Reduced Order Modeling of PDEs
Reduced Order Modeling of PDEs
Zulkeefal Dar
International Center for Numerical Methods in Engineering, e-mail: [email protected]
Joan Baiges
Universitat Politècnica de Catalunya e-mail: [email protected]
Ramon Codina
Universitat Politècnica de Catalunya e-mail: [email protected]
1
Contents
3
4 Contents
List of Acronyms
AE AutoEncoder
AMR Adaptive Mesh-Refinement
ANN Artificial Neural Network
BiLSTM Bidirectional Long Short-Term Memory
DEIM Discrete Empirical Interpolation Method
DMD Dynamic Mode Decomposition
FE Finite Element
FNN Feedforward Neural Network
FOM Full Order Model
LSTM Long Short-Term Memory
ML Machine Learning
NIROM Non-Intrusive Reduced Order Model
NN Neural Network
PDE Partial Differential Equation
PG Petrov-Galerkin
POD Proper Orthogonal Decomposition
POD-G Proper Orthogonal Decomposition based Galerkin projection
PROM Parametric Reduced Order Model
RIC Relative Information Content
ROM Reduced Order Model
SGS Sub-Grid Scales
SINDy Sparse Identification of Nonlinear Dynamics
SUPG Streamline-Upwind Petrov–Galerkin
SVD Singular Value Decomposition
VMS Variational MultiScale
Abstract This chapter presents an overview of the most popular reduced order
models found in the approximation of partial differential equations and their con-
nection with machine learning techniques. Although the presentation is applicable
to many problems in science and engineering, the focus is first order evolution prob-
lems in time and, more specifically, flow problems. Particular emphasis is put in the
distinction between intrusive models, that make use of the physical problem being
modeled, and non-intrusive models, purely designed from data using machine learn-
ing strategies. For the former, models based on proper orthogonal decomposition and
Galerkin projection are described in detail, whereas alternatives are only mentioned.
Likewise, some modifications that are crucial in the applications are detailed. The
progressive incorporation of machine learning methods is described, yielding first
hybrid formulations and ending with pure data-driven approaches. An effort has
been made to include references with applications of the methods being described.
Contents 5
1 Introduction
In the current data-centric era, Machine Learning (ML) has emerged as a viable
tool for reduced order modeling. ML has revolutionized a wide range of fields over
the past few decades. Scientific computing, and in particular reduced order modeling,
is no exception. Although the interest in exploring ML techniques for reduced order
modeling is relatively new, it has already shown great potential by replacing in part,
or entirely, the offline and online steps.
A variety of conventional (see Remark 2) and ML-based techniques have been
developed for offline and online phases till date and applied successfully in a variety
of contexts, e.g., solid mechanics [53, 147], material science [80], fluid mechanics
[15, 16, 32, 64, 68, 87, 98, 114, 130, 141], shape optimization [5, 25, 94, 93, 121]
and flow control [11, 70, 103, 111] problems. The proper orthogonal decomposition
based Galerkin projection (POD-G) method can be considered to be the most well-
established and commonly used method for reduced order modeling. This method
uses proper orthogonal decomposition (POD) to find the basis, called POD modes,
of the low-dimensional space. The FOM is then projected in an intrusive manner
onto these POD modes using mostly Galerkin projection to solve for the unknowns.
The organization of the chapter is as follows. Section 2 describes POD along with
its essential ingredient, the singular value decomposition (SVD). Section 3 describes
the Galerkin projection, hyperreduction, and stabilization of POD-ROMs.Section4
describes the non-intrusive ROMs with a brief explanation of dynamic mode de-
composition (DMD).Section 5 deals with the description of parametric ROMs. Fi-
nally,Section 6 describes the ML techniques used for the online and offline phases
of reduced order modeling.
Remark 1 In literature, the term reduced basis is also sometimes used for projection
based methods. However, more commonly reduced basis methods are meant to refer
to a particular class of projection based methods based on greedy algorithms1. This
later usage is also applicable in the context of this chapter.
The basis commonly used, e.g., piecewise-linear basis in the finite element (FE)
method, Fourier modes, etc., can solve a large number of dynamical systems, but
1 Greedy algorithms are a class of algorithm which are based on choosing the option which produces
the largest immediate reward with the expectation that the successive application of greedy sampling
will lead to a global optimum. Greedy algorithms may use an error estimator to guide the sampling.
Contents 7
these basis are generic and do not correspond to the intrinsic modes of the systems
which they solve. Hence, a large number of such basis functions need to be used
to capture the solution. The intrinsic modes which form the basis of the solution
space can be found using proper generalized decomposition, reduced basis method
or proper orthogonal decomposition (POD), among others. Proper generalized de-
composition and reduced basis methods are commonly based on a greedy approach
and a comprehensive review on them can be found in [42, 43] and [76], respectively.
POD [37] is perhaps the most commonly used method to find the basis, called POD
modes in the context of POD.
For clarity, let us first introduce the concept of function and vector based descrip-
tion of a variable in the context of numerical methods. Suppose that the analysis
domain is spatially discretized using an interpolation based method like finite ele-
ments, finite volumes, spectral elements, etc. The variable of interest can then either
be represented in the vectorial form as the collection of coefficients that multiply
the basis functions, or in a corresponding functional form which relies on the inter-
polation to define the variable over the entire domain. Throughout this chapter, the
functional representation is denoted using the lowercase letters, like 𝒒, and the vecto-
rial representation using the uppercase letters, like 𝑸. In the case of Greek alphabets,
where the case of alphabets is not obvious, functional representation is denoted by
showing the dependence on the spatial coordinates 𝒙 explicitly, like 𝜻 (𝒙). Also, a
variable with underbar 𝒂 represents the variable in a general form which could be
either functional or vectorial based on the context. In the case of the FE method, the
vectorial and functional representations of a variable 𝒖 are related as
𝑛𝑛
∑︁
𝒖(𝒙, 𝑡) = 𝝌 𝑘 (𝒙)𝑈 𝑘 (𝑡)
𝑘=1
where 𝒖(𝒙, 𝑡) is the functional representation which depends on the spatial coordi-
nates 𝒙 in addition to time 𝑡, 𝑈 𝑘 the 𝑘−th element of the vector 𝑼, 𝝌 𝑘 (𝒙) the FE
interpolation function for the 𝑘−th node and 𝑛𝑛 the total number of nodes.
Now, POD consists of representing a variable 𝒖 as a linear combination of the
POD modes as
𝑛𝑏
∑︁
𝒖(𝑡, 𝜇) = 𝚿 𝑘 (𝜇)𝑈𝑟𝑘 (𝑡, 𝜇) (1)
𝑘=1
where 𝚿𝑘 is the 𝑘-th POD mode, 𝑈𝑟𝑘the 𝑘-th ROM coefficient, 𝑛𝑏 the number of
basis vectors and 𝜇 a parameter that characterizes the behavior of the system. 𝒖
and 𝚿 could be the functional or vectorial representation, as required, but the same
representation must be used for both variables.
POD relies on a singular value decomposition (SVD) to find the basis. Let us
describe how the basis is determined using the SVD of the data generated using
PDEs.
8 Contents
Let us consider a general unsteady nonlinear PDE describing the behavior of a real-
valued function 𝒖 of 𝑛 components and dependent on a parameter 𝜇. The evolution
of 𝒖 in the spatial domain Ω ⊂ R𝑑 , 𝑑 denoting the dimensions of the problem, and
time interval ]0, 𝑡f [ is given by
where N is a nonlinear operator, 𝒇 the forcing term and 𝜕𝑡 the time derivative.
Equation (2) is further provided with suitable boundary and initial conditions so that
the problem is well-posed. For simplicity, 𝜇 is considered to be fixed for now, and
hence, 𝒖 will not be stated explicitly as a function of the parameter 𝜇 from now on
tillSection 5, when parametric ROMs are discussed.
After the advent of the computational era, the most commonly used technique
to solve (2) is to discretize it in space using a discretization technique, e.g., the FE
method. This discretization leads to the a system of ordinary differential equations
(ODEs) which reads: find 𝑼 :]0, 𝑡f [→ R𝑛 𝑝 such that
𝑨(𝑼)𝑼 = 𝑹, (4)
where 𝑨(𝑼) ∈ R𝑛 𝑝×𝑛 𝑝 is the nonlinear system matrix and 𝑹 ∈ R𝑛 𝑝 is the right-
hand-side which takes into account the contributions of the previous values of 𝑼 as
well. Equation (4) can then be solved for nt time-steps to get nt solution vectors.
To find the POD basis, all solution vectors are not generally required. Rather
it is desired to select a minimum, but sufficient, number of solution vectors that
contain all the important dynamic features of the system. A simple approach for a
uniform step time-integration scheme is to use solution vectors after every 𝑖-th time-
step, where 𝑖 is a natural number [17]. Another approach could be to capture each
cycle of a periodic phenomenon using a certain number of solution vectors [107].
Suppose, that we are able to gather a set of ns solution vectors, called snapshots,
{𝑼 1 , 𝑼 2 , ..., 𝑼 𝑛𝑠 } carrying all the important features of the system dynamics. To
simplify the exposition, we have assumed that the snapshots correspond to the first
𝑛𝑠 consecutive solution vectors, but this can be easily generalized. Note that the
term snapshots will be interchangeably used for the solution vectors, as well as
Contents 9
their mean-subtracted form discussed in Section 2.2. Once the solution set has been
gathered, the SVD is used to find the basis or POD modes.
SVD, also known as principal component analysis, is one of the most important
matrix factorization techniques used across many fields. The SVD of a matrix is
guaranteed to exist and can be considered unique for the basis generation purposes.
To perform the SVD, it is customary to first subtract the mean value 𝑼 ∈ R𝑛 𝑝 from the
solution vectors 𝑼 𝑗 to obtain 𝑺 𝑗 = 𝑼 𝑗 − 𝑼, for 𝑗 = 1, 2, ..., 𝑛𝑠. The mean-subtracted
snapshots are then arranged into a matrix 𝑺 ∈ R𝑛 𝑝×𝑛𝑠 as follows:
1 2
𝑺 = 𝑺 𝑺 . . . 𝑺 ,
𝑛𝑠
a tall skinny matrix as, in general, 𝑛𝑠 ≪ 𝑛𝑝. It is now desired to find the basis of the
space B ⊂ B 𝑓 to which these snapshots belong using SVD. The SVD of 𝑺 gives
𝑺 = 𝚿 ′ 𝚲′ 𝑽 𝑇 , (5)
where 𝚿′ ∈ R𝑛 𝑝×𝑛 𝑝 contains the left singular vectors, 𝚲′ ∈ R𝑛 𝑝×𝑛𝑠 contains the
singular values and 𝑽 ∈ R𝑛𝑠×𝑛𝑠 contains the right singular vectors. Some important
properties of the SVD are listed below:
• Matrices 𝚿′ and 𝑽 are orthogonal i.e.
Now, 𝚲′ has
at most 𝑛𝑠 non-zero values. Assuming such a case, it is possible to
′ 𝚲
write 𝚲 = , where 𝚲 ∈ R𝑛𝑠×𝑛𝑠 . Using this, the full SVD (5) can be converted to
0
its economy or reduced SVD form as
𝚲 𝑇
𝑺 = 𝚿′ 𝑽 = 𝚿𝚲𝑽 𝑇 , (8)
0
where 𝚿 ∈ R𝑛 𝑝×𝑛𝑠 . The full and reduced versions of SVD are shown in Fig. 1. Note
that (8) represents the exact decomposition of 𝑺. The set of columns {𝚿 𝑗 } 𝑛𝑠
𝑗=1 of the
10 Contents
So, the dimension of the solution space has been reduced from 𝑛𝑝 to 𝑛𝑠, with
𝑛𝑠 ≪ 𝑛𝑝. However, 𝑛𝑠 still could be of the order of hundreds or even thousands
and could still be considered computationally demanding. So, to unlock the full
potential of ROMs in terms of computational savings, truncation is performed to
yield a smaller number of basis vectors than the one provided by the economy SVD.
