0% found this document useful (0 votes)
29 views29 pages

LFA MGjumped

Uploaded by

1017673438
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
29 views29 pages

LFA MGjumped

Uploaded by

1017673438
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

SIAM J. SCI. COMPUT.

c 2019 Society for Industrial and Applied Mathematics


Vol. 41, No. 3, pp. A1385–A1413

ON LOCAL FOURIER ANALYSIS OF MULTIGRID METHODS FOR


PDEs WITH JUMPING AND RANDOM COEFFICIENTS∗
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

PRASHANT KUMAR† , CARMEN RODRIGO‡ , FRANCISCO J. GASPAR† ,


AND CORNELIS W. OOSTERLEE§

Abstract. In this paper, we propose a novel nonstandard local Fourier analysis (LFA) variant for
accurately predicting the multigrid convergence of problems with random and jumping coefficients.
This LFA method is based on a specific basis of the Fourier space rather than the commonly used
Fourier modes. To show the utility of this analysis, we consider, as an example, a simple cell-
centered multigrid method for solving a steady-state single phase flow problem in a random porous
medium. We successfully demonstrate the predictive capability of the proposed LFA using a number
of challenging benchmark problems. The information provided by this analysis could be used to
estimate a priori the time needed for solving certain uncertainty quantification problems by means
of a multigrid multilevel Monte Carlo method.

Key words. PDEs, random coefficients, multigrid, local Fourier analysis, multilevel Monte
Carlo, uncertainty quantification

AMS subject classifications. 65F10, 65M22, 65M55

DOI. 10.1137/18M1173769

1. Introduction. A number of problems in science and engineering involve solv-


ing partial differential equations (PDEs) with random parameters or coefficients. The
solution to such PDEs is of a stochastic nature, and the aim is to compute expected
values and corresponding variances of a functional of the solution. For such uncer-
tainty quantification (UQ) problems, Monte Carlo (MC) type methods are preferred
due to their dimension-independent convergence. For any sampling-based approach,
the availability of a highly efficient and robust (w.r.t. the random inputs) iterative
solver becomes critical. In general, the samplewise computational cost can become
highly heterogeneous, depending on the random inputs. Therefore, if the performance
statistics of such solvers were known a priori, one could utilize this information to op-
timize and parallelize the MC simulations efficiently.
In this paper, we present a nonstandard local Fourier analysis (LFA) technique
to predict the convergence rate of multigrid solvers for problems involving random
and jumping coefficients. Standard LFA techniques are typically based on constant
coefficient discretization stencils, whereas for stochastic PDEs we encounter varying
coefficients throughout the computational domain, due to the randomness. One of
the main contributions of this work is to generalize the LFA toward problems with

∗ Submitted to the journal’s Methods and Algorithms for Scientific Computing section March 5,

2018; accepted for publication (in revised form) March 19, 2019; published electronically May 2,
2019.
http://www.siam.org/journals/sisc/41-3/M117376.html
Funding: The work of the second author was supported by the Spanish project FEDER/MCYT
MTM2016-75139-R and the Diputación General de Aragón (Grupo de referencia APEDIF, ref.
E24 17R). The work of the third author was supported by the European Union’s Horizon 2020 re-
search and innovation program under Marie Sklodowska-Curie grant agreement 705402, POROSOS.
† CWI, Centrum Wiskunde and Informatica, Amsterdam, The Netherlands ([email protected],

[email protected], http://www.unizar.es/pde/fjgaspar/).
‡ IUMA and Applied Mathematics Department, University of Zaragoza, Zaragoza, Spain

([email protected]).
§ CWI, Centrum Wiskunde and Informatica, Amsterdam, The Netherlands, and DIAM, Delft

University of Technology, The Netherlands ([email protected]).


