License: arXiv.org perpetual non-exclusive license
arXiv:2402.17698v1 [math.NA] 27 Feb 2024

Learning reduced-order Quadratic-Linear models in Process Engineering using Operator Inference

Ion Victor Gosea [email protected] Luisa Peterson [email protected] Pawan Goyal [email protected] Jens Bremer [email protected] Kai Sundmacher [email protected] Peter Benner [email protected]
Abstract

In this work, we address the challenge of efficiently modeling dynamical systems in process engineering. We use reduced-order model learning, specifically operator inference. This is a non-intrusive, data-driven method for learning dynamical systems from time-domain data. The application in our study is carbon dioxide methanation, an important reaction within the Power-to-X framework, to demonstrate its potential. The numerical results show the ability of the reduced-order models constructed with operator inference to provide a reduced yet accurate surrogate solution. This represents an important milestone towards the implementation of fast and reliable digital twin architectures.

keywords:
scientific machine learning, operator inference, non intrusive methods, quadratic systems, process engineering, chemical engineering, methanation reactor.
\novelty

This contribution illustrates the robustness and approximation capabilities of the operator inference method for a test case from process engineering.

1 Introduction

Dynamical models are particularly relevant for several important tasks and applications in the field of engineering science, such as analyzing transient behavior under operating conditions, enforcing parameter optimization, enabling long-time horizon prediction, and aiding control techniques. Especially in the case of high-fidelity models, their state-space dimension can easily grow to the order of tens or even hundreds of thousands. In such cases, handling such models becomes computationally prohibitive with respect to storage or time constraints.

Model order reduction (MOR) and reduced-order modeling (RoMod) refer to classes of methodologies that can be used to simplify complex dynamical models or to identify the underlying dynamics from data while enforcing an accurate approximation and using as little computational effort as possible. Traditionally, reduced-order models (ROMs) constructed through intrusive MOR approaches rely on the availability of the underlying mathematical equations (state-space realizations), i.e., on the actual models. For an overview of classical, intrusive (projection-based) approaches, we refer the reader to [1, 6]. An advantage of such techniques is the existence of rigorous theoretical guarantees by means of a posteriori error estimation, or by stability/passivity enforcement. Moreover, established MOR techniques represent a backbone for development of digital twin frameworks [15, 24]. A potential drawback is given by the fact that the ROMs computed via intrusive methods employ explicit projection of the governing equations onto low-dimensional dominant subspaces [5]. Hence, access to the underlying mathematical equations that produced the (high-fidelity) simulations is typically required. Such access is not always provided and hence, represents a challenge for incorporating classical MOR tools for efficient deployment to predominant data-based environments. For the latter, such an accurate/exact description is unavailable (or not fully available), in many practical scenarios. These include applications from electrical, mechanical or from process/chemical engineering, as in the current work.

An alternative approach to classical (intrusive) methods of MOR, which rely on explicit access to a large-scale model, is the use of data-driven (non-intrusive) techniques based on measurements or simulation data. Unlike intrusive methods, data-driven RoMod does not require explicit knowledge of the model structure or matrices. Instead, low-order models can be directly constructed by using solely time-domain data, such as snapshots of the system’s state-space evolution, along with snapshots of control inputs or of the observed outputs. Among such methods, Operator Inference (OpInf) introduced in [26], has established itself in the last decade as a reliable RoMod and learning approach that exploits the inherent structure of physics-based models to efficiently capture the underlying system’s dynamics. OpInf constructs ROMs in a non-intrusive manner via a data-driven regression problem that learns reduced matrices from snapshot data. Typically, the learned dynamics of the ROM depend linearly and quadratically on the state variables (in reduced coordinates), although extensions to fitting other structures were also proposed.

The efficient modeling, simulation, and complexity reduction of dynamical systems in the field of process and chemical engineering has become pivotal in today’s industrial landscape, especially with the ever-growing digitalization trend in modern plants and units in the age of Industry 4.0. Driven by this motivation, our work proposes an application of established reduced-order modeling techniques, focusing on the OpInf method, for a test case involving a CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation reactor [34]. The preliminary results reported here provide insights into the usability of the surrogate ROMs computed with OpInf. The approximation quality attained offers promising prospects for further developments and extensions. Finally, this approach is well-suited for handling the intricacies of the highly nonlinear process under study. This is essential to ensure computational efficiency and approximation accuracy for applications in process engineering and other related disciplines.

This paper is organized as follows. After the introduction and motivation are set up, in Section 2 we provide a comprehensive account of state-of-the-art methods, together with their applications and extensions. Then, in Section 3, we introduce the main method of interest, OpInf, together with its theoretical background and some recent innovations. Section 4 illustrates the application of interest, a CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation reactor characterized by coupled nonlinear PDEs. We present various numerical results when applying OpInf to this test case, while in Section 5 the conclusions and future research endeavors are summarized.

2 Data-driven methods

Non-intrusive data-driven RoMod methods offer an alternative to conventional MOR methods, by means of constructing ROMs directly from data, bypassing the need for accessing an explicit realization. System identification (SI) methods [19] are considered the precursor of RoMod methods in the field of systems and control engineering. By means of incorporating input-output data in the identification process, such methods were fairly prolific in the last decades. The subspace identification method [32] is one of these and it estimates linear time-invariant state-space models using only samples of the input and output signals. Another class of methods uses rational approximation techniques to identify a ROM from frequency-domain data, such as Loewner-matrix approaches [21]. Dynamic Mode Decomposition (DMD) [18] is used to extract the dominant dynamic modes of a system and is closely connected to the concept of Koopman operator. In the field of Scientific Machine Learning (SciML), several techniques have emerged to address the challenges and requirements of dealing with complex scientific problems, in the presence of simulation data. Selecting the most appropriate ML algorithm can be both a challenging and a tedious task, due to the variety of algorithms, computer architectures, and ML models available. Artificial Neural Network (ANN) architectures are particularly attractive for SciML applications due to their ability to handle noisy or inexact data, perform automatic differentiation, and their flexibility as mesh-free models. Therefore, there is a growing interest in combining ANN-based solutions with traditional physics-based approaches to improve performance and reliability. Some recent works in this direction are [29, 20, 11]. SINDy, as introduced by [8], utilizes scientific knowledge to improve the fitted model performance and constructs parsimonious models by selecting only a few terms from a library of candidate functions (sparse identification).