…
…
= Ψ Ψ┴ =
ns x ns ns x ns ns x ns
np x ns np x np np x ns np x ns
Fig. 1 Matrix representation of full and economy SVD. The lightening of the color of the circles
represents the ordered decreasing of the diagonal values of 𝚲′ and 𝚲.
The singular and RIC values for the classical problem of the flow over a cylinder
approximated using the FE method are shown in Fig. 2. It can be seen that only a few
initial values contain the most of the energy. So, instead of using 150 basis vectors,
it is possible to use just a few of them to describe the flow dynamics around the
cylinder. Based on the reasoning discussed above, RIC is widely used as a truncation
criterion. The number of POD modes can then be decided such that RIC is equal
to a desired value, e.g., 0.9, meaning that the POD modes will retain 90% of the
Contents 11
𝑺 ≈ 𝑺ˆ = 𝚿
ˆ𝚲ˆ 𝑽ˆ 𝑇 , (9)
Thus, the presence of patterns in the high dimensional data, shown by rapidly
decreasing singular values, and the optimality of SVD guaranteed by the Eckart–
Young theorem have resulted in the wide usage of POD to find the basis of the
reduced spaces.
0
10
0.6
RIC
0.4
10-2
0.2
10-4
0
0 20 40 60 80 100 120 140
no. of POD modes
Many times we are dealing with functions, e.g., when using FE methods, defined
over the entire domain Ω, and not vectors corresponding to the degrees of freedom
12 Contents
of the approximation. In such cases, it could be desired for the properties to hold in
the functional (continuous) sense rather than in the algebraic (discrete) sense. So,
let us suppose that instead of solving the minimization problem (10), it is desired to
minimize its functional counterpart
2
𝑛𝑠
∑︁ 𝑟 ∫
∑︁
𝑗 ( 𝜙ˆ1 (𝒙), ..., 𝜙ˆ𝑟 (𝒙)) = 𝑠𝑖 (𝒙) − 𝑠𝑖 (𝒙) 𝜙ˆ 𝑗 (𝒙) 𝜙ˆ 𝑗 (𝒙) ,
𝑖=1 𝑗=1 Ω
𝐿 2 (Ω)
∫
subject to 𝜙ˆ𝑖 (𝒙) 𝜙ˆ 𝑗 (𝒙) = 𝛿𝑖 𝑗 , (11)
Ω
where 𝑠𝑖 (𝒙) is the functional form of the vectors 𝑺𝑖 , 𝑖 = 1, ..., 𝑛𝑠, and 𝜙ˆ 𝑗 (𝒙),
𝑗 = 1, ..., 𝑟, are the basis functions, which are 𝐿 2 (Ω)-orthogonal. Note that (11)
minimizes the difference over the entire domain Ω. For the sake of clarity, we have
considered that the unknown of the problem is a scalar function, and so are the
snapshots and the basis, but the extension to the vector case is straightforward. Also
note that the difference between 𝚿ˆ and 𝜙(𝒙)
ˆ is not only of the vectorial and functional
ˆ ˆ
representation. Rather 𝚿 and 𝜙(𝒙) are two different bases having different orthogonal
properties. The functional SVD (11) can be shown to be the same as minimizing
2
𝑛𝑠
∑︁ 𝑟
∑︁
ˆ 1 , ..., 𝚽
ˆ 𝑟) = 1/2 𝑖 1/2 𝑖 𝑇 1/2 ˆ𝑗 𝑴 1/2 ˆ𝑗
𝐽 (𝚽 𝑴 𝑺 − (𝑴 𝑺) 𝑴 𝚽 𝚽 ,
𝑖=1 𝑗=1
R𝑛 𝑝
𝑇
ˆ 𝑴𝚽
subject to 𝚽 ˆ = 𝑰 𝑟 , (12)
where 𝑴 is the mass-matrix as in (3) and 𝑰 𝑟 ∈ R𝑟 ×𝑟 is the identity matrix. Note that
ˆ
the 𝐿 2 (Ω)-orthogonality of basis functions 𝝓(𝒙) translates to orthogonality of the
corresponding basis vectors 𝚽 ˆ with respect to the mass-matrix 𝑴 as 𝚽 ˆ 𝑇 𝑴𝚽ˆ = 𝑰𝑟 .
ˆ we perform the SVD of 𝑺 = 𝑴 𝑺:
To find 𝚽, e 1/2
𝑺=𝚽
e e𝚲ˆ 𝑽ˆ 𝑇 ,
Using the functional SVD produced more accurate results in [54]. Note, however,
that throughout this chapter the SVD term will refer to the one that solves problem
(10), unless stated otherwise, e.g., in Section 3.3.3.
Contents 13
ˆ forms the basis of the reduced solution space B𝑟 , and was calculated from
As 𝚿
mean-subtracted snapshots, decomposition (1) can be written as
ˆ 𝑟 + 𝑼,
𝑼 ≈ 𝚿𝑼 (13)
ˆ 𝑇 𝑨𝚿𝑼
𝚿 ˆ 𝑇 ( 𝑹 − 𝑨𝑼).
ˆ 𝑟 =𝚿 (15)
which is the Galerkin projection of the full order system (4) onto the reduced space.
Let us write (15) compactly as
𝑨𝑟 𝑼 𝑟 = 𝑹 𝑟 (16)
where
𝑇
ˆ 𝑨𝚿
𝑨𝑟 := 𝚿 ˆ ∈ R𝑟 ×𝑟 ,
ˆ 𝑇 ( 𝑹 − 𝑨𝑼)
𝑹𝑟 := 𝚿 ∈ R𝑟 .
Applicable for the general matrices 𝑨, the so-called Petrov-Galerkin (PG) pro-
jection is found to provide more stable results, as compared to Galerkin projection,
in the case of 𝑨 not being a SPD matrix [35]. Using 𝚿 ˆ 𝑇 𝑨 as a suitable PG projector,
14 Contents
𝑨𝑟 𝐴𝑼𝑟 = 𝑹𝑟 𝐴, (17)
where now
𝑇
ˆ 𝑨𝑇 𝑨 𝚿
𝑨𝑟 𝐴 := 𝚿 ˆ ∈ R𝑟 ×𝑟 ,
ˆ 𝑇 𝑨𝑇 ( 𝑹 − 𝑨𝑼)
𝑹𝑟 𝐴 := 𝚿 ∈ R𝑟 .
This corresponds to a least squares strategy for solving (14) with respect to the
standard Euclidean norm in R𝑛 𝑝 . Irrespective of the type of projection used, both
the final reduced order systems (16) and (17) are 𝑟 × 𝑟 systems as opposed to the
full order system (4) of size 𝑛𝑝 × 𝑛𝑝, with 𝑟 ≪ 𝑛𝑝. Thus the reduced order system
can be solved at a fraction of the cost of the full order system. All the concepts
described later apply to both the Galerkin and PG ROMs; however for simplicity, we
will describe them using the Galerkin-ROM (15).
3.2 Hyperreduction
The ROM discussed above can be solved at a reduced computational expense. How-
ever, assembling the system matrices has a cost of the same order as that of the FOM.
For linear problems, the assembling of matrices needs to be done once, and hence, is
not considered a bottle-neck to achieving reduced computation times. However, for
nonlinear problems the system matrices need to be assembled for every nonlinear it-
eration, i.e., multiple times for every time-step in general, and will lead to a significant
cost. Thus, it is important to use some techniques to determine the nonlinear terms
at a reduced cost. This is achieved using hyperreduction techniques and the resulting
models are called hyper-ROMs. There are many methods used for hyperreduction
including, but not limited to, empirical interpolation method [23] or its discrete
version discrete empirical interpolation method (DEIM) [38], Gauss-Newton with
approximate tensors [35], missing point estimator approach [12], cubature based
approximation method [9], energy conserving sampling and weighting method [61]
and adaptive mesh-refinement (AMR) based hyperreduction [118]. Here we briefly
describe DEIM and AMR based hyperreduction.
DEIM is a greedy algorithm and its origin can be traced back to the gappy POD
method [59] which was originally designed for image reconstruction. Just as ROM
approximates the solution space by a subspace, DEIM does the same but for non-
linear terms only. However, DEIM uses interpolation indices to find the temporal
coefficients instead of solving the reduced system.
Let us denote the vector of nonlinear terms as 𝑵(𝜃) ∈ R𝑛 𝑝 , depending on 𝜃. 𝜃 can
represent time 𝑡 or any other parameter in the case of parametric ROMs. However,
here we explain DEIM in the context of non-parametric nonlinear ROMs with 𝜃 = 𝑡.
For DEIM applied to parametric ROMs, see [10]. DEIM proposes approximating
the space to which the 𝑵 belongs by a subspace of lower dimension 𝑠, i.e., 𝑠 ≪ 𝑛𝑝
and not necessarily equal to the dimension 𝑟 of the ROM space. Let this subspace
be spanned by the basis 𝑩 = [𝑩1 , ..., 𝑩 𝑠 ] ∈ R𝑛 𝑝×𝑠 . Thus we can write
where 𝒅(𝑡) is the vector of coefficients. For simplicity, from now on the dependence
on 𝑡 will be omitted from the notation.
An efficient way to determine 𝒅 is to sample 𝑠 spatial points and use them to
determine 𝒅. This can be achieved using a sampling matrix 𝑯 defined as
𝑯 = [𝑯 𝑠1 , ..., 𝑯 𝑠𝑠 ] ∈ R𝑛 𝑝×𝑛𝑠
where 𝑯 𝑠 𝑗 = [0, ..., 0, 1, 0, ...0] 𝑇 ∈ R𝑛 𝑝 , for 𝑗 = 1, ..., 𝑠, is the 𝑠 𝑗 -th column of the
identity matrix 𝑰 𝑛 𝑝 ∈ R𝑛 𝑝×𝑛 𝑝 . Using the sampling matrix 𝑯 we can write
𝑯𝑇 𝑵 = 𝑯𝑇 𝑩𝒅.
𝒅 = (𝑯𝑇 𝑩) −1 𝑯𝑇 𝑵. (19)
𝑵 ≈ 𝑩(𝑯𝑇 𝑩) −1 𝑯𝑇 𝑵. (20)
Now we need to define the basis 𝑩 and the sampling points 𝑠 𝑗 , 𝑗 = 1, ..., 𝑠, called
interpolation indices in DEIM, to approximate 𝑵 using (20) at a reduced cost. The
basis 𝑩 is found using POD for the nonlinear vector 𝑵. During the simulations, the
nonlinear vectors at different time-steps are gathered to form a snapshot matrix 𝑺 𝑁
of nonlinear terms as
1 2
𝑺 𝑁 = 𝑵 𝑵 . . . 𝑵 .
𝑛𝑠
The truncated SVD of rank-𝑠 is then performed as
16 Contents
𝑺ˆ 𝑁 = 𝑩𝚲 𝑁 𝑽 𝑇𝑁 (21)
to obtain the basis 𝑩 of the desired order. The interpolation indices are then selected
iteratively using the basis 𝑩. This approach is shown in Algorithm 1, where [|𝜌| 𝑦]
= max{| 𝑿 |} means that 𝑦 is the index of the maximum value of the components of
vector 𝑿 = [𝑋1 , ..., 𝑋𝑛 𝑝 ], i.e., 𝑋 𝑦 ≥ 𝑋𝑧 , for 𝑧 = 1, ..., 𝑛𝑝. The smallest 𝑦 is taken if
more than one component corresponds to the maximum value.