A1385

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1386 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

random and jumping coefficients, with the aim of predicting, a priori, the total time
needed to solve UQ problems. Some efforts have already been made in [1] regarding
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

the generalization of LFA for jumping coefficients. The novelty of our approach lies
in the choice of basis functions. Here, we utilize a new basis from the Fourier space
rather than the standard Fourier modes. We benchmark the prediction capability of
the proposed LFA technique using a set of challenging jumping coefficient problems
and a number of spatially correlated random fields with varying heterogeneity.
One of the results of this paper is the ability of the aforementioned LFA technique
to predict accurately the multigrid convergence rates for elliptic PDEs with hetero-
geneous random coefficients, as encountered in the stochastic Darcy flow problem.
In stochastic subsurface flow modeling, a lognormal random field often represents the
permeability of an occurring heterogeneous porous medium [2, 3, 4]. Uncertainty may
then be assessed by means of stochastic collocation approaches [5, 6], where the use of
multigrid has been analyzed in [7]. In the current paper, we are interested in solving
the PDE in the context of the multilevel Monte Carlo (MLMC) technique [8, 9, 10].
Within MC methods, samples of the stochastic random field are generated and the
corresponding solution of the PDE with “frozen stochastic coefficients” is computed.
Plain MC methods may require many thousands of random samples on a fine compu-
tational mesh, before the mean and variance of the numerical solution stabilize and
converge. This may cost substantial CPU-time. MLMC methods are based on the de-
composition of the expectation and variance operators into sequences of differences of
these quantities on fine and coarse problem scales. The number of iterations required
on each scale is different, where fewer samples (fewer random fields) are typically
required on the finer scales, and larger numbers of samples are needed on the coarser,
cheaper to compute, scales.
Combining MLMC methods with multigrid seems natural, where the multigrid
method is employed for the numerical solution of a PDE which is based on a sample
of the stochastic quantity on a certain (fine or coarse) scale. Due to the stochasticity,
however, we deal with PDE problems with jumping coefficients, where different jump
patterns are encountered each time a new random field is generated. The generaliza-
tion of the LFA toward these PDEs will provide us insight into the average number
of multigrid iterations and the spread of the convergence factors, among other things.
This information helps to estimate the total CPU-time needed for the multiple multi-
grid computations in a MLMC setting. We consider it useful to construct a technique
to assess the quality of the choice of the multigrid components in the context of
the PDEs with random problems, before the actual multigrid computation has taken
place.
In this work, we will employ a basic cell-centered multigrid (CCMG) algorithm for
solving elliptic PDEs with a variable coefficients field. The components of this algo-
rithm include a simple Gauss–Seidel iteration as the smoother, a piecewise constant
prolongation operator and its adjoint as the restriction, and a direct discretization
technique to define the discrete operators on the coarse grids. We show that for this
special combination, the coarse-grid discretization operators are equivalent to the
ones obtained from commonly used Galerkin operators [11]. We utilize this CCMG
method to perform MLMC simulations with different permeability parameters. It
may be surprising that such a basic algorithm converges well in the context of the
generated random fields, where computation takes place for thousands of different
samples. Although we restrict ourselves to this basic CCMG to demonstrate the ac-
curacy of the predictions of the novel LFA technique, we emphasize that this approach
can be used for a wider range of problems, discretizations, and multigrid methods.
The proposed LFA technique allows us to deal with several challenging problems for

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1387

which it is not easily possible to apply the classical LFA.


The paper is organized as follows. In section 2 we introduce the context of PDEs
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

with jumping and random coefficients, together with their discretization by a cell-
centered finite volume scheme. A discussion on multigrid methods for this type of
problem is also included, and the multigrid components that will be considered in this
work are defined. Section 3 is devoted to the generalization of the LFA to deal with
jumping coefficients and problems with random fields. In section 4, we present results
obtained by this analysis for different benchmark problems with jumping coefficients.
Section 5 presents the LFA results for problems with random coefficients, and in
section 6 MLMC computations for PDEs with random coefficients are presented.
Finally, in section 7 conclusions are drawn.
2. Jumping coefficients, random coefficients, multigrid. Robust and effi-
cient iterative solution methods are very relevant for PDEs with variable coefficients.
For PDEs with jumping coefficients, multigrid methods have already been shown to
be this type of solver. When using MC methods, in the case of elliptic PDEs with
random coefficient fields, many samples of the random field are generated and for each
field the numerical solution should be computed. This can take substantial CPU-time
if very many samples are required. For a fixed sample of the random field, we deal with
an elliptic PDE with varying coefficients, due to the randomness. Multigrid comes in
naturally as a highly efficient solution method for the resulting PDEs. In this section,
we explain this setting and we briefly describe an efficient multigrid method based on
a cell-centered grid and a finite volume discretization.
2.1. PDEs with jumping and with random coefficients. We start with
classical PDE problems with jumping coefficients. In particular, we deal with the
following two-dimensional diffusion equation on the square domain D = (0, `)2 :
(1) −∇ · (k(x)∇u(x)) = f (x), x ∈ D,
(2) u(x) = g(x), x ∈ ∂D,
where k(x) is a function which may be discontinuous across internal boundaries.
To discretize this problem, we use a cell-centered finite volume method based on
the harmonic average of the diffusion coefficient k(x). We consider a uniform grid Dh
with the same step size h = `/M, M ∈ N in both directions,
(3) Dh = {(xi1 , xi2 ); xiα = (iα − 1/2)h, iα = 1, . . . , M, α = 1, 2}.