OpInf involves fitting structured models in reduced coordinates by means of solving a regression problem formed from snapshot data (trajectories of the state evolution in the time domain). This approach provides a versatile framework for the analysis and modeling of continuous-time processes in chemical engineering. For example, the continuous-time behavior of the temperature and conversion profile of a chemical reactor can be accurately represented using OpInf, as will be demonstrated in this paper. As far as the authors are aware, together with the recent contribution [27], the current work is the first contribution towards applying OpInf to process engineering problems. Typically, OpInf was successfully applied to test cases from computational fluid dynamics [3], mechanics [9], or aeronautics [22]. In recent years, OpInf was extended also to cope with higher-order polynomial terms [28], or even with non-polynomial terms [4]. Additionally, parametric problems were treated [33, 23]. Then, contributions that propose regularizing the learning problem to enable performance on large-scale systems were proposed in [22], while problems with noisy or low-quality data sets were treated in [31]. Dealing with differential-algebraic equations (DAEs) for incompressible flow problems was treated in [3]. The problem of stability of the ROMs constructed via OpInf was tackled in [13, 30]. In [11], the authors propose combining the OpInf approach with certain deep ANN approaches, to infer the unknown nonlinear dynamics of the system. For more references, detailed principles, and technical details of OpInf, we refer the reader to the recent comprehensive survey paper [17]. One possible downside of this approach may be that it requires direct access or approximation of temporal gradient information. From a numerical point of view, errors might arise due to the inherent gradient approximation. This issue has been recently addressed in [12], in which Runge-Kutta numerical integration schemes have been successfully integrated into the system identification process, thereby avoiding the need for collecting derivative information.

3 The method under investigation: operator inference

3.1 General description

One of the strengths of the OpInf methodology is that it makes use of the known physical structures and constraints at a qualitative level. That is, one can assume nonlinearities of quadratic structure (these appear naturally in many established flow problems, such as in the Navier-Stokes or Burgers’ equations). Even when the dynamics are characterized by other nonlinearities (high-degree polynomials or more generic, analytic ones), one can equivalently embed them into a quadratic manifold by means of lifting [14, 2].

Consider a quadratic full-order model (FOM) described by

𝐱˙(t)=𝐀𝐱(t)+𝐇(𝐱(t)𝐱(t))+𝐁,˙𝐱𝑡𝐀𝐱𝑡𝐇tensor-product𝐱𝑡𝐱𝑡𝐁\dot{\mathbf{x}}(t)={\mathbf{A}}{\mathbf{x}}(t)+{\mathbf{H}}({\mathbf{x}}(t)% \otimes{\mathbf{x}}(t))+{\mathbf{B}},over˙ start_ARG bold_x end_ARG ( italic_t ) = bold_Ax ( italic_t ) + bold_H ( bold_x ( italic_t ) ⊗ bold_x ( italic_t ) ) + bold_B , (1)

where 𝐱(t)n,𝐀n×n,𝐇n×n2formulae-sequence𝐱𝑡superscript𝑛formulae-sequence𝐀superscript𝑛𝑛𝐇superscript𝑛superscript𝑛2{\mathbf{x}}(t)\in\mathbb{R}^{n},{\mathbf{A}}\in\mathbb{R}^{n\times n},{% \mathbf{H}}\in\mathbb{R}^{n\times n^{2}}bold_x ( italic_t ) ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , bold_A ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × italic_n end_POSTSUPERSCRIPT , bold_H ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, and 𝐁n𝐁superscript𝑛{\mathbf{B}}\in\mathbb{R}^{n}bold_B ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Data preparation is an important step of the method, i.e., the matrix 𝐗𝐗{\mathbf{X}}bold_X of snapshots of the state variables in (5). As we will see below, we can perform this step also in reduced dimension. One needs to build, in the beginning, the following time-domain snapshot matrix for time instances ti>0subscript𝑡𝑖0t_{i}>0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > 0 and 𝐱i:=𝐱(ti)assignsubscript𝐱𝑖𝐱subscript𝑡𝑖{\mathbf{x}}_{i}:={\mathbf{x}}(t_{i})bold_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT := bold_x ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) with 1ik1𝑖𝑘1\leq i\leq k1 ≤ italic_i ≤ italic_k, as

𝐗𝐗\displaystyle{{\mathbf{X}}}bold_X :=[   𝐱0𝐱1𝐱k   ]n×(k+1).assignabsentdelimited-[]  missing-subexpression subscript𝐱0subscript𝐱1subscript𝐱𝑘  missing-subexpression superscript𝑛𝑘1\displaystyle:=\left[\begin{array}[]{llll}\leavevmode\nobreak\ \rule[-4.30554% pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385% pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}\\ {\mathbf{x}}_{0}&{\mathbf{x}}_{1}&\cdots&{\mathbf{x}}_{k}\\ \leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak% \ \rule[-4.30554pt]{0.5pt}{10.76385pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]% {0.5pt}{10.76385pt}\end{array}\right]\in\mathbb{R}^{n\times(k+1)}.:= [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL bold_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL ⋯ end_CELL start_CELL bold_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW end_ARRAY ] ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × ( italic_k + 1 ) end_POSTSUPERSCRIPT . (5)

Then we compute a projection matrix 𝐕n×r𝐕superscript𝑛𝑟{\mathbf{V}}\in\mathbb{R}^{n\times r}bold_V ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × italic_r end_POSTSUPERSCRIPT using r𝑟ritalic_r dominant Proper Orthogonal Decomposition (POD) basis vectors (see Ch.3 in [16] for more details) so that 𝐗𝐕𝐕𝐗τnorm𝐗superscript𝐕𝐕top𝐗𝜏\|{\mathbf{X}}-{\mathbf{V}}{\mathbf{V}}^{\top}{\mathbf{X}}\|\leq\tau∥ bold_X - bold_VV start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_X ∥ ≤ italic_τ, depending on a tolerance value τ>0𝜏0\tau>0italic_τ > 0. Typically, the matrix 𝐕𝐕{\mathbf{V}}bold_V is chosen by performing a Singular Value Decomposition (SVD) of 𝐗𝐗{\mathbf{X}}bold_X, and by assembling the leading r𝑟ritalic_r right singular vectors of 𝐗𝐗{\mathbf{X}}bold_X.