Instabilities can arise when PDEs are solved using numerical methods, usually in
singular perturbation problems or when the approximation spaces of the different
unknowns need to satisfy compatibility conditions. This issue is further exacerbated
when POD-G is used to develop a ROM. This has to do with the fact that the ROM
Contents 17
does not account for the impact of the FOM scales that are not captured by the
low-order space. This problem is well-known in other computational mechanics set-
tings, such as finite elements, where stabilized formulations have been developed to
address the instability of the Galerkin projection. The Variational Multiscale (VMS)
framework, originally proposed in [79], is a popular framework used to develop
stabilized formulations taking into account the effect of the discarded scales in a
multi-scale problem. A comprehensive review of VMS-based stabilization meth-
ods developed for fluid problems is provided in [51]. Inspired by this, VMS based
stabilization methods have been developed for projection based ROMs [118] and
successfully applied in the context of flow problems [119], fluid-structure interac-
tion [133, 134] and adaptive-mesh based hyperreduction [118]. A comprehensive
description of it is provided in [49, 118]. However, a summary of the method, which
uses the same VMS formulation to stabilize both FOM and ROM, is presented here
for completeness. Let us describe the formulation using a general unsteady nonlinear
convection-diffusion-reaction transport equation.
Let us consider again problem (2) and write it in a slightly modified form, along
with the boundary and initial conditions. Let the boundary Γ of the domain Ω be
split into non-overlapping Dirichlet, Γ𝐷 , and Neumann, Γ𝑁 , parts. Given the initial
condition for the unknown 𝒖 0 , the problem aims at finding 𝒖 of 𝑛 components that
satisfies
𝜕𝑡 𝒖 + N (𝒖; 𝒖) = 𝒇 in Ω, 𝑡 ∈ ]0, 𝑡f [,
D𝒖 = D𝒖 0 on Γ𝐷 , 𝑡 ∈ ]0, 𝑡f [,
F (𝒖; 𝒖) = 𝒇 𝑁 on Γ𝑁 , 𝑡 ∈ ]0, 𝑡f [,
0
𝒖=𝒖 in Ω, 𝑡 = 0,
where 𝒏 is the external unit normal to the boundary Γ with 𝑛𝑖 being its 𝑖-th component.
To write the weak form of the problem, let the integral of the product of two
functions, 𝒇 and 𝒈 over the domain 𝜔 be defined by ⟨ 𝒇 , 𝒈⟩ 𝜔 . For simplicity the
18 Contents
subscript 𝜔 is omitted in the case 𝜔 = Ω. Let us also introduce the form 𝐵 𝜔 and the
linear form 𝐿 𝜔 as
Let 𝒖(., 𝑡) and 𝒗 belong to the space B𝑐 , the solution space of the continuous problem.
The weak form of the problem (in space) consists of finding 𝒖 :]0, 𝑡f [→ B𝑐 such
that
for all 𝒗 ∈ B𝑐,0 , where B𝑐,0 is the space of time independent test functions that
satisfy D𝒗 = 0 on Γ𝐷 . For simplicity, we assume in what follows homogeneous
Dirichlet conditions so that 𝒗 ∈ B𝑐 = B𝑐,0 .
VMS method can be applied to other discretization techniques, but in what follows
we shall concentrate on the FE method. Thus, let us discretize the domain using
FEs. Let Pℎ = {𝐾 } be a FE partition of the domain Ω, assumed quasi-uniform for
simplicity, with elements of size ℎ. From this, a conforming FE space Bℎ ⊂ B𝑐 may
be constructed using a standard approach. Note that now Bℎ = B 𝑓 , i.e., the FE space
is a particular realization of the FOM space introduced earlier.
Any time integration scheme may be used for the time discretization. For con-
ciseness, we shall assume that a backward difference scheme is employed with a
uniform time step Δ𝑡 and the time discretization is represented by replacing 𝜕𝑡 with
𝛿𝑡 , where 𝛿𝑡 involves 𝒖 𝑛ℎ , 𝒖 𝑛−1
ℎ , ..., depending on the order of the scheme used. Us-
ing a superscript 𝑛 for the time step counter, and a subscript ℎ for FE quantities, the
fully discretized Galerkin approximation of problem (22) is to find {𝒖 𝑛ℎ } ∈ Bℎ , for
𝑛 = 1, ..., 𝑛𝑡, 𝑛𝑡 being the number of time steps, that satisfy
⟨𝛿𝑡 𝒖 ℎ , 𝒗 ℎ ⟩ + 𝐵(𝒖 ℎ ; 𝒖 ℎ , 𝒗 ℎ ) = 𝐿 (𝒗 ℎ ) ∀ 𝒗 ℎ ∈ Bℎ ,
where we have omitted the initial conditions and the time step superscript for sim-
plicity. This problem may suffer from instabilities and, hence, it requires the use of
stabilization methods like those based on VMS method.
The core idea of the VMS approach lies in the splitting B𝑐 = Bℎ ⊕ B ′ , where
B is any space that completes Bℎ in B𝑐 . We call B ′ the space of sub-grid scales
′
or subscales (SGS), and the functions in the SGS spaces will be identified with the
superscript ′ . Using the splitting 𝒖 = 𝒖 ℎ + 𝒖 ′ and similarly for the test function
𝒗 = 𝒗 ℎ + 𝒗 ′ , the continuous problem (22) splits into
Contents 19
Important Considerations/Assumptions
N (𝒖; 𝒖 ′ )| 𝐾 ≈ 𝝉 𝐾
−1
(𝒖)𝒖 ′ | 𝐾 ,
⟨𝛿𝑡 𝒖 ℎ , 𝒗 ℎ ⟩ + 𝐵 ℎ (𝒖 ∗ ; 𝒖 ℎ , 𝒗 ℎ ) = 𝐿 ℎ (𝒖 ∗ ; 𝒗 ℎ ), ∀ 𝒗 ℎ ∈ Bℎ , (26)
𝐵 ℎ (𝒖 ∗ ; 𝒖 ℎ , 𝒗 ℎ ) = 𝐵(𝒖 ∗ ; 𝒖 ℎ , 𝒗 ℎ ) + 𝐵′ (𝒖 ∗ ; 𝒖 ℎ , 𝒗 ℎ ) (27)
∗ ′ ∗
𝐿 ℎ (𝒖 ; 𝒗 ℎ ) = 𝐿(𝒗 ℎ ) + 𝐿 (𝒖 ; 𝒗 ℎ ) (28)
20 Contents
A ROM for the FOM discussed above can be developed by constructing a ROM
space B𝑟 ⊂ Bℎ ⊂ B𝑐 . Using the POD relying on SVD for functions, described in
Section 2.2.2, we may obtain a ROM space of dimension 𝑟
ˆ 1, 𝚽
B𝑟 = span{𝚽 ˆ 2 , ..., 𝚽
ˆ 𝑟 },
B = Bℎ ⊕ B ′ = B𝑟 ⊕ B ′′ ,
Then, since the basis vectors obtained from the POD are 𝐿 2 (Ω)-orthogonal, choosing
orthogonal subgrid scales allows us to write
B ′′ = span{𝚽
ˆ 𝑟+1 , 𝚽 ˆ 𝑛 𝑝 } ⊕ B′,
ˆ 𝑟+2 , ..., 𝚽
i.e., we have an explicit representation of the ROM space of SGSs. So, when VMS-
ROM is used to approximate ROM SGSs, it accounts for the FOM subscales, present
in the subspace B ′ , as well as the SGSs arising as a result of ROM trunctaion, present
in the subspace spanned by {𝚽 ˆ 𝑟+1 , 𝚽
ˆ 𝑟+2 , ..., 𝚽
ˆ 𝑛 𝑝 }.
Having in mind the previous discussion, the final reduced order problem can be
written as finding 𝒖𝑟𝑛 ∈ B𝑟 , for 𝑛 = 1, ..., 𝑛𝑡, that satisfy
⟨𝛿𝑡 𝒖𝑟 , 𝒗𝑟 ⟩ + 𝐵𝑟 (𝒖 ∗ ; 𝒖𝑟 , 𝒗𝑟 ) = 𝐿 𝑟 (𝒖 ∗ ; 𝒗𝑟 ), ∀ 𝒗𝑟 ∈ B𝑟 , (29)
Contents 21
𝐵𝑟 (𝒖 ∗ ; 𝒖𝑟 , 𝒗𝑟 ) = 𝐵(𝒖 ∗ ; 𝒖𝑟 , 𝒗𝑟 ) + 𝐵′ (𝒖 ∗ ; 𝒖𝑟 , 𝒗𝑟 ) (30)
∗ ′ ∗
𝐿 𝑟 (𝒖 ; 𝒗𝑟 ) = 𝐿(𝒗𝑟 ) + 𝐿 (𝒖 ; 𝒗𝑟 ). (31)
It can be seen that the Equations (29)-(31) look exactly the same as (26)-(28).
Furthermore, the expressions for 𝐵′ and 𝐿 ′ are also the same for the FOM and the
ROM if the same choices are made for both, regarding the considerations discussed in
Section 3.3.2. The only difference between the FOM and the ROM formulation is that
in the case of ROM, functions are approximated in B𝑟 instead of Bℎ . 𝐵′ (𝒖 ∗ ; 𝒖𝑟 , 𝒗𝑟 )
and 𝐿 ′ (𝒖 ∗ ; 𝒗𝑟 ) for different combination of choices can be found in [118].
FOM and the ROM implementations completely and are particularly useful in the
cases where the code used for the FOM is not open-source. NIROMs can be obtained
using conventional or ML-based techniques and the recent large-scale adoption of
NIROMs can be attributed to the increasing popularity of ML in scientific computing.
The ML-based NIROMs are later discussed in Section 6.2. For now, we describe
dynamic mode decomposition (DMD), which can be considered a conventional
non-intrusive extension of the POD.
𝑺∗ = 𝑻𝑺 (32)
Now, it is desired to find the eigenvalues and eigenvectors of matrix 𝑻. This matrix
𝑻 is a 𝑛𝑝 × 𝑛𝑝 matrix, and hence, it is impractical to perform its eigendecomposition.
DMD provides an efficient way of finding the 𝑟 leading eigenvalues and eigenvectors
of matrix 𝑻. Using 𝑺+ as the pseudo-inverse of 𝑺, (32) can be written as
𝑻 = 𝑺∗ 𝑺+ (33)
SVD can be used approximate the pseudo-inverse 𝑺+ . Using a truncated SVD, it can
be written
𝑺≈𝚿 ˆ𝚲ˆ 𝑽ˆ 𝑇 . (34)
As 𝚿ˆ and 𝑽ˆ are orthogonal and satisfy 𝚿ˆ 𝑇𝚿ˆ = 𝑰 𝑛 𝑝 and 𝑽ˆ 𝑇 𝑽ˆ = 𝑰 𝑛𝑠 , using (34)
allows us to write (33) as
𝑻 ≈ 𝑺∗ 𝑽ˆ 𝚲ˆ −1 𝚿
ˆ 𝑇,
where 𝚲 ˆ is a diagonal matrix and can be easily inverted. We are only interested in
the first 𝑟 eigenvalues and eigenvectors of matrix 𝑻. The 𝑟-rank approximation of 𝑻,
denoted by 𝑻 𝑟 ∈ R𝑟 ×𝑟 , is achieved by projecting 𝑻 on to the reduced space using
basis 𝚿ˆ as
𝑇
ˆ 𝑻𝚿
𝑻𝑟 = 𝚿 ˆ
=𝚿 ˆ −1 𝚿
ˆ 𝑇 𝑺∗𝑽ˆ 𝚲 ˆ 𝑇𝚿
ˆ
ˆ −1 .
