 field, and the PICKLE method is more accurate for the RF2 field for most considered cases

  3. The execution time of PICKLE and MAP increases and the accuracy decreases as the roughness of the parameter field increases. This is expected for most inversion methods for ill-posed problems that rely on a smoothness-enforcing regularization. Therefore, for smooth fields, the regularized inverse solutions are expected to be more accurate. For the same reason, iterative optimization methods for inverse problems are expected to converge faster for smooth fields

  4. In the PICKLE method, the execution time and accuracy increase with the increasing number of CKL terms. In this work, as a baseline we used Ny = Nu = 1000 that corresponds to rtol < 10−6. We stipulate that this criterion is sufficient to obtain a convergent estimate of y with respect to the number of CKL terms

  5. The training of the PICKLE model should be performed only for one value of the boundary conditions and does not need to be updated as the boundary conditions change, which significantly reduces its cost

  6. The accuracy of the PICKLE method depends on the ability of the truncated CKLEs to accurately approximate y and u, which requires a certain degree of smoothness of the considered fields. We demonstrated that for y and u fields that are representative of the Handford Site, the CKLE approximations of the fields lead to results that are comparable in accuracy to the MAP method. However, CKLEs can also be used to approximate fields exhibiting step-like changes (e.g., at the boundaries of different geological formations) using a logistic function as was shown in A. Tartakovsky et al. (2020)

  7. In the PICKLE method, computing the covariance function of u from MCS can become a computational bottleneck for large-scale problems. However, MCS can be replaced with more computationally efficient alternatives, including the multilevel MC method, generative physics-informed machine learning models, Polynomial Chaos and other surrogate models, and the moment equation method

Finally, it is important to note that the PICKLE method is not limited to steady-state problems. In fact, the extension of the PICKLE formulation (11) to time-dependent problem is straightforward and only requires replacing the evaluation of the residuals of the steady-state flow equation with the time-dependent flow equation. Constructing the CKLE (6) for a time-dependent solution field is the same as for one that is only a function of space. For a time-space-dependent problem, x in Section 6 will be a vector of n spatial coordinates and time. The covariance of u, which is needed to compute the eigenfunctions and values in Section 6, will be in both space and time—however, it can be computed from the MC solution just as in the considered steady-state problem.

Acknowledgments

This research was partially supported by the U.S. Department of Energy (DOE) Advanced Scientific Computing (ASCR) program. Pacific Northwest National Laboratory is operated by Battelle for the DOE under Contract DE-AC05-76RL01830.

    Appendix A: Finite Volume Discretization

    Figure A1 shows two adjacent cells in the finite-volume discretization model, with their centers labeled as pi and pj, respectively. In this model, we assume that the transmissivity is constant within each cell i and its average value Ti is at its center. The edge shared by the two cells, ei,j, has length |ei,j|. The stiffness matrix A in Equation 4 is defined as
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0254(A1)
    where
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0255
    Details are in the caption following the image

    Two adjacent cells used to define the two-point finite-volume discretization model.

    and Ti is the transmissivity of the ith cell. The right-hand side b describes the boundary conditions, and is defined as
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0256(A2)
    where τi is the transmissivity between the ith cell and the boundary.

    Appendix B: Computing MAP Estimates

    In this work, we compute MAP estimates by recasting the PDE-constrained optimization problem of Equation 5 into an unconstrained nonlinear least-squares problem. Specifically, we aim to solve the problem
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0257(B1)
    with cost vector given by
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0258(B2)
    where Hu and Hy are the observation matrices in Equation 5. Note that in Equation B1, we fold the PDE constraint into the cost function by treating u explicitly as a function of the parameters p.
    For solving this least-square minimization problem, it is necessary to compute the Jacobian matrix of the cost vector with respect to the parameters p than is given by
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0259(B3)
    It can be seen that the first block row of Equation B3 corresponds to the Jacobian of u with respect to the parameters p, for which we derive a formula as follows: Differentiating Equation 4 with respect to the ith component of p, we obtain
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0260
    therefore,
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0261(B4)

    Appendix C: Solver Optimization

    We implemented our solvers for both PICKLE and MAP in Python. For both methods, we employ TPFA-FV as the forward problem. Although we did not parallelize the solvers used in this work, we optimized the codes in several ways described as follows.

    C1 Precomputing Matrices

    Because the properties of each cell, the observation locations of us, ys, and the topology of the cell connections are fixed, the structure (i.e., the location of non-zero entries) of the observation matrices Hu, Hy, the regularization matrix D, the stiffness matrix A in Equation 4, and the partial derivatives in the first block row of Equation B3 also remain unchanged throughout the least-squares minimization of Equations 5 and 14. Thus, these fixed structures can be identified in advance, and only the values of A and the partial derivatives in Equation B4 need to be updated at each minimization iteration. In addition, when the boundary conditions are known and constant in time, the aggregated contribution of the prescribed hydraulic head urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0262 and normal flux urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0263 to each FV cell iurn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0264 and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0265 in Equation A2—can also be precomputed. For MAP, the second to the fourth block rows and Hu in the first block row of the Jacobian in Equation B3 are also constant throughout minimization because they only depend on the topology of the mesh. Therefore, these elements can also be precomputed ahead of time.