By setting 𝐱^(t):=𝐕𝐱(t)assign^𝐱𝑡superscript𝐕top𝐱𝑡\hat{{\mathbf{x}}}(t):={\mathbf{V}}^{\top}{\mathbf{x}}(t)over^ start_ARG bold_x end_ARG ( italic_t ) := bold_V start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_x ( italic_t ), the intrusively calculated ROM via the POD approach is computed via a Galerkin projection: 𝐀^=𝐕𝐀𝐕,𝐇^=𝐕𝐇(𝐕𝐕)formulae-sequence^𝐀superscript𝐕top𝐀𝐕^𝐇superscript𝐕top𝐇tensor-product𝐕𝐕\hat{{\mathbf{A}}}={\mathbf{V}}^{\top}{\mathbf{A}}{\mathbf{V}},\hat{{\mathbf{H% }}}={\mathbf{V}}^{\top}{\mathbf{H}}({\mathbf{V}}\otimes{\mathbf{V}})over^ start_ARG bold_A end_ARG = bold_V start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_AV , over^ start_ARG bold_H end_ARG = bold_V start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_H ( bold_V ⊗ bold_V ), and 𝐁^=𝐕𝐁^𝐁superscript𝐕top𝐁\hat{{\mathbf{B}}}={\mathbf{V}}^{\top}{\mathbf{B}}over^ start_ARG bold_B end_ARG = bold_V start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_B. However, the key feature of OpInf is to infer these matrices in a non-intrusive manner, i.e., without using the FOM matrices in (1).

The reduced-order state data matrices can be put together using the compressed snapshots 𝐱^i=𝐕𝐱isubscript^𝐱𝑖superscript𝐕topsubscript𝐱𝑖\hat{\mathbf{x}}_{i}={\mathbf{V}}^{\top}{\mathbf{x}}_{i}over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = bold_V start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, 𝐱^i=𝐱^i𝐱^isuperscriptsubscript^𝐱𝑖tensor-producttensor-productsubscript^𝐱𝑖subscript^𝐱𝑖\hat{\mathbf{x}}_{i}^{\otimes}=\hat{\mathbf{x}}_{i}\otimes\hat{\mathbf{x}}_{i}over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT = over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⊗ over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and also the time-derivative data matrix can be put together from estimated snapshots 𝐱^˙isubscript˙^𝐱𝑖\dot{\hat{\mathbf{x}}}_{i}over˙ start_ARG over^ start_ARG bold_x end_ARG end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT using 𝐗^^𝐗\hat{{\mathbf{X}}}over^ start_ARG bold_X end_ARG, e.g. by employing a time derivative approximation, as follows:

𝐗^^𝐗\displaystyle\hat{{\mathbf{X}}}over^ start_ARG bold_X end_ARG :=[   𝐱^0𝐱^1𝐱^k   ],𝐗^:=[   𝐱^0𝐱^1𝐱^k   ],𝐗^˙=[   𝐱˙^0𝐱^˙1𝐱^˙k   ].formulae-sequenceassignabsentdelimited-[]  missing-subexpression subscript^𝐱0subscript^𝐱1subscript^𝐱𝑘  missing-subexpression formulae-sequenceassignsuperscript^𝐗tensor-productdelimited-[]  missing-subexpression superscriptsubscript^𝐱0tensor-productsuperscriptsubscript^𝐱1tensor-productsuperscriptsubscript^𝐱𝑘tensor-product  missing-subexpression ˙^𝐗delimited-[]  missing-subexpression subscript^˙𝐱0subscript˙^𝐱1subscript˙^𝐱𝑘  missing-subexpression \displaystyle:=\left[\begin{array}[]{llll}\leavevmode\nobreak\ \rule[-4.30554% pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385% pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}\\ \hat{\mathbf{x}}_{0}&\hat{\mathbf{x}}_{1}&\cdots&\hat{\mathbf{x}}_{k}\\ \leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak% \ \rule[-4.30554pt]{0.5pt}{10.76385pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]% {0.5pt}{10.76385pt}\end{array}\right],\quad\hat{{\mathbf{X}}}^{\otimes}:=\left% [\begin{array}[]{llll}\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt% }&\leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&&\leavevmode% \nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}\\ \hat{\mathbf{x}}_{0}^{\otimes}&\hat{\mathbf{x}}_{1}^{\otimes}&\cdots&\hat{% \mathbf{x}}_{k}^{\otimes}\\ \leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak% \ \rule[-4.30554pt]{0.5pt}{10.76385pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]% {0.5pt}{10.76385pt}\end{array}\right],\quad\dot{\hat{{\mathbf{X}}}}=\left[% \begin{array}[]{llll}\leavevmode\nobreak\ \rule{0.4pt}{6.45831pt}&\leavevmode% \nobreak\ \rule{0.4pt}{6.45831pt}&&\leavevmode\nobreak\ \rule{0.4pt}{6.45831pt% }\\ \hat{\dot{\mathbf{x}}}_{0}&\dot{\hat{\mathbf{x}}}_{1}&\cdots&\dot{\hat{\mathbf% {x}}}_{k}\\ \leavevmode\nobreak\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\leavevmode\nobreak% \ \rule[-4.30554pt]{0.5pt}{10.76385pt}&&\leavevmode\nobreak\ \rule[-4.30554pt]% {0.5pt}{10.76385pt}\end{array}\right].:= [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL ⋯ end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW end_ARRAY ] , over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT := [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT end_CELL start_CELL ⋯ end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW end_ARRAY ] , over˙ start_ARG over^ start_ARG bold_X end_ARG end_ARG = [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over^ start_ARG over˙ start_ARG bold_x end_ARG end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL over˙ start_ARG over^ start_ARG bold_x end_ARG end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL ⋯ end_CELL start_CELL over˙ start_ARG over^ start_ARG bold_x end_ARG end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW end_ARRAY ] .