This gives, for each interior cell with center (xi1 , xi2 ), denoted by Dhi1 ,i2 , a five-point
scheme
(4) chi1 ,i2 ui1 ,i2 + wih1 ,i2 ui1 −1,i2 + ehi1 ,i2 ui1 +1,i2 + shi1 ,i2 ui1 ,i2 −1 + nhi1 ,i2 ui1 ,i2 +1 = fih1 ,i2 ,
where
2 ki1 ,i2 ki1 −1,i2 2 ki1 ,i2 ki1 +1,i2
wih1 ,i2 = − , ehi1 ,i2 = − ,
h2 ki1 ,i2 + ki1 −1,i2 h2 ki1 ,i2 + ki1 +1,i2
2 ki1 ,i2 ki1 ,i2 −1 2 ki1 ,i2 ki1 ,i2 +1
shi1 ,i2 =− 2 , nhi1 ,i2 =− 2 ,
h ki1 ,i2 + ki1 ,i2 −1 h ki1 ,i2 + ki1 ,i2 +1
chi1 ,i2 = −(wih1 ,i2 + ehi1 ,i2 + nhi1 ,i2 + shi1 ,i2 ),

with, for instance, ki1 ,i2 the diffusion coefficient associated with the cell Dhi1 ,i2 . By
interior cell we mean a cell for which none of its edges lies at the boundary of the
domain. This scheme is changed appropriately for the cells close to the boundary.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1388 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

As mentioned, we also consider elliptic PDEs with random coefficient fields. The
PDE of our interest describes the steady-state single-phase flow in a random porous
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

medium. Denoting by ω an event in the probability space (Ω, F, P), with sample space
Ω, σ-algebra F, and probability measure P, the permeability in the porous medium is
described by k(x, ω) : D × Ω → R+ . The PDE is then given by
(5) −∇ · (k(x, ω)∇u(x, ω)) = f (x), x ∈ D, ω ∈ Ω,
with f as a source term.
The engineering interest in the solution of (5) is typically found in expected values
of linear functionals of the solution u, denoted by Q := Q(u).
To discretize these problems, we use the same cell-centered finite volume method
based on the harmonic average of the random diffusion coefficient as previously de-
scribed for problems with jumping coefficients. We make the common assumption
that the permeability random field is constant over each cell of the grid.
2.2. Multigrid for PDEs with jumping and with random coefficients.
In this work, the multigrid components for the above cell-centered discrete problems
are chosen as follows. We use a lexicographic Gauss–Seidel iteration as the multigrid
smoother and we consider standard coarsening obtained by doubling the mesh size
in both directions. Each coarse cell is the union of four fine cells, and, since the
unknowns are located in the cell centers, this results in a nonnested hierarchy of
h
grids. We consider a simple prolongation operator P2h , that is, the piecewise constant
interpolation operator. In stencil notation, it is given by
 h
1 1
(6) h
P2h = ?  ,
1 1 2h
where ? denotes the position of a coarse-grid unknown. The classical stencil notation
shows the contribution of the coarse-grid node to the neighboring fine-grid nodes. The
restriction operator Rh2h is chosen as the scaled adjoint of the prolongation, given in
stencil form by
 2h
1 1
1
(7) Rh2h =  ?  .
4
1 1 h
The coarse-grid operators are constructed by direct discretization defining the diffu-
sion coefficients at the edges of the coarse cells appropriately, which we will describe
in more detail. We assume that the diffusion coefficient k(x) is piecewise constant on
the fine grid. The flux over an edge, dependent on the solution in the two adjacent
cells, is calculated based on the harmonic average. The values of the diffusion coef-
ficients at a coarse edge located between two coarse cells, however, are calculated as
the arithmetic average of the corresponding fine-grid coefficients; see Figure 1 for a
more detailed description. As pointed out in [12], this direct discretization procedure
is equivalent to the often used Galerkin approach, i.e., L2h = 21 Rh2h Lh P2h
h
, but com-
putationally more efficient. The factor 1/2 in the previous expression is due to the
lack of consistency of the operator Rh2h Lh P2h
h
with the differential operator [13]. In
the next result, we prove that both discretizations are indeed equivalent.
Proposition 1. Let Lh be the fine grid operator based on the cell-centered finite
volume discretization of problem (5) on a uniform grid of mesh size h = `/M with