    C2 Sparsity

    Sparsity is maintained throughout the evaluations of the objective functions of both PICKLE and MAP, including the residual l(u, y, q) in Equation 4, as well as their corresponding Jacobian matrices. This significantly reduces the storage and computation overhead because the increase in the resolution of the mesh quadruples the size of the matrices. However, the SciPy implementation of the sparse linear solver (spsolve) does not support sparse right-hand-side vectors and matrices. Furthermore, partial solvers that only compute solutions at measurement locations and reuse sparse structural reordering are not supported by the package. Future optimization using these techniques would further reduce the execution times of both the MAP and PICKLE methods.

    Appendix D: Perturbative Expression for the Head Covariance

    In this section, we aim to derive closed-form perturbative approximations to the covariance of hydraulic head by treating the transmissivity field as a random field. Let urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0266 denote ensemble averaging. We assume that the transmissivity field is written as ensemble average, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0267, plus a zero-mean random fluctuation T′(x), that is, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0268, with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0269. Similarly, the hydraulic head is decomposed into its ensemble mean and a zero-mean deviation, that is, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0270, with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0271.

    According to D. M. Tartakovsky and Neuman (1998), for Equations 1–3 the hydraulic head deviation is given by
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0272(D1)
    where G(y, x) is the Green's function that solves the deterministic PDE problem.
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0273(D2)
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0274(D3)
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0275(D4)
    From the above, it follows that the hydraulic head covariance is given by
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0276
    Employing the localization approximation (D. M. Tartakovsky et al., 2002), the integrals above can be approximated by “pulling” urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0277 outside the integral, leading to the approximations
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0278
    where urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0279, and
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0280(D5)
    Employing the definition of g(y, z) given above, Equation D5 can be rewritten as
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0281(D6)
    with the tensor A(x, z) given by
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0282(D7)
    The expression for the hydraulic head covariance is completed by providing an approximation for urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0283. Specifically, for small variance of K′, urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0284 approximately solves the “mean field” problem.
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0285(D8)
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0286(D9)
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0287(D10)

    In Equation D6, A(x, z) depends on the homogeneous Dirichlet and Neumann boundary conditions, while urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0289 depends on the heterogeneous Dirichlet and Neumann boundary conditions that describe the actual boundary conditions of the modeled system. Therefore, to compute urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0290 for any heterogeneous boundary conditions one only needs to solve the deterministic Equation D8 subject to these heterogeneous boundary conditions. The expensive part of computing urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0291 is to evaluate A(x, z), which needs to be done only once for the homogeneous boundary conditions.

    Evaluating A(x, z) from Equation D5 requires computing the Green's function, which involves solving the deterministic problem (EquationsD2–D4) N times (this number can be reduced if the symmetry G(y, x) = G(x, y) is taken into account), where N is the number of nodes in the computational domain discretization. In this work, we employ a different strategy. We compute urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0292 for one set of heterogeneous boundary conditions urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0293 (and urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0294) corresponding to the average state of the modeled system using the Monte Carlo simulation method described in Section 3.2. Next, we perform the following order of magnitude analysis to describe the change in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0295 due to an ϵ change in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0296:

    It follows from the linearity of Equations D8–D10 with respect to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0297 that O(ϵ) changes in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0298 lead to O(ϵ) changes in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0299. Replacing urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0300 with urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0301 in Equation D6 we obtain
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0302
    Note that urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0303, where L is the domain size in the direction of the mean flow. Then, it follows that
    urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0304

    which states that an ϵ change in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0305 leads to an O(ϵ/L) change in the covariance of hydraulic head. At the Hanford Site, L is on the order of 104 m and ϵ (the natural variations in Dirichlet boundary conditions reflecting water level changes in the Columbia river) is on the order of 1 m. Therefore, for the Hanford Site, the natural variations in the Dirichlet boundary condition lead to urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0306 changes in urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0307. We assume that this sufficiently small change can be disregarded, and that the covariance urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0308, which is computed for the boundary condition urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0309, can be used to model u data corresponding to the boundary condition urn:x-wiley:00431397:media:wrcr25973:wrcr25973-math-0310.

    Data Availability Statement

    The data and codes used in this paper are available at https://zenodo.org/record/6512404#.YnA5hPPMKBR.