Then, a ROM of the form

𝐱^˙(t)=𝐀^𝐱^(t)+𝐇^(𝐱^(t)𝐱^(t))+𝐁^,˙^𝐱𝑡^𝐀^𝐱𝑡^𝐇tensor-product^𝐱𝑡^𝐱𝑡^𝐁\dot{\hat{\mathbf{x}}}(t)=\hat{\mathbf{A}}\hat{\mathbf{x}}(t)+\hat{\mathbf{H}}% (\hat{\mathbf{x}}(t)\otimes\hat{\mathbf{x}}(t))+\hat{\mathbf{B}},over˙ start_ARG over^ start_ARG bold_x end_ARG end_ARG ( italic_t ) = over^ start_ARG bold_A end_ARG over^ start_ARG bold_x end_ARG ( italic_t ) + over^ start_ARG bold_H end_ARG ( over^ start_ARG bold_x end_ARG ( italic_t ) ⊗ over^ start_ARG bold_x end_ARG ( italic_t ) ) + over^ start_ARG bold_B end_ARG ,

can be obtained using projected data by solving the following optimization problem:

min𝐀^,𝐇^,𝐁^𝐗^˙𝐀^𝐗^𝐇^𝐗^𝐁^+α(𝐀^,𝐇^).subscript^𝐀^𝐇^𝐁norm˙^𝐗^𝐀^𝐗^𝐇superscript^𝐗tensor-product^𝐁𝛼^𝐀^𝐇\min_{\hat{\mathbf{A}},\hat{\mathbf{H}},\hat{\mathbf{B}}}\left\|\dot{\hat{% \mathbf{X}}}-\hat{\mathbf{A}}\hat{\mathbf{X}}-\hat{\mathbf{H}}\hat{\mathbf{X}}% ^{\otimes}-\hat{\mathbf{B}}\right\|+\alpha{\cal R}(\hat{\mathbf{A}},\hat{% \mathbf{H}}).roman_min start_POSTSUBSCRIPT over^ start_ARG bold_A end_ARG , over^ start_ARG bold_H end_ARG , over^ start_ARG bold_B end_ARG end_POSTSUBSCRIPT ∥ over˙ start_ARG over^ start_ARG bold_X end_ARG end_ARG - over^ start_ARG bold_A end_ARG over^ start_ARG bold_X end_ARG - over^ start_ARG bold_H end_ARG over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT - over^ start_ARG bold_B end_ARG ∥ + italic_α caligraphic_R ( over^ start_ARG bold_A end_ARG , over^ start_ARG bold_H end_ARG ) . (6)

Typically, the least squares (LS) problem in (6) is ill-conditioned, and hence requires regularization techniques (such as the Tikhonov approach, see [33, 22] in the context of OpInf), i.e., encoded in the term (𝐀^,𝐇^)^𝐀^𝐇{\cal R}(\hat{\mathbf{A}},\hat{\mathbf{H}})caligraphic_R ( over^ start_ARG bold_A end_ARG , over^ start_ARG bold_H end_ARG ) above. We refer the reader to Section 4 in [3], for details on implementation aspects for this basic, “vanilla version” of OpInf.

3.2 Solving the optimization problem

We note that, in the OpInf algorithm, the FOM is never explicitly used, at any step. When α=0𝛼0\alpha=0italic_α = 0 (no regularization case), the LS problem can be written as

min𝐀^,𝐇^,𝐁^𝐗^˙[𝐀^𝐇^𝐁^]𝒟,where𝒟=[𝐗^𝐗^𝟏T],subscript^𝐀^𝐇^𝐁norm˙^𝐗matrix^𝐀^𝐇^𝐁𝒟where𝒟matrix^𝐗superscript^𝐗tensor-productsuperscript1𝑇\displaystyle\min_{\hat{\mathbf{A}},\hat{\mathbf{H}},\hat{\mathbf{B}}}\left\|% \dot{\hat{\mathbf{X}}}-\begin{bmatrix}\hat{\mathbf{A}}&\hat{\mathbf{H}}&\hat{% \mathbf{B}}\end{bmatrix}\mathcal{D}\right\|,\ \text{where}\ \mathcal{D}=\begin% {bmatrix}\hat{{\mathbf{X}}}\\ \hat{{\mathbf{X}}}^{\otimes}\\ \mathbf{1}^{T}\end{bmatrix},roman_min start_POSTSUBSCRIPT over^ start_ARG bold_A end_ARG , over^ start_ARG bold_H end_ARG , over^ start_ARG bold_B end_ARG end_POSTSUBSCRIPT ∥ over˙ start_ARG over^ start_ARG bold_X end_ARG end_ARG - [ start_ARG start_ROW start_CELL over^ start_ARG bold_A end_ARG end_CELL start_CELL over^ start_ARG bold_H end_ARG end_CELL start_CELL over^ start_ARG bold_B end_ARG end_CELL end_ROW end_ARG ] caligraphic_D ∥ , where caligraphic_D = [ start_ARG start_ROW start_CELL over^ start_ARG bold_X end_ARG end_CELL end_ROW start_ROW start_CELL over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT ⊗ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL bold_1 start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] , (11)

with the matrix 𝒟(r+r2+1)×(k+1)𝒟superscript𝑟superscript𝑟21𝑘1\mathcal{D}\in\mathcal{R}^{(r+r^{2}+1)\times(k+1)}caligraphic_D ∈ caligraphic_R start_POSTSUPERSCRIPT ( italic_r + italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 ) × ( italic_k + 1 ) end_POSTSUPERSCRIPT constructed in terms of the projected states. Based on the singular value decay of 𝒟𝒟\mathcal{D}caligraphic_D (and on a tolerance value), one can choose a suitable truncation order r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG. More precisely, one can write the solution of (11) as 𝐗^˙𝒟˙^𝐗superscript𝒟\dot{\hat{\mathbf{X}}}\mathcal{D}^{\dagger}over˙ start_ARG over^ start_ARG bold_X end_ARG end_ARG caligraphic_D start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, where 𝒟superscript𝒟\mathcal{D}^{\dagger}caligraphic_D start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT is the pseudo-inverse of matrix 𝒟𝒟\mathcal{D}caligraphic_D. This can be computed through a truncated SVD of 𝒟𝒟\mathcal{D}caligraphic_D, kee** only the r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG dominant singular values and vectors [3].