ˆ 𝑇 𝑺∗𝑽ˆ 𝚲
=𝚿
ˆ −1 𝑬
𝝋 = 𝑺∗𝑽ˆ 𝚲
where the columns of 𝝋 ∈ R𝑛 𝑝×𝑟 , called DMD modes, are the eigenvectors of 𝑻.
Once DMD eigenvalues and eigenvectors have been determined, the state of the
system at the 𝑘-th time-step, 𝑼 𝑘 , is given by
𝑼 𝑘 = 𝝋Υ 𝑘−1 𝑫,
where 𝑫 ∈ R𝑟 is the vector of mode amplitudes that can be computed using initial
conditions. The DMD of a flow over a cylinder is illustrated in Fig. 3.
24 Contents
1
1.5
mode1 mode2 mode3
1
S
0.5
2
DMD
0
+
-0.5
3
-1
S*
-1.5
0 50 100 150
t
Fig. 3 Illustration of DMD applied to a flow over a cylinder. Three DMD modes and the temporal
evolution of their coefficients is shown.
In the previous sections, we have discussed how to build a ROM during an offline
stage and how to use it for getting results quickly during an online stage. So far, we
have assumed that the unknown 𝑼(𝑡, 𝜇) was a function of 𝑡 only and the parameter
𝜇 ∈ D ⊂ R, was kept constant. So, in essence, the ROM was used to solve exactly
the same problem whose solution was used to generate the snapshots to be used
for the ROM basis generation. The aim of reduced order modeling is to perform
the computationally expensive offline stage once (or a few times) and then use the
generated ROM to perform many simulations in the cheap online phase for the new
values of the parameter 𝜇. This situation arises routinely in optimization and control
problems governed by parametric PDEs. The parameter can represent anything
including boundary conditions, geometry, viscosity, Reynold’s number, etc. For
simplicity, we assume that the parameter represents a scalar and its different values,
𝜇1 , ..., 𝜇 𝑝𝑠 , represent different configurations, however, the subsequent discussion
is equally valid where 𝜇 represents more than one parameters. The difficulty with
parametric reduced order models (PROMs) lies in the fact that the basis 𝚿 𝜇1 obtained
for 𝜇1 is unlikely to perform well for 𝜇2 as the behavior captured by the basis 𝚿 𝜇1
might be different from the behavior exhibited by the system for 𝜇2 , i.e.
𝑼(𝜇1 ) ≈ 𝚿 𝜇1 𝑼𝑟 (𝜇1 ),
but
𝑼(𝜇2 ) 0 𝚿 𝜇1 𝑼𝑟 (𝜇2 ).
Several techniques have been developed to obtain a suitable basis for PROMs.
Hyperreduction techniques, like DEIM described in Section 3.2.1, can also be used
for PROMS [10] with 𝜃 = 𝜇. Here, we describe two popular techniques to obtain
a basis for PROMs, the global basis method and the local basis with interpolation
method. These techniques commonly use a greedy approach to sample suitable
Contents 25
parameter values to obtain the snapshots. Thus, they are commonly referred to as
POD-greedy approaches.
Probably the most obvious approach is to sample different parameter values, obtain
snapshots corresponding to them and perform the SVD on all the snapshots to obtain
ˆ such that
a single global basis 𝚿
ˆ 𝑟 (𝜇),
𝑼(𝜇) ≈ 𝚿𝑼 ∀ 𝜇 ∈ D. (35)
A greedy approach can be used to sample 𝑝𝑠 parameter values to obtain the snapshots.
The global basis approach can provide a compact 𝑟 dimensional basis satisfying (35)
if the solution is not very sensitive to the parameter 𝜇, i.e., the solution manifold
has rapidly decaying Kolmogorov 𝑛-width. If the solution manifold has slow decay-
ing Kolmogorov 𝑛-width, it might require obtaining snapshots at a lot of sampled
parameter values, which can lead to a prohibitively expensive offline phase. Even
if the computational expense of the offline phase is completely ignored, achieving
a reasonable accuracy in the online phase will require a lot of POD modes. Hence,
truncating the global basis to a rank 𝑟, ensuring a real-time execution of the online
phase with reasonable accuracy, will not be possible.
In the case that the global basis approach is not feasible, local basis can be developed
and used with interpolation. Similar to the global basis approach, 𝑝𝑠 parameter
values are sampled and the snapshots are obtained for them. However, instead of
performing a SVD of the matrix containing all the snapshots, a separate SVD is
performed for the snapshot matrix for every sampled parameter value 𝜇 to obtain a
corresponding local basis 𝚿 𝜇𝑖 , for 𝑖 = 1, ..., 𝑝𝑠. Now, the basis 𝚿 𝜇∗ can be obtained
at a requested, but unsampled, parameter value 𝜇∗ using interpolation.
If conventional interpolation techniques are used, the interpolated basis 𝚿 𝜇∗ is
likely to lose the key properties, e.g., orthogonality, after interpolation. Hence, inter-
polation using property preserving matrix manifolds is recommended to preserve the
𝑝𝑠
key properties. Let G be such a manifold of orthogonal matrices. Also, let {𝜇𝑖 }𝑖=1 be
𝑝𝑠
the set of sampled parameter values and {𝚿 𝜇𝑖 }𝑖=1 be the set of corresponding bases.
The basis 𝚿 𝜇∗ for the unsampled point 𝜇∗ ∉ {𝜇𝑖 }𝑖=1 𝑝𝑠
can be obtained as follows.
First, a tangent space T (𝚿 𝜇˜ ) is defined such that it is tangent to G at a reference
𝑝𝑠
point 𝚿 𝜇˜ ∈ {𝚿 𝜇𝑖 }𝑖=1 . Now, 𝚿 𝜇𝑖 , 𝑖 = 1, ..., 𝑝𝑠, except the reference point 𝚿 𝜇˜ , are
projected to the tangent space using a logarithmic map defined as
𝑇 (𝚿 𝜇𝑖 ) = 𝑹 𝜇𝑖 tan−1 (𝚵 𝜇𝑖 )𝑾 𝑇𝜇𝑖
26 Contents
𝑇 (𝚿 𝜇∗ ) = 𝑹 𝜇∗ 𝚵 𝜇∗ 𝑾 𝑇𝜇∗ .
𝑇(𝚿𝜇 2 )
𝑇(𝚿𝜇 ∗ ) 𝑇(𝚿𝜇 𝑝𝑠 )
𝑇(𝚿𝜇 1 )
𝚿𝜇
𝚿𝜇 ∗ 𝚿𝜇 𝑝𝑠
𝚿𝜇 2
𝚿𝜇 1
𝑝𝑠
Fig. 4 Interpolation of a set of matrices {𝚿 𝜇𝑖 }𝑖=1 using the matrix manifold G and the tangent
plane T (𝚿 𝜇˜ ).
Remark 5 The above described manifold based interpolation has been shown to be
applicable to the direct interpolation of reduced order system matrices/vectors as
well for linear systems [7]. Consider a spatially discretized reduced order parametric
linear system
𝑴 𝑟 (𝜇∗ )𝜕𝑡 𝑼𝑟 (𝜇∗ ) + 𝑳 𝑟 (𝜇∗ )𝑼𝑟 (𝜇∗ ) = 𝑭𝑟 (𝜇∗ ).
𝑝𝑠
If {𝑴 𝑟 (𝜇𝑖 )}𝑖=1 𝑝𝑠
, {𝑳 𝑟 (𝜇𝑖 )}𝑖=1 𝑝𝑠
and {𝑭𝑟 (𝜇𝑖 )}𝑖=1 are obtained offline, 𝑴 𝑟 (𝜇∗ ), 𝑳 𝑟 (𝜇∗ )
∗
and 𝑭𝑟 (𝜇 ) can be obtained during the online phase using the manifold based inter-
polation. This ensures that the key properties, e.g., symmetric positive definiteness
(SPD), is preserved after the interpolation. This direct interpolation is more efficient
than first finding the interpolated basis 𝚿 𝜇∗ and then finding the reduced matrix
𝑿 𝑟 (𝜇∗ ) using 𝚿𝑇𝜇∗ 𝑿𝚿 𝜇∗ . However, this direct interpolation has been shown to work
Contents 27
for linear problems only so far. The appropriate logarithmic and exponential maps
to be used to preserve different matrix properties can be found in [7].
The impact of ML has been profound on scientific computing. In this section, the
applications of ML in the context of projection based ROMs are explored. However, it
is pertinent to mention the natural suitability of ML techniques to develop extremely
computationally inexpensive models, even beyond the context of projection based
ROMs. A lot of the success of ML techniques can be attributed to their ability to
find and learn nonlinear mappings that govern a physical system. For any system,
a few key inputs can be selected and a ML technique can be applied to learn the
(non)linear mapping that exists between its outputs and the selected inputs. Since
the online (testing) phase of ML algorithms is very fast, any such application would
result in a computationally inexpensive model, i.e., a ROM.
In the context of projection based ROMs, ML techniques have been applied to
achieve higher accuracy, improvement in speeds or a combination of both. ML
techniques have been applied to obtain nonlinear reduced spaces for ROMs which
offer a more compact representation than linear POD spaces. ML techniques have
also been used to obtain NIROMs by directly learning the nonlinear evolution of
reduced coordinates, previously referred to as ROM coefficients in the context of
POD. The term reduced coordinates is more popular in the literature in the context of
ML-based ROMs, and hence, it will be used from here on. ML can also improve the
accuracy of the intrusive Galerkin ROMs by providing closure models or corrections
based on finer solutions. Purely data-driven ML techniques can be very data hungry.
In order to reduce the reliance on data, and improve the generalization of the ML
models, physics has been embedded in ML techniques and then applied to reduced
order modeling. ML has even been used for system identification to discover simple
equations for the evolution of reduced coordinates. Let us describe the state-of-art
in the above-mentioned application domains.
POD (or DMD) provides modes that approximate a linear subspace. However, the
evolution of many dynamical systems lies in nonlinear spaces. The linear approxima-
tion can lead to two issues. First, the POD modes might be unable to capture highly
nonlinear phenomena. Second, in the case that the linear representation can cap-
ture the dynamics with reasonable accuracy, using POD may require more reduced
coordinates than the nonlinear representation. So, instead of the linear mapping pro-
vided by the POD (13), a more robust mapping would be using a nonlinear function
𝜗𝐷 : R𝑟 → R𝑛 𝑝 given by
28 Contents
𝑼 ≈ 𝑼 𝐷 = 𝜗𝐷 (𝑼𝑟 ), (36)
where 𝑼 𝐷 ∈ R𝑛 𝑝 is the mapped value and 𝑼 is the FOM solution. This nonlinear
mapping can be achieved using autoencoders (AEs) [21].
AEs are artificial neural networks (ANNs) used widely for dimension reduction.
The simplest AE, called undercomplete AE, consists of input and output layers of
the same size as the size of the FOM, 𝑛𝑝 in this case. Furthermore, it has a bottleneck
layer in the middle of the desired size 𝑟, the same as the size of the reduced space.
The architecture of an undercomplete AE is shown in Fig. 5. In general, AEs are
quite deep, i.e., they consist of many layers, and use nonlinear activation functions.
Based on the task performed, an AE can be subdivided into two sub-parts: an
encoder and a decoder. The encoder compresses the high-dimensional data succes-
sively through its many layers to produce the low-dimensional representation. The
encoder can be represented by a function 𝜗𝐸 : R𝑛 𝑝 → R𝑟 as
𝑼𝑟 = 𝜗𝐸 (𝑼).