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1389
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

AM AM

HM

h 2h 4h

Fig. 1. Schematic representation of permeability upscaling used in the multigrid hierarchy (h-
2h-4h). (Top) Permeability values generated at cell centers (blue dots). (Bottom left) Permeability
values at face centers (red dots) obtained from the harmonic mean (HM) of permeabilities from two
adjacent cell centers. (Bottom middle) Permeability at face centers (bigger red dots) of 2h-grid is the
arithmetic mean (AM) of permeabilities from face centers of the h-grid. (Bottom right) Permeability
at face centers (biggest red dots) of 4h-grid is the arithmetic mean of permeabilities from face centers
of the 2h-grid.

h
M even. Let P2h be the piecewise constant prolongation operator and Rh2h its adjoint.
Then, the Galerkin coarse-grid operator L2h = 21 Rh2h Lh P2hh
is equivalent to a direct
discretization on the coarse-grid based on the arithmetic average of the corresponding
fine-grid coefficients.
i1 ,i2
Proof. We prove the equivalence for a coarse-grid cell D2h such that none of its
edges lies on the boundary of the domain. The equivalence for coarse cells close to
boundaries with Dirichlet or Neumann boundary conditions can be proven similarly.
By applying the restriction operator Rh2h in (7), the equation associated with the cell
i1 ,i2
D2h by using the Galerkin approach is given by
1 2h h 1 h h h
(R Lh P2h u)i1 ,i2 = ((Lh P2h u)2i1 ,2i2 + (Lh P2h u)2i1 −1,2i2 + (Lh P2h u)2i1 ,2i2 −1
2 h 8
h
(8) + (Lh P2h u)2i1 −1,2i2 −1 ).
Taking into account that the prolongation operator is piecewise constant, we obtain
h
(Lh P2h u)2i1 ,2i2 = eh2i1 ,2i2 ui1 +1,i2 + w2i
h
u
1 ,2i2 i1 ,i2
+ nh2i1 ,2i2 ui1 ,i2 +1
+ sh2i1 ,2i2 ui1 ,i2 + ch2i1 ,2i2 ui1 ,i2
and similar expressions for the other terms in (8). By substituting these expressions
i1 ,i2
in (8), the following discretization for the coarse cell D2h is obtained:
1 2h h
(R Lh P2h u)i1 ,i2 = c2h 2h 2h
i1 ,i2 ui1 ,i2 + wi1 ,i2 ui1 −1,i2 + ei1 ,i2 ui1 +1,i2
2 h
(9) + s2h 2h
i1 ,i2 ui1 ,i2 −1 + ni1 ,i2 ui1 ,i2 +1 ,

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1390 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

where
1 h  1 h 
wi2h = w2i1 −1,2i2 + w2i h
, e2hi1 ,i2 = e2i1 ,2i2 + eh2i1 ,2i2 −1 ,
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

1 ,i2 1 −1,2i2 −1
8 8
1 h  1 h 
s2h = s h 2h
+ s2i1 −1,2i2 −1 , ni1 ,i2 = n + nh2i1 ,2i2 ,
i1 ,i2
8 2i1 ,2i2 −1 8 2i1 −1,2i2
c2h
i1 ,i2 = −(wi2h
1 ,i2
+ e2h 2h 2h
i1 ,i2 + ni1 ,i2 + si1 ,i2 ).