However, matrix 𝒟𝒟\mathcal{D}caligraphic_D may be ill-conditioned, in practice. Typically, a mixed SVD-regularization approach is to be used in practice, since choosing fewer singular values from 𝒟𝒟\mathcal{D}caligraphic_D (small r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG) yields a larger mismatch of the data fidelity term, whereas, for larger r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG, the problem becomes more and more ill-conditioned. An important property of OpInf is that it recovers the intrusive POD reduced model if data are Markovian, as shown in [25]. Then, a paramount innovation in the OpInf framework was made in [13], in which global-stability-informed learning was achieved through quadratic system parameterizations. There, the authors first provide ways of enforcing local asymptotic stability, by imposing that the matrix 𝐀^^𝐀\hat{{\mathbf{A}}}over^ start_ARG bold_A end_ARG of the learned model has all its eigenvalues in the left-half complex plane (by imposing a parameterization of 𝐀^^𝐀\hat{{\mathbf{A}}}over^ start_ARG bold_A end_ARG in terms of symmetric positive definite (SPD) and skew-symmetric matrices). Then, an extension is proposed, that enforces global asymptotic stability by means of specially-tailored parameterization for the 𝐇^^𝐇\hat{{\mathbf{H}}}over^ start_ARG bold_H end_ARG operator of the learned quadratic model. Then, a gradient-based approach is proposed in order to obtain a solution to the optimization problem in (6), which now has additional constraints on the matrices involved. By relaxing the SPD condition, and by rewriting the problem of inferring operators via an integral form of the differential equation. This circumvents the need to incorporate derivative information, at the cost of a more complicated setup to be solved, involving integrals. By making use of a fourth-order Runge-Kutta scheme, the latter problem is approximated and solved iteratively, as described in detail in Section 6 of [13]. In the numerics part, i.e., in Section 4.1, we leverage these results and provide additional explanations.

4 The application of interest

The application of interest, i.e., CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation, represents an important process in the Power-to-X framework, which plays a key role in the storage of energy from renewable sources [10]. CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation, i.e., the conversion of carbon dioxide (CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT) and hydrogen to methane and water, is a key aspect of this framework, facilitating the recycling of CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT emissions. The exothermic methanation reaction releases heat, which commonly leads to hotspots (localized areas of elevated temperatures) that affect temperature control and reactor efficiency. Effective management of these hotspots is critical, especially in reactors with variable energy inputs from renewable sources, such as wind or solar energy [7].

Our study investigates a one-dimensional, pseudo-homogeneous reactor model for catalytic CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation, following [34]. We employ a simplification by excluding the consideration of effective axial mass dispersion in our analysis. The behavior of the reactor is captured by coupled PDEs. These governing equations are written in terms of the energy balance, represented by the temperature variable T𝑇Titalic_T, and the mass balance, represented by the CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT conversion variable X𝑋Xitalic_X, as:

εRXtsubscript𝜀R𝑋𝑡\displaystyle\varepsilon_{\textrm{R}}\frac{\partial X}{\partial t}italic_ε start_POSTSUBSCRIPT R end_POSTSUBSCRIPT divide start_ARG ∂ italic_X end_ARG start_ARG ∂ italic_t end_ARG =uXz+MCO2ρyCO2,in(1εR)ξσeff,absent𝑢𝑋𝑧subscript𝑀subscriptCO2𝜌subscript𝑦subscriptCO2in1subscript𝜀R𝜉subscript𝜎eff\displaystyle=-u\frac{\partial X}{\partial z}+\frac{M_{\mathrm{CO_{2}}}}{\rho y% _{\mathrm{CO_{2},in}}}(1-\varepsilon_{\textrm{R}})\xi\sigma_{\textrm{eff}},= - italic_u divide start_ARG ∂ italic_X end_ARG start_ARG ∂ italic_z end_ARG + divide start_ARG italic_M start_POSTSUBSCRIPT roman_CO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ italic_y start_POSTSUBSCRIPT roman_CO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_in end_POSTSUBSCRIPT end_ARG ( 1 - italic_ε start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ) italic_ξ italic_σ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT , (12)
(ρcp)effTtsubscript𝜌subscript𝑐peff𝑇𝑡\displaystyle\left(\rho c_{\textrm{p}}\right)_{\textrm{eff}}\frac{\partial T}{% \partial t}( italic_ρ italic_c start_POSTSUBSCRIPT p end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT divide start_ARG ∂ italic_T end_ARG start_ARG ∂ italic_t end_ARG =uinρincpTz+z[ΛaxTz]+4UD(TTcool)ΔHR(1εR)ξσeff,absentsubscript𝑢insubscript𝜌insubscript𝑐p𝑇𝑧𝑧delimited-[]subscriptΛax𝑇𝑧4𝑈𝐷𝑇subscript𝑇coolΔsubscript𝐻R1subscript𝜀R𝜉subscript𝜎eff\displaystyle=-u_{\textrm{in}}\rho_{\textrm{in}}c_{\textrm{p}}\frac{\partial T% }{\partial z}+\frac{\partial}{\partial z}\left[\Lambda_{\textrm{ax}}\frac{% \partial T}{\partial z}\right]+\frac{4U}{D}(T-T_{\textrm{cool}})-\Delta H_{% \textrm{R}}(1-\varepsilon_{\textrm{R}})\xi\sigma_{\textrm{eff}},= - italic_u start_POSTSUBSCRIPT in end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT in end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT p end_POSTSUBSCRIPT divide start_ARG ∂ italic_T end_ARG start_ARG ∂ italic_z end_ARG + divide start_ARG ∂ end_ARG start_ARG ∂ italic_z end_ARG [ roman_Λ start_POSTSUBSCRIPT ax end_POSTSUBSCRIPT divide start_ARG ∂ italic_T end_ARG start_ARG ∂ italic_z end_ARG ] + divide start_ARG 4 italic_U end_ARG start_ARG italic_D end_ARG ( italic_T - italic_T start_POSTSUBSCRIPT cool end_POSTSUBSCRIPT ) - roman_Δ italic_H start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ( 1 - italic_ε start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ) italic_ξ italic_σ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ,

with initial and boundary conditions as given in [34] (L𝐿Litalic_L is the length of the reactor):