The decoder reproduces the high dimensional representation from the low dimen-
sional one as per the mapping (36). To train an AE, 𝑼 is given both as the input and
the desired output and the loss function ∥𝑼 𝐷 − 𝑼∥ 22 is minimized, i.e., AE as a whole
is expected to behave like an identity map. Interestingly, the optimal solution using
the encoder and decoder of single layers with linear activation functions is shown
to be closely related to POD [20], i.e., POD can be considered to be a linear AE.
AEs have been used with simple feedforward [102], as well as convolutional [69,
92, 108] layers and have shown performance enhancement as compared to the linear
dimension reduction techniques. Furthermore, time-lagged AEs have been shown to
capture the slowly evolving dynamics of chemical processes with higher precision
[142].
An important issue of AEs is that they do not provide a systematic way of
determining the suitable dimension of the reduced space as they do not provide
hierarchical reduced coordinates, as POD does based on the RIC index. The number
of reduced coordinates needs to be provided a priori to an AE as the size of the
bottleneck. A smaller number of reduced coordinates cannot be selected as the
coordinates are not distributed hierarchically and each coordinate may correspond
to roughly the same RIC. A trial-and-error approach can be used to find the optimal
dimension of the reduced space, but this is not an efficient approach. Variational
autoencoders [90] can resolve this issue and provide a parsimonious set of the reduced
coordinates. In [58], 𝛽-Variational autoencoders [77] uses 𝛽 as a hyperparameter to
promote the sparsity of reduced coordinates by deactivating the unimportant nodes
of the bottleneck. 𝛽-Variational autoencoders were shown to represent ∼ 90% of the
energy for the problem of a flow over an urban environment, using five modes only,
in contrast to just ∼ 30% captured by the first five POD modes. AEs have also been
applied to discover nonlinear coordinate spaces for DMD [99, 112, 132].
Contents 29
Bottleneck
Encoder Decoder
Fig. 5 Architecture of an undercomplete autoencoder with encoder-decoder parts. The number of
layers and the size of the bottleneck is set to three and two, respectively, for illustration purposes.
LSTM and BiLSTM NNs have been widely used to predict the temporal evolution
of systems based on the past values. LSTM and BiLSTM NNs were used to model
isotropic and magnetohydrodynamic turbulence in [103], where the Hurst Expo-
nent [81] was employed to study and quantify the memory effects captured by the
LSTM/BiLSTM NNs. LSTM NNs were used in [137] to predict high-dimensional
chaotic systems and were shown to outperform Gaussian processes. The improved
performance of LSTM NNs was also shown for reduced order modeling of near-wall
turbulence as compared to FNNs in [130]. Finally, LSTM NNs were also used to
model a synthetic jet and three dimensional flow in the wake of a cylinder in [1].
Training ML algorithms in general, and LSTM/BiLSTM NNs in particular, can
be very computationally demanding. Transfer learning can be used to speed up the
training phase. Instead of initializing the weights of a network randomly, transfer
learning relies on using weights of a network previously trained for a closely related
problem for initialization. Providing better initial weights allows the training to
converge faster to the optimal weights. Transfer learning was used to speed-up the
training of LSTM and BiLSTM NNs modeling the three-dimensional turbulent wake
of a finite wall-mounted square cylinder in [146]. The flow was analyzed on 2D planes
at different heights, each modeled using a separate LSTM/BiLSTM NN. After the
first LSTM/BiLSTM NN was trained, the others were initialized using its weights,
as the flow in different planes is correlated.
Gaussian process regression [117] can be used to build NIROMs alongside pro-
viding uncertainty quantification. Gaussian process regression has been used to
develop NIROM for shallow water equations [101], chaotic systems like climate
forecasting models [138] and nonlinear structural problems [73]. In the domain of
unsupervised learning, cluster reduced order modeling has been applied to mixing
layers [86]. The cluster reduced order modeling groups the ensemble of data (snap-
shots) into clusters and the transitions between the states are dynamically modeled
using a Markov process.
Addi�onal Addi�onal
inputs inputs
𝜕𝑡 𝑼𝑟 = 𝑮 (𝑼𝑟 , 𝑼).
e
Using the truncated basis to build the ROM implicitly implies, abusing of the nota-
tion,
𝜕𝑡 𝑼𝑟 = 𝑮 (𝑼𝑟 , 0) = 𝑮 (𝑼𝑟 ),
which is not true in the nonlinear cases, as the behavior of the resolved scales is
governed by their interaction with the unresolved ones as well. So, it is desired to
model this interaction as a term 𝑪 (𝑼𝑟 ), which is a function of the resolved scales
𝑼𝑟 , so that
𝜕𝑡 𝑼𝑟 = 𝑮 (𝑼𝑟 ) + 𝑪 (𝑼𝑟 ). (38)
32 Contents
Closure error
modeled by C(Ur)
Projection error
Fig. 7 An illustration of the ROM closure problem. The projection error is due to the use of the
space B𝑟 ⊂ B 𝑓 for approximating the unknowns. The closure error is due to neglecting the effect
of nonlinear interaction on the evolution of the resolved coordinates.
where the memory integral is a non-Markovian 2 term and takes into account the
contribution of the past values of the resolved coordinates 𝑼𝑟 to model the unresolved
ones 𝑼.
e This memory integral is very computationally expensive to compute. To
evaluate it efficiently, neural closure models using neural delay differential equations
were proposed in [149]. The number of past values required to accurately determine
2 A non-Markovian term implies that the future state depends not only on the current values, but
also on the past value, i.e., such processes have memory effects of the past values.
Contents 33
the closure term was also determined. A conditioned LSTM NN was used to model
the memory term in [140]. To ensure the computational efficiency, the authors further
used explicit time integration of the memory term, while using implicit integration
for the remaining terms of the discrete system. The Mori-Zwanzig formalism and
LSTM NNs were shown to have a natural analogy between them in [100]. This
analogy was also used to develop a systematic approach for constructing memory
based ROMs. Recently, reinforcement learning techniques have also been applied to
build unsupervised reward-oriented closure models for ROMs [24, 126].
𝑨𝑼 𝑐 = 𝑹. (40)
Let the fine solution 𝑼 𝑓 be also available for the given problem. Let us assume that
the projection of the fine solution onto the coarse solution space, denoted by 𝑼 𝑐 𝑓 , is
the best possible coarse solution. In this case, a correction vector 𝑪 can be added to
modify system (40) to obtain a new system
𝑨𝑼 𝑐 𝑓 + 𝑪 = 𝑹, (41)
with 𝑼 𝑐 𝑓 as its solution. When the fine solution is not known, 𝑪 needs to be modeled
as a function of the coarse solution, i.e., 𝑪 = 𝑪 (𝑼 𝑐 ).
A correction term using a least-squares (LS) approach was proposed for POD-
ROM in [17]. Special considerations regarding gathering training data and using
the appropriate initial conditions were also addressed with least-squares providing
a linear model of the correction term. A nonlinear correction term was determined
using ANN in [18]. The correction term was determined for a coarse mesh based
ROM using the solution of an AMR based FOM, and applied to fluid, structure and
FSI problems . A comparison of linear least-squares and nonlinear ANNs based
corrections was carried out for the wave equation in [60]. A nonlinear ANN based
correction for POD-ROM was used in [54]. Different combination of features to be
provided as the inputs to the ANNs were evaluated to develop an accurate model
while minimizing the complexity. The implicit and explicit treatment of the ANN
based correction was also evaluated. It was shown that the ROM was able to produce
good results for parametric interpolation, as well as temporal extrapolation. All of
34 Contents
the above mentioned works relied on significantly less training data as compared to
NIROMs. A training-free correction was further proposed in [19] to account for the
loss of information due to the adaptive coarsening of the coarse mesh based ROM.
The correction was based solely on the data generated within the same simulation
and did not require any external data.
Remark 6 The correction based on fine solutions discussed in Section 6.4 and the
closure modeling discussed in Section 6.3 are similar to some extent. However, there
is a difference in their motivation, as well as the accepted definition in the literature.
Closure modeling for ROMs is understood to account for the error generated due
to the Galerkin projection, i.e., spatial discretization as given by (38). On the other
hand, the correction based on fine solutions works by introducing a correction at the
fully discrete level given by (41). Hence, the two approaches have been discussed
separately.
ML has also been used to obtain nonlinear solution manifolds for parametric
ROMs [92]. Deep convolutional auto-encoders were used in [69] to generate a
reduced parametric trial manifold. It was then combined with a LSTM NN to model
the dynamics of the reduced coordinates. Nonlinear reduced parametric manifolds
were also learned using AEs in [62] and a variation of it was developed in [63] to
speed up the offline training process.
J𝑃 = ∥ 𝑨(𝑼 𝑀 𝐿 ) ∥ 2 , (43)
J𝐷 = ∥𝑼 𝑀 𝐿 − R (𝑼) ∥ 2 . (44)
In general, the training phase involves minimizing the mean of (43) and (44) for
multiple time-steps and/or parameter values. The total loss function is given by
J = J𝐷 + 𝜀J𝑃 , (45)
incorporating the residual, arising from violating the conservation laws using finite
volume discretization, in the loss function. Embedding physics in a data-driven
ROM closure model reduced the data requirement by 50% in [105]. The physics was
incorporated by requiring some terms of the closure model to be energy dissipating,
while others to be energy conserving.
Introducing physics using (45), so that the training phase becomes a constrained
optimization problem, can be considered as applying weak constraints (note that
J𝑃 = 0 if 𝑼 𝑀 𝐿 is replaced by 𝑼𝑟 in Eq. (43)). The violation of the physics leads
to a large loss function, and hence, in an effort to minimize the loss function, the
ANN tries to adhere to the physics as well. The physics embedded in such a way is
prone to be violated in the testing phase when the ANN is exposed to cases beyond
the training phase. This is because architecturally, the ANN is still unaware of the
physics. Furthermore, 𝜀 acts as an additional hyperparameter which needs to be
tuned.
An alternative strategy is to amend the ANN structurally so that it enforces the
physical laws strongly. Such an ANN is hoped to be more robust in the testing phase
as it is not blind to the physics. Embedded physics in a coarse grained ROM for 3D
turbulence via hard constraints was achieved in [104]. The divergence free condition
was enforced using curl-computing layers which formed a part of the backpropagation
process as well. Backpropagation through the curl operator ensured that the network
is not blind and has intimate knowledge of the constraints through which it must
make predictions. Another way of embedding physics in the layers was proposed in
[115]. A physics-guided machine learning framework was introduced which injected
the desired physics in one of the layers of the LSTM NN. By incorporating physics,
an ANN applicable to more generalized cases was achieved as compared to the
purely data-driven approach.
ML can also be used to find the equations representing the dynamics of a system
using data. An equation comprising of different terms with adjustable coefficients
is assumed to model the behavior of a system. Data is then used to find the value
of these adjustable coefficients. This technique is known as system identification
and is of particular interest in two scenarios. First, if the equations are not known,
as in the case of modeling climate, epidemiology, neuroscience, etc. Second, when
the equations describing the behavior of the reduced coordinates are required. The
resulting equation based representation of a system provides generalizability and
interpretability, not achievable by simply constructing a regression model based on
data. System identification is a broad field and many techniques have been applied
in this context, see [95] and [84] for details.
Reduced system identification aims to obtain sparse equations (consisting of a few
simple terms) to describe the evolution of reduced coordinates of a projection based
ROM. The sparse identification of nonlinear dynamics (SINDy) [28] algorithm can
Contents 37
be used to get a simplistic dynamic model for the reduced coordinates. A library
of simple nonlinear terms, like polynomials or trigonometric functions, is provided.