We observe that this scheme is equivalent to a direct discretization on the coarse


grid where the diffusion coefficients on the edges are the arithmetic averages of the
corresponding fine-grid coefficients.
Remark. In order to achieve a mesh-independent multigrid convergence following
the analysis from [14], the following condition must be satisfied:

(10) mp + mr > Mpde ,

where the orders mp and mr are the highest degree plus one of the polynomials that
h
are exactly interpolated by P2h and Rh2h , respectively, and Mpde is the order of the
PDE to be solved. For PDE (5), we have Mpde = 2, and for the considered operators
(6) and (7), we get mp = mr = 1, which does not satisfy the inequality (10). In [15] it
is shown, however, that this condition is not needed to prove uniform convergence [16].
2.3. Discussion about other multigrid methods for jumping coefficients.
In the context of algebraic multigrid methods for the numerical solution of PDEs, basi-
cally two prevailing methods have proved their use for multiple engineering problems,
i.e., algebraic multigrid and aggregation-based multigrid methods [17, 18, 19, 20, 21].
These methods converge remarkably well, for example, for scalar PDEs with jumping
coefficients. It is not always easily understood why these methods, and particularly
the aggregation-based method, converge so well.
The origin of these algebraic methods may be found in the early days of multigrid,
where a black-box multigrid with operator-dependent transfer operators (restriction
and prolongation) and Galerkin coarse-grid operators for structured vertex-centered
Cartesian grids was proposed in [22, 23, 24]. This can be seen as a predecessor
of classical AMG, where these components were essentially enhanced by a flexible
coarsening strategy.
The aggregation-based multigrid methods, with their origin in the work by Man-
del [21] (smoothed aggregation), may be related to the CCMG methods as proposed
in [25, 26]. In [26], it was shown that constant, i.e., operator-independent, transfer
operators, in combination with Galerkin coarse-grid discretization, provided highly
efficient multigrid results for cell-centered discretizations of elliptic PDEs that in-
cluded jumping coefficients. These CCMG components were augmented with robust
smoothing, like incomplete lower-upper decomposition (ILU) relaxation. The individ-
ual contributions of the coarse-grid correction and the smoothing parts were difficult
to distinguish. Also, a CCMG based on coarsening by a factor of three together with
operator-dependent interpolations was explored in [27].
3. Local Fourier analysis for variable coefficients. In this section we de-
scribe LFA in a setting which allows us to estimate the multigrid convergence factors
for problems with jumping coefficients and problems with random fields. A discrete
linear operator with constant coefficients, which is formally defined on an infinite grid,
is usually assumed for carrying out a standard LFA. As we will show, this assump-
tion can be relaxed by considering a discrete operator with constant coefficients in

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1391
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Fig. 2. Infinite grid Dh divided into the corresponding subgrids for n = 2.

appropriate infinite subgrids. This allows us to generalize the analysis to problems for
which the discrete operator consists of different stencils. A key point in this improved
analysis is to consider a specific basis of the Fourier space, rather than the standard
basis which is based on the Fourier modes. The use of this new basis will simplify the
analysis.
We start from a regular infinite grid Dh with grid size h in both directions.
Such an infinite grid will be split into n × n subgrids in the following way. First of
all, a window comprising n × n cells of the original grid is adequately chosen, and,
subsequently, we consider its periodic extension. The choice of the size of the n × n
window is made such that the variability of the discrete operator in the computational
grid can be appropriately represented, as will be explained by means of examples of
different nature. Once n is fixed, the infinite subgrids are defined as follows (see
Figure 2 for an example with n = 2),

(11) Dhkl = {(k, l)h + (nk1 , nk2 )h | k1 , k2 ∈ Z} , k, l = 0, . . . , n − 1.

For each low frequency, θ 00 ∈ Θnh = (−π/nh, π/nh]2 , we introduce the grid-
functions:

(12) ψhkl (θ 00 , x) = ϕh (θ 00 , x)χDhkl (x), k, l = 0, . . . , n − 1, x ∈ Dh ,

00
where ϕh (θ 00 , x) = eıθ ·x is the standard Fourier mode on Dh corresponding to the
frequency θ 00 . It is easy to see that the subspace generated by these n2 grid-functions,
2
(13) Fhn (θ 00 ) = span{ψhkl (θ 00 , ·), k, l = 0, . . . , n − 1},

is the same as the one spanned by the n2 Fourier modes ϕh (θ 00


kl , ·) associated with
the frequencies:


(14) θ 00
kl = θ
00
+ (k, l) , k, l = 0, . . . , n − 1.
nh

In the case n = 2, the basis {ψh00 (θ 00 , ·), ψh11 (θ 00 , ·), ψh10 (θ 00 , ·), ψh01 (θ 00 , ·)} is related
to the standard basis of Fourier modes {ϕh (θ 00 , ·), ϕh (θ 11 , ·), ϕh (θ 10 , ·), ϕh (θ 01 , ·)} in
the following way:

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1392 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

(15)
 00 00      
ψh (θ , ·) ϕh (θ 00 , ·) 1 1 1 1 ϕh (θ 00 , ·)
 ψ 11 (θ 00 , ·)   11    
−1   ϕh (θ 11 , ·) 
 = M  ϕh (θ 10 , ·)  = 1  1 1 −1
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

 h 00 .
 ψ 10 (θ , ·)   ϕh (θ , ·)  4  1 −1 −1 1   ϕh (θ 10 , ·) 
h
ψh01 (θ 00 , ·) ϕh (θ 01 , ·) 1 −1 1 −1 ϕh (θ 01 , ·)
It is well-known that Fourier modes are eigenfunctions of any constant coefficient
linear discrete operator Lh , that is, Lh ϕh (θ, x) = L e h (θ)ϕh (θ, x). Therefore, the
2
representation of Lh with respect to a basis of n Fourier modes is a diagonal matrix
with diagonal elements L e h (θ kl ) with k, l = 0, . . . , n − 1. In general, the Fourier
representation with respect to the basis of functions {ψhkl }n−1 k,l=0 is a dense matrix. We
b 00
will denote it by Lh (θ ).
If we consider the five-point standard discretization of the Laplace operator on a
uniform grid of mesh size h,
 
−1
1 
(16) −1 4 −1  ,
h2
−1
its Fourier symbol with respect to the standard basis of Fourier modes is a diagonal
matrix with diagonal elements equal to
1
(4 − 2 cos(θxkl ) − 2 cos(θykl ))
h2
(see [28], for instance), whereas the Fourier representation with respect to the new
basis in the case n = 2 is given by
(17)
 
2 0 cos(θx00 ) cos(θy00 )
2  0 2 cos(θy00 ) cos(θx00 ) 
b h (θ ) =
L 00   with θ 00 = (θx00 , θy00 ).
h2  cos(θx00 ) cos(θy00 ) 2 0 
cos(θy00 ) cos(θx00 ) 0 2
Notice that, for example, the first row of the previous symbol is obtained by looking
at the decomposition of the stencil (16) into the connections among the unknowns
located at the different subgrids Dhkl defined in (11). In particular, following the
notation in Figure 2, the •-•, • −  and • − ◦ connections are given by the stencils
 
  −1
1 1 1 
[4] , −1 • −1 , • ,
h2 h2 h2
−1
giving rise to the symbols
4 2 2
, cos(θx00 ), cos(θy00 ),
h2 h2 h2
which appear in the first row of (17), whereas there is no • − × connection. The rest
of the rows are analogously computed.
The procedure to obtain the Fourier symbol of a smoothing operator Sh , which

is based on a splitting of the discrete operator Lh = L+h + Lh , is analogous with the
new basis. The smoothing iteration is given by

L+
h w h + Lh wh = fh

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1393

with wh the approximation of the solution before the smoothing step and wh the

approximation after the smoothing step. By computing the symbols of L+
h and Lh
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

as before, the Fourier symbol of the smoothing operator is given by

(18) Sbh (θ 00 ) = −(L


b + )−1 (θ 00 )L
h
b − (θ 00 ).
h

The Fourier symbol corresponding to a lexicographic Gauss–Seidel iteration for the


five-point standard discretization of the Laplace operator on a uniform grid of mesh
size h, in the case n = 2, is as in (18), where

(19)
 
2 0 0 0
2  0 2 cos(θy00 ) cos(θx00 ) 
b + (θ 00 ) =
L  b − (θ 00 ) = L
, L b h (θ 00 ) − L
b + (θ 00 ).
h
h2  cos(θ 00
x ) 0 2 0  h h

cos(θy00 ) 0 0 2

Once the Fourier representation of the smoothing operator with respect to the new
basis is obtained, we can define the smoothing factor by using the change of basis
matrix. For example, for the case n = 2, the smoothing factor is obtained by