X|z=0evaluated-at𝑋𝑧0\displaystyle X|_{z=0}italic_X | start_POSTSUBSCRIPT italic_z = 0 end_POSTSUBSCRIPT =0,ΛaxdTdz|z=0=uinρincp(TTin),formulae-sequenceabsent0evaluated-atsubscriptΛax𝑑𝑇𝑑𝑧𝑧0subscript𝑢insubscript𝜌insubscript𝑐𝑝𝑇subscript𝑇in\displaystyle=0,\ \Lambda_{\textrm{ax}}\frac{dT}{dz}|_{z=0}=u_{\textrm{in}}% \rho_{\textrm{in}}c_{p}(T-T_{\textrm{in}}),= 0 , roman_Λ start_POSTSUBSCRIPT ax end_POSTSUBSCRIPT divide start_ARG italic_d italic_T end_ARG start_ARG italic_d italic_z end_ARG | start_POSTSUBSCRIPT italic_z = 0 end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT in end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT in end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_T - italic_T start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ) , (13)
X|t=0evaluated-at𝑋𝑡0\displaystyle X|_{t=0}italic_X | start_POSTSUBSCRIPT italic_t = 0 end_POSTSUBSCRIPT =X0,T|t=0=T0,Tz|Z=L=0.formulae-sequenceabsentsubscript𝑋0formulae-sequenceevaluated-at𝑇𝑡0subscript𝑇0evaluated-at𝑇𝑧𝑍𝐿0\displaystyle=X_{0},\ T|_{t=0}=T_{0},\ \frac{\partial T}{\partial z}|_{Z=L}=0.= italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_T | start_POSTSUBSCRIPT italic_t = 0 end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , divide start_ARG ∂ italic_T end_ARG start_ARG ∂ italic_z end_ARG | start_POSTSUBSCRIPT italic_Z = italic_L end_POSTSUBSCRIPT = 0 . (14)

In the formulation from (12), various parameters appear such as the reaction rate σeffsubscript𝜎eff\sigma_{\textrm{eff}}italic_σ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT, the thermal gas conductivity ΛaxsubscriptΛax\Lambda_{\textrm{ax}}roman_Λ start_POSTSUBSCRIPT ax end_POSTSUBSCRIPT, heat transfer coefficient U𝑈Uitalic_U, enthalpy of reaction Hrsubscript𝐻𝑟H_{r}italic_H start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, gas mixture density ρ𝜌\rhoitalic_ρ, and surface gas velocity u𝑢uitalic_u. Additionally, there also some constants appearing, such as the tube diameter D𝐷Ditalic_D, heat capacity cpsubscript𝑐𝑝c_{p}italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, catalyst practical fraction ξ𝜉\xiitalic_ξ. For more details, we refer the reader to [34].

In our current study, we extend the analysis from our previous work [27] and concentrate on the start-up phase of the reactor. Initially, the CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT conversion is zero throughout the reactor (denoted as X(z,0)=0𝑋𝑧00X(z,0)=0italic_X ( italic_z , 0 ) = 0 for all z𝑧zitalic_z), and the internal temperature is uniformly set to the cooling temperature, Tcoolsubscript𝑇coolT_{\textrm{cool}}italic_T start_POSTSUBSCRIPT cool end_POSTSUBSCRIPT. These initial conditions are crucial for analyzing the early dynamics of the reactor and provide a baseline for observing temporal changes. Formally, the initial conditions are

X(z,0)=0,z[0,L],T(z,0)=Tcool,z[0,L],formulae-sequence𝑋𝑧00formulae-sequencefor-all𝑧0𝐿formulae-sequence𝑇𝑧0subscript𝑇coolfor-all𝑧0𝐿X(z,0)=0,\quad\forall z\in[0,L],\ \ T(z,0)=T_{\textrm{cool}},\quad\forall z\in% [0,L],italic_X ( italic_z , 0 ) = 0 , ∀ italic_z ∈ [ 0 , italic_L ] , italic_T ( italic_z , 0 ) = italic_T start_POSTSUBSCRIPT cool end_POSTSUBSCRIPT , ∀ italic_z ∈ [ 0 , italic_L ] , (15)

where L𝐿Litalic_L is the length of the reactor. This setup allows a clear examination of the state evolution of the reactor from a consistent and well-defined starting point. In our previous study [27], we investigated the dynamics of a reactor during a sudden increase in the cooling temperature, Tcoolsubscript𝑇coolT_{\textrm{cool}}italic_T start_POSTSUBSCRIPT cool end_POSTSUBSCRIPT, from 270 to 280°C.

Using a finite volume method, we semi-discretize the PDEs describing the reactor dynamics over 200200200200 equally-spaced control volumes. Such a high spatial resolution improves the fidelity of the dynamics captured but increases the computational requirements. In terms of temporal resolution, particular attention was paid to ensuring stability and convergence of the solution. The initial and boundary conditions align with the reactor start-up scenario mentioned earlier. Notably, alongside integrating at specific time points, we concurrently evaluate the right-hand side of the model to obtain accurate time derivatives. To solve the discretized equations, we used the Kvaerno5 integrator from the diffrax 111See https://github.com/patrick-kidger/diffrax for details library in Python 3.10, chosen for its balance between accuracy and computational efficiency. Computational performance, including run-time and resource utilization, was monitored to ensure the effectiveness of the simulation approach.

Figure 1 depicts the evolution of conversion and temperature within the reactor during the start-up scenario, showcasing both the temporal and spatial domains. For the conversion variable X𝑋Xitalic_X, the results show a progressive increase along the length of the reactor, reaching over 80 % conversion towards the end. This indicates the effectiveness of the catalytic process in converting CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT. The temperature variable T𝑇Titalic_T varies between 550 K and just over 800 K during the simulation. A sharp peak in temperature is observed, representing the ignition of the reactor. This “shock-like” behavior is a critical feature of the start-up phase, demonstrating the dynamic nature of the system as it approaches a steady state. By the end of the simulation time, both the conversion and temperature profiles show convergence to a steady state, highlighting the ability of the reactor to stabilize under the given conditions.

Refer to caption
Figure 1: 3D plots for 𝖢𝖮2subscript𝖢𝖮2\textsf{CO}_{2}CO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT conversion (left) and temperature (right) over time and reactor length.