SINDy then tries to find a mapping for the provided input-output data using the
minimum number of terms of the library, thus providing a minimalistic interpretable
model offering a balance of efficiency and accuracy.
SINDy has been applied to recover models for a variety of flow behaviors includ-
ing shear layers, laminar and turbulent wakes, and convection [33, 55, 96, 97]. The
vortex shedding behind a cylinder, for example, can be captured using three modes
only [109], first two POD modes and a shift mode as
where 𝑈𝑟1 , 𝑈𝑟2 and 𝑈𝑟3 are the reduced coordinates related to first two POD modes
and the shift mode. SINDy was able to recover this minimalistic model using data,
identifying the dominant terms and the associated coefficients correctly [28]. SINDy
was also combined with an autoencoder to find the low dimensional nonlinear
representation, as well as to model the dynamics of the corresponding reduced
coordinates [36]. To improve the performance of SINDy, physics was also embedded
in it in the form of symmetry in [72] and of conservation laws in [97].
7 Concluding Remarks
been developed based entirely on ML. Thus, a range of reduced order modeling
techniques are at disposal with purely conventional techniques at one end of the
spectrum to purely ML techniques at the other end, with the hybrid techniques lying
in-between.
References
[1] R. Abadı́a-Heredia et al. “A Predictive Hybrid Reduced Order Model Based on Proper Orthogonal Decom-
position Combined with Deep Learning Architectures”. In: Expert Systems with Applications 187 (2022),
p. 115910.
[2] H. F. Ahmed et al. “Machine Learning–Based Reduced-Order Modeling of Hydrodynamic Forces Using
Pressure Mode Decomposition”. In: Proceedings of the Institution of Mechanical Engineers, Part G: Journal
of Aerospace Engineering 235.16 (2021), pp. 2517–2528.
[3] S. E. Ahmed et al. “A Long Short-Term Memory Embedding for Hybrid Uplifted Reduced Order Models”. In:
Physica D: Nonlinear Phenomena 409 (2020), p. 132471.
[4] S. E. Ahmed et al. “On Closures for Reduced Order Models—A Spectrum of First-Principle to Machine-Learned
Avenues”. In: Physics of Fluids 33.9 (2021), p. 091301.
[5] I. Akhtar, J. Borggaard, and A. Hay. “Shape Sensitivity Analysis in Flow Models Using a Finite-Difference
Approach”. In: Mathematical Problems in Engineering 2010 (2010).
[6] A. Alla and J. N. Kutz. “Nonlinear Model Order Reduction via Dynamic Mode Decomposition”. In: SIAM
Journal on Scientific Computing 39.5 (2017), B778–B796.
[7] D. Amsallem and C. Farhat. “An Online Method for Interpolating Linear Parametric Reduced-Order Models”.
In: SIAM Journal on Scientific Computing 33.5 (2011), pp. 2169–2198.
[8] D. Amsallem and C. Farhat. “Stabilization of Projection-Based Reduced-Order Models”. In: International
Journal for Numerical Methods in Engineering 91.4 (2012), pp. 358–377.
[9] S. S. An, T. Kim, and D. L. James. “Optimizing Cubature for Efficient Integration of Subspace Deformations”.
In: ACM Transactions on Graphics 27.5 (2008), 165:1–165:10.
[10] H. Antil, M. Heinkenschloss, and D. C. Sorensen. “Application of the Discrete Empirical Interpolation Method
to Reduced Order Modeling of Nonlinear and Parametric Systems”. In: Reduced Order Methods for Modeling
and Computational Reduction. Ed. by A. Quarteroni and G. Rozza. MS&A - Modeling, Simulation and
Applications. Cham: Springer International Publishing, 2014, pp. 101–136.
[11] E. Arian, M. Fahl, and E. W. Sachs. Trust-Region Proper Orthogonal Decomposition for Flow Control. Tech.
rep. Institute for Computer Applications in Science and Engineering, Hampton VA, 2000.
[12] P. Astrid et al. “Missing Point Estimation in Models Described by Proper Orthogonal Decomposition”. In:
IEEE Transactions on Automatic Control 53.10 (2008), pp. 2237–2251.
[13] M. Azaı̈ez, T. Chacón Rebollo, and S. Rubino. “A Cure for Instabilities Due to Advection-Dominance in
POD Solution to Advection-Diffusion-Reaction Equations”. In: Journal of Computational Physics 425 (2021),
p. 109916.
[14] J. Baiges and R. Codina. “A Variational Multiscale Method with Subscales on the Element Boundaries for the
Helmholtz Equation”. In: International Journal for Numerical Methods in Engineering 93.6 (2013), pp. 664–
684.
[15] J. Baiges, R. Codina, and S. Idelsohn. “A Domain Decomposition Strategy for Reduced Order Models. Ap-
plication to the Incompressible Navier–Stokes Equations”. In: Computer Methods in Applied Mechanics and
Engineering 267 (2013), pp. 23–42.
[16] J. Baiges, R. Codina, and S. Idelsohn. “Explicit Reduced-Order Models for the Stabilized Finite Element
Approximation of the Incompressible Navier–Stokes Equations”. In: International Journal for Numerical
Methods in Fluids 72.12 (2013), pp. 1219–1243.
[17] J. Baiges, R. Codina, and S. Idelsohn. “Reduced-Order Subscales for POD Models”. In: Computer Methods in
Applied Mechanics and Engineering 291 (2015), pp. 173–196.
[18] J. Baiges et al. “A Finite Element Reduced-Order Model Based on Adaptive Mesh Refinement and Artificial
Neural Networks”. In: International Journal for Numerical Methods in Engineering 121.4 (2020), pp. 588–601.
[19] J. Baiges et al. “An Adaptive Finite Element Strategy for the Numerical Simulation of Additive Manufacturing
Processes”. In: Additive Manufacturing 37 (2021), p. 101650.
[20] P. Baldi and K. Hornik. “Neural Networks and Principal Component Analysis: Learning from Examples without
Local Minima”. In: Neural Networks 2.1 (1989), pp. 53–58.
[21] D. H. Ballard. “Modular Learning in Neural Networks”. In: Proceedings of the Sixth National Conference on
Artificial Intelligence - Volume 1. AAAI’87. Seattle, Washington: AAAI Press, 1987, pp. 279–284.
[22] F. Ballarin et al. “Supremizer Stabilization of POD–Galerkin Approximation of Parametrized Steady Incom-
pressible Navier–Stokes Equations”. In: International Journal for Numerical Methods in Engineering 102.5
(2015), pp. 1136–1161.
Contents 39
[23] M. Barrault et al. “An ‘Empirical Interpolation’Method: Application to Efficient Reduced-Basis Discretization
of Partial Differential Equations”. In: Comptes Rendus Mathematique 339.9 (2004), pp. 667–672.
[24] M. Benosman, A. Chakrabarty, and J. Borggaard. “Reinforcement Learning-based Model Reduction for Partial
Differential Equations”. In: IFAC-PapersOnLine. 21st IFAC World Congress 53.2 (2020), pp. 7704–7709.
[25] M. Bergmann, L. Cordier, and J.-P. Brancher. “Drag Minimization of the Cylinder Wake by Trust-Region
Proper Orthogonal Decomposition”. In: Active Flow Control. Springer, 2007, pp. 309–324.
[26] D. Bertsimas and J. Dunn. “Optimal Classification Trees”. In: Machine Learning 106.7 (2017), pp. 1039–1082.
[27] A. N. Brooks and T. J. R. Hughes. “Streamline Upwind/Petrov-Galerkin Formulations for Convection Dominated
Flows with Particular Emphasis on the Incompressible Navier-Stokes Equations”. In: Computer Methods in
Applied Mechanics and Engineering 32.1 (1982), pp. 199–259.
[28] S. L. Brunton, J. L. Proctor, and J. N. Kutz. “Discovering Governing Equations from Data by Sparse Identifi-
cation of Nonlinear Dynamical Systems”. In: Proceedings of the National Academy of Sciences 113.15 (2016),
pp. 3932–3937.
[29] S. L. Brunton et al. “Chaos as an Intermittently Forced Linear System”. In: Nature Communications 8.1 (2017),
p. 19.
[30] T. Bui-Thanh, K. Willcox, and O. Ghattas. “Model Reduction for Large-Scale Systems with High-Dimensional
Parametric Input Space”. In: SIAM Journal on Scientific Computing 30.6 (2008), pp. 3270–3288.
[31] S. Buoso et al. “Stabilized Reduced-Order Models for Unsteady Incompressible Flows in Three-Dimensional
Parametrized Domains”. In: Computers & Fluids 246 (2022), p. 105604.
[32] J. Burkardt, M. Gunzburger, and H.-C. Lee. “POD and CVT-based Reduced-Order Modeling of Navier–Stokes
Flows”. In: Computer methods in applied mechanics and engineering 196.1-3 (2006), pp. 337–355.
[33] J. L. Callaham et al. “An Empirical Mean-Field Model of Symmetry-Breaking in a Turbulent Wake”. In: Science
Advances 8.19 (2022), eabm4786.
[34] K. Carlberg, M. Barone, and H. Antil. “Galerkin v. Least-Squares Petrov–Galerkin Projection in Nonlinear
Model Reduction”. In: Journal of Computational Physics 330 (2017), pp. 693–734.
[35] K. Carlberg, C. Bou-Mosleh, and C. Farhat. “Efficient Non-Linear Model Reduction via a Least-Squares
Petrov–Galerkin Projection and Compressive Tensor Approximations”. In: International Journal for Numerical
Methods in Engineering 86.2 (2011), pp. 155–181.
[36] K. Champion et al. “Data-Driven Discovery of Coordinates and Governing Equations”. In: Proceedings of the
National Academy of Sciences 116.45 (2019), pp. 22445–22451.
[37] A. Chatterjee. “An Introduction to the Proper Orthogonal Decomposition”. In: Current Science 78.7 (2000),
pp. 808–817.
[38] S. Chaturantabut and D. C. Sorensen. “Nonlinear Model Reduction via Discrete Empirical Interpolation”. In:
SIAM Journal on Scientific Computing 32.5 (2010), pp. 2737–2764.
[39] K. K. Chen, J. H. Tu, and C. W. Rowley. “Variants of Dynamic Mode Decomposition: Boundary Condition,
Koopman, and Fourier Analyses”. In: Journal of Nonlinear Science 22.6 (2012), pp. 887–915.
[40] W. Chen et al. “Physics-Informed Machine Learning for Reduced-Order Modeling of Nonlinear Problems”. In:
Journal of Computational Physics 446 (2021), p. 110666.
[41] Z. Chen, Y. Zhao, and R. Huang. “Parametric Reduced-Order Modeling of Unsteady Aerodynamics for Hyper-
sonic Vehicles”. In: Aerospace Science and Technology 87 (2019), pp. 1–14.
[42] F. Chinesta, A. Ammar, and E. Cueto. “Recent Advances and New Challenges in the Use of the Proper
Generalized Decomposition for Solving Multidimensional Models”. In: Archives of Computational Methods in
Engineering 17.4 (2010), pp. 327–350.
[43] F. Chinesta, P. Ladeveze, and E. Cueto. “A Short Review on Model Order Reduction Based on Proper Gener-
alized Decomposition”. In: Archives of Computational Methods in Engineering 18.4 (2011), p. 395.
[44] R. Codina. “On Stabilized Finite Element Methods for Linear Systems of Convection–Diffusion-Reaction
Equations”. In: Computer Methods in Applied Mechanics and Engineering 188.1 (2000), pp. 61–82.
[45] R. Codina. “Stabilization of Incompressibility and Convection through Orthogonal Sub-Scales in Finite Element
Methods”. In: Computer methods in applied mechanics and engineering 190.13-14 (2000), pp. 1579–1599.