µ(Sh ) = sup ρ(Qh MSbh (θ 00 )M−1 ),


θ 00 ∈Θ2h

where Qh is the projection operator onto the space of high-frequency components and
M is the change of basis matrix given in (15).
3.1. LFA formulations for cell-centered grids. In section 2, we didn’t dis-
tinguish between cell- and vertex-centered grids. The generalized LFA indeed works
well for both types of discretization. By introducing the coarse grids and their relation
with the fine grids, we need to fix the approach of interest. Since here we will focus
on cell-centered discretizations, from now on the description of the analysis will be
given for this case, although it may be applied to the vertex-centered case in a similar
way by defining appropriately the coarse meshes.
According to the location of the coarse-grid points in a regular cell-centered grid,
we define for a fixed n the following infinite coarse subgrids of D2h :
kl
(20) D2h = {(h/2, h/2) + (k, l)2h + (nk1 , nk2 )h | k1 , k2 ∈ Z} , k, l = 0, . . . , n/2 − 1.

Due to the relation between the grid-functions of the new Fourier basis given in (12)
and the standard Fourier modes, it can be shown that the coarse-grid correction
h −1 2h h
operator Ch = Ih − P2h L2h Rh Lh , where P2h and Rh2h are the prolongation and
restriction operators, Lh and L2h are the fine- and coarse-grid operators, and Ih is
the identity, satisfies the following invariance property:
2 2
Ch : Fhn (θ 00 ) → Fhn (θ 00 ).

More concretely, for θ 00 ∈ Θ2nh = (−π/2nh, π/2nh]2 , the following properties of the
operators in Ch are fulfilled:
2 2
1. Lh , Ih : Fhn (θ 00 ) → Fhn (θ 00 ),
n2 /4 n2 /4
2. L2h : F2h (2θ 00 ) → F2h (2θ 00 ),
2 n2 /4
3. Rh2h : Fhn (θ 00 ) → F2h (2θ 00 ),
2
n /4 2
h
4. P2h : F2h (2θ 00 ) → Fhn (θ 00 ).

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1394 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

From these invariance properties we can compute the Fourier representations of the
corresponding operators. As an example, we will describe next the representation of
n/2−1
(2θ 00 )}k,l=0 and {ψhkl (θ 00 )}n−1
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Rh2h with respect to the grid-functions {ψ2hkl


k,l=0 , for the
restriction operators considered in this work.
We first consider the basic restriction operator obtained as the adjoint of the piece-
wise constant prolongation operator with stencil form (7). Its Fourier representation
with respect to the new Fourier basis is given by
 
b2h (θ 00 ) = 1 e−ı(θx00 +θy00 )/2
R eı(θx
00
+θy00 )/2 00
eı(θx −θy00 )/2 00
eı(θy
00
−θx )/2 .
h
4
In the case of the cell-centered restriction operator by Khalil and Wesseling [25], that
is,
 2h
1 1 0 0
 1 3 2 0 
1 


 ,
(21) Rh2h = ?
16 
 0

2 3 1 
0 0 1 1 h

the Fourier representation is given by


 00 00 00 00 T
e−ı(θx +θy )/2 (2 + e2ıθy + e2ıθx )
 00 00 00 00 
bh2h (θ 00 ) = 1  eı(θx +θy )/2 (2 + e−2ıθy + e−2ıθx ) 
R  00 00 00 00  .
16  eı(θx −θy )/2 (3 + e2ı(θy −θx ) 
00 00 00 00
eı(θy −θx )/2 (3 + e2ı(θx −θy ) ).