4.1 Numerical results

In our study of the dynamic changes within the reactor, we applied Principal Component Analysis (PCA) via SVD to the collected state snapshots and initiated the OpInf procedure. Figure 2 shows the decay pattern of the singular values resulting from the PCA. Note that the decay of the singular values for the conversion variable X𝑋Xitalic_X is somewhat steeper than that for the temperature variable T𝑇Titalic_T. This indicates a more diverse distribution of singular values for the temperature, suggesting its more complex and non-linear characteristics compared to the conversion process. As a result, it is more difficult to capture the nuances of the temperature dynamics, reflecting its complexity. Aiming to capture 99.90  to 99.99 % of the total variance in the data, we developed ROMs across multiple dimensions. The variance of the data is measured by the cumulative sum of the squared singular values, which represents the information richness of the system state. The number of singular values required to reach this predefined variance threshold was determined by their cumulative contribution to the total variance.

Refer to caption
Figure 2: Representation of the singular value decay and the cumulative energy encompassed by the primary dominant modes.

In constructing our low-dimensional data and its derivatives, we projected the high-dimensional data onto the dominant modes using a reduced-order basis. This projection and subsequent operator inference were performed in Python 3.10, taking advantage of the language’s computational efficiency and robust library support. Specifically, we used PyTorch’s 222see https://github.com/pytorch/pytorch for details library Adam as optimization algorithm combined with a CyclicLR scheduler for a cyclic learning rate policy. The scheduler, operating in a triangular2 mode with no cycle dynamics, adjusted the learning rate between 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT and 0.50.50.50.5. The optimization was designed to stop early after 500 epochs if minimal improvement was observed, thereby increasing efficiency. A regularization factor of 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT was applied to the quadratic matrix for stability and to prevent over-fitting. The reduced operators 𝐀^^𝐀\hat{\mathbf{A}}over^ start_ARG bold_A end_ARG, 𝐇^^𝐇\hat{\mathbf{H}}over^ start_ARG bold_H end_ARG, and 𝐁^^𝐁\hat{\mathbf{B}}over^ start_ARG bold_B end_ARG were derived following these methods, in accordance with the stability parameterization suggested by [13].

Refer to caption
Figure 3: Flattened 3D representation comparing the true response with the inferred model, highlighting deviations in both conversion and temperature.

Figure 3 shows a flattened 3D plot comparing the actual data with our ROM, which captures 99.90 % of the energy for both conversion and temperature. The model, at a rank of r=7𝑟7r=7italic_r = 7 (rX=2subscript𝑟X2r_{\mathrm{X}}=2italic_r start_POSTSUBSCRIPT roman_X end_POSTSUBSCRIPT = 2 and rT=5subscript𝑟T5r_{\mathrm{T}}=5italic_r start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT = 5), closely matches the actual data. However, discrepancies are observed in the regions representing the ignition phase, a near-shock scenario. Accurately capturing such abrupt and nonlinear transitions is a well-known challenge in modeling, often due to the inherent limitations of linear and quadratic terms in representing extreme state changes. These observations are critical for further refinement of the model, particularly in improving its ability to more accurately represent rapid dynamic changes. Quantitatively, the Frobenius norm shows a small deviation of 0.45 % from the original model, underscoring the accuracy of the ROM. The computational efficiency of the ROM is quite high, in that it needs only 0.46 % of the time required by the full mechanistic model.

5 Conclusion and outlook into the future

OpInf combines the advantages of physics-based modeling with optimization and learning techniques by integrating first-principles models with data-driven regression. The results reported in this work have successfully demonstrated the ability of the OpInf ROMs to capture complex system dynamics in process engineering with increased computational efficiency. As an outlook, one option would be to allow variable parameters, by including variable input loads that affect species volume flow. In addition, the incorporation of specific control terms into the model will enhance its adaptability. By refining the tuning capabilities of the fitted models across a wide range of parameters, control settings, and many-query environments, they will hopefully be integrated into an effective digital twin environment.

Acknowledgement

This work is partially funded by the Bundesministerium für Bildung und Forschung (BMBF) and Project Management Jülich (PtJ) under grant 03HY302R. This work is also part of the research initiative “SmartProSys: Intelligent Process Systems for the Sustainable Production of Chemicals” funded by the Ministry for Science, Energy, Climate Protection and the Environment of the State of Saxony-Anhalt.