[46] R. Codina. “Stabilized Finite Element Approximation of Transient Incompressible Flows Using Orthogonal
Subscales”. In: Computer methods in applied mechanics and engineering 191.39-40 (2002), pp. 4295–4321.
[47] R. Codina and J. Baiges. “Finite Element Approximation of Transmission Conditions in Fluids and Solids
Introducing Boundary Subgrid Scales”. In: International Journal for Numerical Methods in Engineering 87.1-
5 (2011), pp. 386–411.
[48] R. Codina, J. Principe, and J. Baiges. “Subscales on the Element Boundaries in the Variational Two-Scale Finite
Element Method”. In: Computer methods in applied mechanics and engineering 198.5-8 (2009), pp. 838–852.
[49] R. Codina, R. Reyes, and J. Baiges. “A Posteriori Error Estimates in a Finite Element VMS-based Reduced
Order Model for the Incompressible Navier-Stokes Equations”. In: Mechanics Research Communications.
Special Issue Honoring G.I. Taylor Medalist Prof. Arif Masud 112 (2021), p. 103599.
[50] R. Codina et al. “Time Dependent Subscales in the Stabilized Finite Element Approximation of Incompressible
Flow Problems”. In: Computer Methods in Applied Mechanics and Engineering 196.21-24 (2007), pp. 2413–
2430.
[51] R. Codina et al. “Variational Multiscale Methods in Computational Fluid Dynamics”. In: Encyclopedia of
computational mechanics (2018), pp. 1–28.
[52] N. Dal Santo et al. “An Algebraic Least Squares Reduced Basis Method for the Solution of Nonaffinely
Parametrized Stokes Equations”. In: Computer Methods in Applied Mechanics and Engineering 344 (2019),
pp. 186–208.
40 Contents
[53] T. Daniel et al. “Model Order Reduction Assisted by Deep Neural Networks (ROM-net)”. In: Advanced Modeling
and Simulation in Engineering Sciences 7.1 (2020), p. 16.
[54] Z. Dar, J. Baiges, and R. Codina. “Artificial neural network based correction models for reduced order models
in computational fluid mechanics”. In: Submitted (2022).
[55] N. Deng et al. “Low-Order Model for Successive Bifurcations of the Fluidic Pinball”. In: Journal of Fluid
Mechanics 884 (2020), A37.
[56] R. Dupuis, J.-C. Jouhaud, and P. Sagaut. “Surrogate Modeling of Aerodynamic Simulations for Multiple
Operating Conditions Using Machine Learning”. In: AIAA Journal 56.9 (2018), pp. 3622–3635.
[57] C. Eckart and G. Young. “The Approximation of One Matrix by Another of Lower Rank”. In: Psychometrika
1.3 (1936), pp. 211–218.
[58] H. Eivazi et al. “Towards Extraction of Orthogonal and Parsimonious Non-Linear Modes from Turbulent
Flows”. In: Expert Systems with Applications 202 (2022), p. 117038.
[59] R. Everson and L. Sirovich. “Karhunen–Loève Procedure for Gappy Data”. In: JOSA A 12.8 (1995), pp. 1657–
1664.
[60] A. Fabra, J. Baiges, and R. Codina. “Finite Element Approximation of Wave Problems with Correcting Terms
Based on Training Artificial Neural Networks with Fine Solutions”. In: Computer Methods in Applied Mechanics
and Engineering 399 (2022), p. 115280.
[61] C. Farhat, T. Chapman, and P. Avery. “Structure-Preserving, Stability, and Accuracy Properties of the Energy-
Conserving Sampling and Weighting Method for the Hyper Reduction of Nonlinear Finite Element Dynamic
Models”. In: International Journal for Numerical Methods in Engineering 102.5 (2015), pp. 1077–1110.
[62] S. Fresca, L. Dede’, and A. Manzoni. “A Comprehensive Deep Learning-Based Approach to Reduced Order
Modeling of Nonlinear Time-Dependent Parametrized PDEs”. In: Journal of Scientific Computing 87.2 (2021),
p. 61.
[63] S. Fresca and A. Manzoni. “POD-DL-ROM: Enhancing Deep Learning-Based Reduced Order Models for Non-
linear Parametrized PDEs by Proper Orthogonal Decomposition”. In: Computer Methods in Applied Mechanics
and Engineering 388 (2022), p. 114181.
[64] B. Galletti et al. “Low-Order Modelling of Laminar Flow Regimes Past a Confined Square Cylinder”. In:
Journal of Fluid Mechanics 503 (2004), pp. 161–170.
[65] B. Garcı́a-Archilla, J. Novo, and S. Rubino. “Error Analysis of Proper Orthogonal Decomposition Data Assim-
ilation Schemes with Grad–Div Stabilization for the Navier–Stokes Equations”. In: Journal of Computational
and Applied Mathematics 411 (2022), p. 114246.
[66] S. Giere and V. John. “Towards Physically Admissible Reduced-Order Solutions for Convection–Diffusion
Problems”. In: Applied Mathematics Letters 73 (2017), pp. 78–83.
[67] S. Giere et al. “SUPG Reduced Order Models for Convection-Dominated Convection–Diffusion–Reaction
Equations”. In: Computer Methods in Applied Mechanics and Engineering 289 (2015), pp. 454–474.
[68] B. Glaz, L. Liu, and P. P. Friedmann. “Reduced-Order Nonlinear Unsteady Aerodynamic Modeling Using a
Surrogate-Based Recurrence Framework”. In: AIAA journal 48.10 (2010), pp. 2418–2429.
[69] F. J. Gonzalez and M. Balajewicz. “Deep Convolutional Recurrent Autoencoders for Learning Low-Dimensional
Feature Dynamics of Fluid Systems”. In: arXiv preprint arXiv:1808.01346 (2018). arXiv: 1808 . 01346
[physics].
[70] W. R. Graham, J. Peraire, and K. Y. Tang. “Optimal Control of Vortex Shedding Using Low-Order Models.
Part I—Open-Loop Model Development”. In: International Journal for Numerical Methods in Engineering
44.7 (1999), pp. 945–972.
[71] A. Graves and J. Schmidhuber. “Framewise Phoneme Classification with Bidirectional LSTM and Other Neural
Network Architectures”. In: Neural Networks. IJCNN 2005 18.5 (2005), pp. 602–610.
[72] Y. Guan, S. L. Brunton, and I. Novosselov. “Sparse Nonlinear Models of Chaotic Electroconvection”. In: Royal
Society Open Science 8.8 (2021), p. 202367.
[73] M. Guo and J. S. Hesthaven. “Reduced Order Modeling for Nonlinear Structural Analysis Using Gaussian
Process Regression”. In: Computer methods in applied mechanics and engineering 341 (2018), pp. 807–826.
[74] M. Guo and J. S. Hesthaven. “Data-Driven Reduced Order Modeling for Time-Dependent Problems”. In:
Computer Methods in Applied Mechanics and Engineering 345 (2019), pp. 75–99.
[75] J. S. Hesthaven and S. Ubbiali. “Non-Intrusive Reduced Order Modeling of Nonlinear Problems Using Neural
Networks”. In: Journal of Computational Physics 363 (2018), pp. 55–78.
[76] J. S. Hesthaven, G. Rozza, and B. Stamm. Certified Reduced Basis Methods for Parametrized Partial Differential
Equations. SpringerBriefs in Mathematics. Cham: Springer International Publishing, 2016.
[77] I. Higgins et al. “Beta-VAE: Learning Basic Visual Concepts with a Constrained Variational Framework”. In:
International Conference on Learning Representations. 2022.
[78] S. Hochreiter and J. Schmidhuber. “Long Short-Term Memory”. In: Neural Computation 9.8 (1997), pp. 1735–
1780.
[79] T. J. R. Hughes et al. “The Variational Multiscale Method—a Paradigm for Computational Mechanics”. In:
Computer Methods in Applied Mechanics and Engineering. Advances in Stabilized Methods in Computational
Mechanics 166.1 (1998), pp. 3–24.
[80] A. Hunter et al. “Reduced-Order Modeling through Machine Learning and Graph-Theoretic Approaches for
Brittle Fracture Applications”. In: Computational Materials Science 157 (2019), pp. 87–98.
[81] H. E. Hurst. “Long-Term Storage Capacity of Reservoirs”. In: Transactions of the American Society of Civil
Engineers 116.1 (1951), pp. 770–799.
Contents 41
[82] V. John, B. Moreau, and J. Novo. “Error Analysis of a SUPG-stabilized POD-ROM Method for Convection-
Diffusion-Reaction Equations”. In: Computers & Mathematics with Applications 122 (2022), pp. 48–60.
[83] L. John Leask. “The Structure of Inhomogeneous Turbulent Flows”. In: The structure of inhomogeneous
turbulent flows (1967), pp. 166–178.
[84] J.-N. Juang. Applied System Identification. Prentice Hall, 1994.
[85] K. Kaheman, S. L. Brunton, and J. N. Kutz. “Automatic Differentiation to Simultaneously Identify Nonlin-
ear Dynamics and Extract Noise Probability Distributions from Data”. In: Machine Learning: Science and
Technology 3.1 (2022), p. 015031.
[86] E. Kaiser et al. “Cluster-Based Reduced-Order Modelling of a Mixing Layer”. In: Journal of Fluid Mechanics
754 (2014), pp. 365–414.
[87] I. Kalashnikova and M. Barone. “Stable and Efficient Galerkin Reduced Order Models for Non-Linear Fluid
Flow”. In: 6th AIAA Theoretical Fluid Mechanics Conference. 2011, p. 3110.
[88] M. G. Kapteyn, D. J. Knezevic, and K. Willcox. “Toward Predictive Digital Twins via Component-Based
Reduced-Order Models and Interpretable Machine Learning”. In: AIAA Scitech 2020 Forum. AIAA SciTech
Forum. American Institute of Aeronautics and Astronautics, 2020.
[89] M. Kast, M. Guo, and J. S. Hesthaven. “A Non-Intrusive Multifidelity Method for the Reduced Order Modeling
of Nonlinear Problems”. In: Computer Methods in Applied Mechanics and Engineering 364 (2020), p. 112947.
[90] D. P. Kingma and M. Welling. “Auto-Encoding Variational Bayes”. In: International Conference on Learning
Representations. 2013.
[91] K. Lee and K. Carlberg. “Deep Conservation: A Latent-Dynamics Model for Exact Satisfaction of Physical
Conservation Laws”. In: arXiv preprint arXiv:1909.09754 (2020). arXiv: 1909.09754 [physics].
[92] K. Lee and K. T. Carlberg. “Model Reduction of Dynamical Systems on Nonlinear Manifolds Using Deep
Convolutional Autoencoders”. In: Journal of Computational Physics 404 (2020), p. 108973.
[93] P. LeGresley and J. Alonso. “Airfoil Design Optimization Using Reduced Order Models Based on Proper
Orthogonal Decomposition”. In: Fluids 2000 Conference and Exhibit. American Institute of Aeronautics and
Astronautics, 2000.
[94] J. Li, X. Du, and J. R. R. A. Martins. “Machine Learning in Aerodynamic Shape Optimization”. In: Progress
in Aerospace Sciences 134 (2022), p. 100849.
[95] L. Ljung. System Identification: Theory for the User. 2nd edition. Upper Saddle River, NJ: Pearson, 1998.
[96] J.-C. Loiseau. “Data-Driven Modeling of the Chaotic Thermal Convection in an Annular Thermosyphon”. In:
Theoretical and Computational Fluid Dynamics 34 (2020), pp. 339–365.
[97] J.-C. Loiseau and S. L. Brunton. “Constrained Sparse Galerkin Regression”. In: Journal of Fluid Mechanics
838 (2018), pp. 42–67.