References

  • [1] Antoulas, A.C.: Approximation of Large-scale Dynamical Systems. Advances in Design and Control. SIAM, Philadelphia, PA, USA (2005). 10.1137/1.9780898718713
  • [2] Benner, P., Breiten, T.: Two-sided projection methods for nonlinear model order reduction. SIAM J. Sci. Comput. 37(2), B239–B260 (2015). 10.1137/14097255X
  • [3] Benner, P., Goyal, P., Heiland, J., Pontes Duff, I.: Operator Inference and Physics-Informed Learning of Low-Dimensional Models for Incompressible Flows. Electron. Trans. Numer. Anal. 56, 28–51 (2022). 10.1553/etna_vol56s28
  • [4] Benner, P., Goyal, P., Kramer, B., Peherstorfer, B., Willcox, K.: Operator Inference for Non-Intrusive Model Reduction of Systems with Non-Polynomial Nonlinear Terms. Comp. Meth. Appl. Mech. Eng. 372 (2020). 10.1016/j.cma.2020.113433
  • [5] Benner, P., Gugercin, S., Willcox, K.: A survey of projection-based model reduction methods for parametric dynamical systems. SIAM review 57(4), 483–531 (2015). 10.1137/130932715
  • [6] Benner, P., Ohlberger, M., Cohen, A., Willcox, K.: Model Reduction and Approximation. Computational Science &\&& Engineering. Society for Industrial and Applied Mathematics, Philadelphia, PA (2017). 10.1137/1.9781611974829
  • [7] Bremer, J., Rätze, K.H., Sundmacher, K.: Co22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation: Optimal start-up control of a fixed-bed reactor for power-to-gas applications. AIChE 63(1), 23–31 (2017). 10.1002/aic.15496
  • [8] Brunton, S.L., Proctor, J.L., Kutz, J.N.: Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proceedings of the National Academy of Sciences 113(15), 3932–3937 (2016). 10.1073/pnas.1517384113
  • [9] Filanova, Y., Duff, I.P., Goyal, P., Benner, P.: An operator inference oriented approach for linear mechanical systems. Mechanical Systems and Signal Processing 200, 110620 (2023). 10.1016/j.ymssp.2023.110620
  • [10] Ghaib, K., Ben-Fares, F.Z.: Power-to-methane: A state-of-the-art review. Renewable and Sustainable Energy Reviews 81, 433–446 (2018). 10.1016/j.rser.2017.08.004
  • [11] Goyal, P., Benner, P.: LQResNet: a deep neural network architecture for learning dynamic processes. arXiv preprint (2021). 10.48550/arXiv.2103.02249
  • [12] Goyal, P., Benner, P.: Discovery of nonlinear dynamical systems using a Runge–Kutta inspired dictionary-based sparse regression approach. Proceedings of the Royal Society A 478(2262), 20210883 (2022). 10.1098/rspa.2021.0883
  • [13] Goyal, P., Duff, I.P., Benner, P.: Guaranteed stable quadratic models and their applications in SINDy and operator inference. arXiv preprint (2023). 10.48550/arXiv.2308.13819
  • [14] Gu, C.: QLMOR: A projection-based nonlinear model order reduction approach using quadratic-linear representation of nonlinear systems. Trans. Comput.-Aided Design Integr. Circ. Syst. 30(9), 1307–1320 (2011). 10.1109/TCAD.2011.2142184
  • [15] Hartmann, D., Herz, M., Wever, U.: Model order reduction a key technology for digital twins. Reduced-Order Modeling (ROM) for Simulation and Optimization: Powerful Algorithms as Key Enablers for Scientific Computing pp. 167–179 (2018)
  • [16] Holmes, P., Lumley, J.L., Berkooz, G.: Turbulence, coherent structures, dynamical systems and symmetry. Cambridge Monographs on Mechanics. Cambridge University Press (2012). 10.1017/CBO9780511622700
  • [17] Kramer, B., Peherstorfer, B., Willcox, K.E.: Learning nonlinear reduced models from data with operator inference. Annual Review of Fluid Mechanics 56 (2024). 10.1146/annurev-fluid-121021-025220
  • [18] Kutz, J.N., Brunton, S.L., Brunton, B.W., Proctor, J.L.: Dynamic Mode Decomposition: Data-Driven Modeling of Complex Systems. SIAM (2016). 10.1137/1.9781611974508
  • [19] Ljung, L.: System Identification: Theory for the User. Information and System Sciences Series. Prentice Hall (1999). Subsequent edition, ISBN: 978-0136566953
  • [20] Lu, L., **, P., Pang, G., Zhang, Z., Karniadakis, G.E.: Learning nonlinear operators via DeepONet based on the universal approximation theorem of operators. Nature Machine Intelligence 3(3), 218–229 (2021). 10.1038/s42256-021-00302-5
  • [21] Mayo, A.J., Antoulas, A.C.: A framework for the solution of the generalized realization problem. Linear Algebra and Its Applications 425(2-3), 634–662 (2007). 10.1016/j.laa.2007.03.008
  • [22] McQuarrie, S.A., Huang, C., Willcox, K.E.: Data-driven reduced-order models via regularised operator inference for a single-injector combustion process. Journal of the Royal Society of New Zealand 51(2), 194–211 (2021). 10.1080/03036758.2020.1863237
  • [23] McQuarrie, S.A., Khodabakhshi, P., Willcox, K.E.: Nonintrusive reduced-order models for parametric partial differential equations via data-driven operator inference. SIAM Journal on Scientific Computing 45(4), A1917–A1946 (2023). 10.1137/21M1452810
  • [24] Niederer, S.A., Sacks, M.S., Girolami, M., Willcox, K.: Scaling digital twins from the artisanal to the industrial. Nature Computational Science 1(5), 313–320 (2021). 10.1038/s43588-021-00072-5
  • [25] Peherstorfer, B.: Sampling low-dimensional Markovian dynamics for preasymptotically recovering reduced models from data with operator inference. SIAM Journal on Scientific Computing 42(5), A3489–A3515 (2020). 10.1137/19M1292448
  • [26] Peherstorfer, B., Willcox, K.: Data-driven operator inference for nonintrusive projection-based model reduction. Comp. Meth. Appl. Mech. Eng. 306, 196–215 (2016). 10.1016/j.cma.2016.03.025
  • [27] Peterson, L., Goyal, P., Gosea, I.V., Bremer, J., Benner, P., Sundamacher, K.: Learning reduced-order models for dynamic CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT methanation using operator inference. Tech. rep. (2023). Accepted for presentation at the Proceedings of the 34th European Symposium on Computer Aided Process Engineering, June 2-6, 2024, Florence, Italy
  • [28] Qian, E., Krämer, B., Peherstorfer, B., Willcox, K.: Lift & learn: Physics-informed machine learning for large-scale nonlinear dynamical systems. Physica D: Nonlinear Phenomena 406(1), art. 132401 (2020). 10.1016/j.physd.2020.132401
  • [29] Raissi, M., Perdikaris, P., Karniadakis, G.E.: Physics-informed neural networks: A deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations. J. Comp. Phys. 378, 686–707 (2019). 10.1016/j.jcp.2018.10.045
  • [30] Sawant, N., Kramer, B., Peherstorfer, B.: Physics-informed regularization and structure preservation for learning stable reduced models from data with operator inference. Comp. Meth. Appl. Mech. Eng. 404, 115836 (2023). 10.1016/j.cma.2022.115836
  • [31] Uy, W.I.T., Wang, Y., Wen, Y., Peherstorfer, B.: Active operator inference for learning low-dimensional dynamical-system models from noisy data. SIAM J. Sci. Comput. 45(4), A1462–A1490 (2023). 10.1137/21M1439729
  • [32] van Overschee, P., de Moor, B.L.: Subspace Identification for Linear Systems: Theory - Implementation - Applications. Kluwer Academic Publishers (1996). 10.1007/978-1-4613-0465-4
  • [33] Yıldız, S., Goyal, P., Benner, P., Karasözen, B.: Learning reduced-order dynamics for parametrized shallow water equations from data. Int. J. Num. Methods Fluids 93(8), 2803–2821 (2021). 10.1002/fld.4998
  • [34] Zimmermann, R.T., Bremer, J., Sundmacher, K.: Load-flexible fixed-bed reactors by multi-period design optimization. Chemical Engineering Journal 428, 130771 (2022). 10.1016/j.cej.2021.130771