[98] D. J. Lucia and P. S. Beran. “Projection Methods for Reduced Order Models of Compressible Flows”. In:
Journal of Computational Physics 188.1 (2003), pp. 252–280.
[99] B. Lusch, J. N. Kutz, and S. L. Brunton. “Deep Learning for Universal Linear Embeddings of Nonlinear
Dynamics”. In: Nature Communications 9.1 (2018), p. 4950.
[100] C. Ma, J. Wang, and W. E. “Model Reduction with Memory and the Machine Learning of Dynamical Systems”.
In: Communications in Computational Physics 25.4 (2019).
[101] R. Maulik et al. “Latent-Space Time Evolution of Non-Intrusive Reduced-Order Models Using Gaussian
Process Emulation”. In: Physica D: Nonlinear Phenomena 416 (2021), p. 132797.
[102] M. Milano and P. Koumoutsakos. “Neural Network Modeling for Near Wall Turbulent Flow”. In: Journal of
Computational Physics 182.1 (2002), pp. 1–26.
[103] A. T. Mohan and D. V. Gaitonde. “A Deep Learning Based Approach to Reduced Order Modeling for Turbulent
Flow Control Using LSTM Neural Networks”. In: arXiv preprint arXiv:1804.09269 (2018). arXiv: 1804.09269
[physics].
[104] A. T. Mohan et al. “Embedding Hard Physical Constraints in Neural Network Coarse-Graining of 3D Turbu-
lence”. In: arXiv preprint arXiv:2002.00021 (2020). arXiv: 2002.00021 [physics].
[105] M. Mohebujjaman, L. Rebholz, and T. Iliescu. “Physically Constrained Data-Driven Correction for Reduced-
Order Modeling of Fluid Flows”. In: International Journal for Numerical Methods in Fluids 89.3 (2019),
pp. 103–122.
[106] H. Mori. “Transport, Collective Motion, and Brownian Motion*)”. In: Progress of Theoretical Physics 33.3
(1965), pp. 423–455.
[107] C. Mou et al. “Data-Driven Variational Multiscale Reduced Order Models”. In: Computer Methods in Applied
Mechanics and Engineering 373 (2021), p. 113470.
[108] T. Murata, K. Fukami, and K. Fukagata. “Nonlinear Mode Decomposition with Convolutional Neural Networks
for Fluid Dynamics”. In: Journal of Fluid Mechanics 882 (2020), A13.
[109] B. R. Noack et al. “A Hierarchy of Low-Dimensional Models for the Transient and Post-Transient Cylinder
Wake”. In: Journal of Fluid Mechanics 497 (2003), pp. 335–363.
[110] B. R. Noack et al. “Recursive Dynamic Mode Decomposition of Transient and Post-Transient Wake Flows”. In:
Journal of Fluid Mechanics 809 (2016), pp. 843–872.
[111] B. R. Noack et al., eds. Reduced-Order Modelling for Flow Control. Vol. 528. CISM International Centre for
Mechanical Sciences. Vienna: Springer, 2011.
[112] S. E. Otto and C. W. Rowley. “Linearly Recurrent Autoencoder Networks for Learning Dynamics”. In: SIAM
Journal on Applied Dynamical Systems 18.1 (2019), pp. 558–593.
[113] P. Pacciarini and G. Rozza. “Stabilized Reduced Basis Method for Parametrized Advection–Diffusion PDEs”.
In: Computer Methods in Applied Mechanics and Engineering 274 (2014), pp. 1–18.
42 Contents
[114] S. Pawar et al. “A Deep Learning Enabler for Nonintrusive Reduced Order Modeling of Fluid Flows”. In:
Physics of Fluids 31.8 (2019), p. 085101.
[115] S. Pawar et al. “Model Fusion with Physics-Guided Machine Learning: Projection-based Reduced-Order Mod-
eling”. In: Physics of Fluids 33.6 (2021), p. 067123.
[116] M. Raissi, P. Perdikaris, and G. E. Karniadakis. “Physics-Informed Neural Networks: A Deep Learning Frame-
work for Solving Forward and Inverse Problems Involving Nonlinear Partial Differential Equations”. In: Journal
of Computational Physics 378 (2019), pp. 686–707.
[117] C. E. Rasmussen and C. K. I. Williams. Gaussian Processes for Machine Learning. 2005.
[118] R. Reyes and R. Codina. “Projection-Based Reduced Order Models for Flow Problems: A Variational Multiscale
Approach”. In: Computer Methods in Applied Mechanics and Engineering 363 (2020), p. 112844.
[119] R. Reyes et al. “Reduced Order Models for Thermally Coupled Low Mach Flows”. In: Advanced Modeling and
Simulation in Engineering Sciences 5.1 (2018), pp. 1–20.
[120] G. Rozza, D. B. P. Huynh, and A. Manzoni. “Reduced Basis Approximation and a Posteriori Error Estimation
for Stokes Flows in Parametrized Geometries: Roles of the Inf-Sup Stability Constants”. In: Rozza (2013).
[121] G. Rozza, T. Lassila, and A. Manzoni. “Reduced Basis Approximation for Shape Optimization in Thermal
Flows with a Parametrized Polynomial Geometric Map”. In: Spectral and High Order Methods for Partial
Differential Equations. Springer, 2011, pp. 307–315.
[122] S. Sahba et al. “Dynamic Mode Decomposition for Aero-Optic Wavefront Characterization”. In: Optical
Engineering 61.1 (2022), p. 013105.
[123] O. San and T. Iliescu. “A Stabilized Proper Orthogonal Decomposition Reduced-Order Model for Large Scale
Quasigeostrophic Ocean Circulation”. In: Advances in Computational Mathematics 41.5 (2015), pp. 1289–
1319.
[124] O. San and R. Maulik. “Extreme Learning Machine for Reduced Order Modeling of Turbulent Geophysical
Flows”. In: Physical Review E 97.4 (2018), p. 42322.
[125] O. San and R. Maulik. “Neural Network Closures for Nonlinear Model Order Reduction”. In: Advances in
Computational Mathematics 44.6 (2018), pp. 1717–1750.
[126] O. San, S. Pawar, and A. Rasheed. “Variational Multiscale Reinforcement Learning for Discovering Reduced
Order Closure Models of Nonlinear Spatiotemporal Transport Systems”. In: arXiv preprint arXiv:2207.12854
(2022). arXiv: 2207.12854 [physics].
[127] P. J. Schmid. “Dynamic Mode Decomposition of Numerical and Experimental Data”. In: Journal of Fluid
Mechanics 656 (2010), pp. 5–28.
[128] P. J. Schmid, D. Violato, and F. Scarano. “Decomposition of Time-Resolved Tomographic PIV”. In: Experiments
in Fluids 52.6 (2012), pp. 1567–1579.
[129] N. V. Shah et al. “Finite Element Based Model Order Reduction for Parametrized One-Way Coupled Steady
State Linear Thermo-Mechanical Problems”. In: Finite Elements in Analysis and Design 212 (2022), p. 103837.
[130] P. A. Srinivasan et al. “Predictions of Turbulent Shear Flows Using Deep Neural Networks”. In: Physical Review
Fluids 4.5 (2019), p. 054603.
[131] J. A. K. Suykens et al. Least Squares Support Vector Machines. WORLD SCIENTIFIC, 2002.
[132] N. Takeishi, Y. Kawahara, and T. Yairi. “Learning Koopman Invariant Subspaces for Dynamic Mode Decom-
position”. In: Proceedings of the 31st International Conference on Neural Information Processing Systems.
NIPS’17. Red Hook, NY, USA: Curran Associates Inc., 2017, pp. 1130–1140.
[133] A. Tello and R. Codina. “Field-to-Field Coupled Fluid Structure Interaction: A Reduced Order Model Study”.
In: International Journal for Numerical Methods in Engineering 122.1 (2021), pp. 53–81.
[134] A. Tello, R. Codina, and J. Baiges. “Fluid Structure Interaction by Means of Variational Multiscale Reduced
Order Models”. In: International Journal for Numerical Methods in Engineering 121.12 (2020), pp. 2601–2625.
[135] G. Tissot et al. “Model Reduction Using Dynamic Mode Decomposition”. In: Comptes Rendus Mécanique.
Flow Separation Control 342.6 (2014), pp. 410–416.
[136] J. H. Tu et al. “On Dynamic Mode Decomposition: Theory and Applications”. In: Journal of Computational
Dynamics 1.2 (Mon Dec 01 01:00:00 CET 2014), pp. 391–421.
[137] P. R. Vlachas et al. “Data-Driven Forecasting of High-Dimensional Chaotic Systems with Long Short-Term
Memory Networks”. In: Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences
474.2213 (2018), p. 20170844.
[138] Z. Y. Wan and T. P. Sapsis. “Reduced-Space Gaussian Process Regression for Data-Driven Probabilistic Forecast
of Chaotic Dynamical Systems”. In: Physica D: Nonlinear Phenomena 345 (2017), pp. 40–55.
[139] Q. Wang, J. S. Hesthaven, and D. Ray. “Non-Intrusive Reduced Order Modeling of Unsteady Flows Using
Artificial Neural Networks with Application to a Combustion Problem”. In: Journal of computational physics
384 (2019), pp. 289–307.
[140] Q. Wang, N. Ripamonti, and J. S. Hesthaven. “Recurrent Neural Network Closure of Parametric POD-Galerkin
Reduced-Order Models Based on the Mori-Zwanzig Formalism”. In: Journal of Computational Physics 410
(2020), p. 109402.
[141] Z. Wang et al. “Proper Orthogonal Decomposition Closure Models for Turbulent Flows: A Numerical Compar-
ison”. In: Computer Methods in Applied Mechanics and Engineering 237 (2012), pp. 10–26.
[142] C. Wehmeyer and F. Noé. “Time-Lagged Autoencoders: Deep Learning of Slow Collective Variables for
Molecular Kinetics”. In: The Journal of Chemical Physics 148.24 (2018), p. 241703.
[143] M. O. Williams, I. G. Kevrekidis, and C. W. Rowley. “A Data–Driven Approximation of the Koopman Operator:
Extending Dynamic Mode Decomposition”. In: Journal of Nonlinear Science 25.6 (2015), pp. 1307–1346.
Contents 43
[144] X. Xie, C. Webster, and T. Iliescu. “Closure Learning for Nonlinear Model Reduction Using Deep Residual
Neural Network”. In: Fluids 5.1 (2020), p. 39.
[145] S. Xu et al. “Multi-Output Least-Squares Support Vector Regression Machines”. In: Pattern Recognition Letters
34.9 (2013), pp. 1078–1084.
[146] M. Z. Yousif and H.-C. Lim. “Reduced-Order Modeling for Turbulent Wake of a Finite Wall-Mounted Square
Cylinder Based on Artificial Neural Network”. In: Physics of Fluids 34.1 (2022), p. 015116.
[147] J. Yvonnet and Q.-C. He. “The Reduced Model Multiscale Method (R3M) for the Non-Linear Homogenization
of Hyperelastic Media at Finite Strains”. In: Journal of Computational Physics 223.1 (2007), pp. 341–368.
[148] H. Zhao. “A Reduced Order Model Based on Machine Learning for Numerical Analysis: An Application to
Geomechanics”. In: Engineering Applications of Artificial Intelligence 100 (2021), p. 104194.
[149] Q. Zhu, Y. Guo, and W. Lin. “Neural delay differential equations”. In: The International Conference on Learning
Representations. 2021, p. 20.
[150] R. Zwanzig. “Ensemble Method in the Theory of Irreversibility”. In: The Journal of Chemical Physics 33.5
(1960), pp. 1338–1341.