As an immediate consequence of these invariance properties and the invariance prop-


erty of the smoothing operator, also the two-grid operator Kh2h = Ch Shν , where ν
2
denotes the number of smoothing steps, leaves the subspaces Fhn (θ 00 ) invariant. Its
Fourier representation is given by
b h2h (θ 00 ) = C
K bh (θ 00 )Sbhν (θ 00 ) = (Ibh (θ 00 )−Pb2h
h b −1 (θ 00 )R
(θ 00 )L bh2h (θ 00 )L
b h (θ 00 ))Sbhν (θ 00 ).
2h

Finally, we can compute the asymptotic two-grid convergence factor as the supremum
of the spectral radii of (n2 × n2 )-matrices, as follows:

ρ(Kh2h ) = sup b 2h (θ 00 )),


ρ(Kh
θ 00 ∈Θ
e 2nh

where Θe 2nh is the subset of Θ2nh in which we remove the frequencies θ 00 such that
the determinant of the Fourier symbol of Lh or L2h vanishes.
4. LFA results for PDEs with jumping coefficients. In this section, we
apply the proposed LFA to predict the two-grid convergence factors for a collection of
benchmark problems with jumping coefficients taken from the literature [22, 29, 30].
The test cases cover a variety of possible inhomogeneities including jumps that are
not aligned with the coarse grid. In all these problems, (1) is numerically solved in
the domain D = (0, 1)2 by using a mesh of grid-size h = 1/128. In particular, the
following jumping coefficient benchmark problems, characterized by the distribution
of the diffusion coefficient, are considered here:

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1395

1. Vertical jump. Function k(x, y) is defined in the following way (see also Figure
3(a)): (
if x < 12 + h,
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

1
k(x, y) =
103 if x ≥ 12 + h.
2. Four corner problem. The domain is divided into four regions in which the
diffusion coefficient is varying; see Figure 3(b). In particular,
 1
2
4
 10 if (x, y) ∈ 0, 2 + h ,



 1  
if (x, y) ∈ 0, 12 + h × 12 + h, 1 ,
k(x, y) =  

 10−2 if (x, y) ∈ 12 + h, 1 × 21 + h, 1 ,



10−4 otherwise.
3. Square inclusion. In this example we assume a square inhomogeneity in one
cell within the square domain; see Figure 3(c). The diffusion coefficient is
defined as ( 2
k0 if (x, y) ∈ 21 − h, 12 ,
k(x, y) =
1 otherwise,
where values k0 = 104 and k0 = 10−4 are considered.
4. Periodic square inclusions. This test is taken from [31]. We consider a struc-
tured pattern of square inclusions of size 2h × 2h as depicted in
Figure 3(d). The diffusion parameter is k(x, y) = 1 inside the dark region
and k(x, y) = 1000 inside the white region.
5. Periodic L-shaped inclusions. In the last test case, we consider a structured
pattern of L-shaped inclusions as in Figure 3(e). The diffusion parameter is
k(x, y) = 104 inside the white region and k(x, y) = 1 inside the dark region.
To perform the theoretical analysis, the periodic extension of a window of size 8×8 has
been chosen, where the diffusion coefficient is prescribed in such a window according
to its definition; see Figure 3 (right side). In all numerical tests a random initial guess
is chosen, and the right-hand side and boundary conditions are set to zero to be able
to determine asymptotic convergence factors. In this way, we avoid round-off errors,
permitting us to perform as many iterations as needed. In practice, we have seen that
50 iterations are sufficient.
Next, we show the excellent correspondence between the theoretical analysis and
the experimental results for these test cases. Two combinations of intergrid transfer
operators are considered. The first combination, denoted here by (CP,CR), is based
on the use of piecewise constant prolongation (6) and its adjoint as the restriction
(7). In the second combination we change to a higher polynomial order interpolation
operator which is the adjoint to the Wesseling–Khalil restriction (21). This choice is
denoted by (WP,CR). Moreover, a standard damped Jacobi smoother (damping with
ω = 0.8) is considered as well as the proposed lexicographic Gauss–Seidel smoother.
In Table 1, for different numbers of smoothing steps, for two different smoothers,
and for the two combinations of restriction and prolongation operators, we provide
the two-grid convergence factors predicted by the novel LFA for each of the pro-
posed numerical experiments. We also display in parentheses the average value after
50 iterations of the experimentally computed convergence factors by using two-grid
multigrid cycles. For all these cases, we observe a very accurate match between the
analysis results and the rates experimentally obtained. Regarding the size of the win-
dow to perform the LFA, we have observed that a window of size 8 × 8 is enough to

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1396 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

Multigrid LFA
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

(a) Vertical jump


(b) Four corner problem
(c) Square inclusions
(d) Periodic sq. inclusions
(e) Periodic L-s. inclusions

Fig. 3. Distribution of the diffusion coefficients for the five considered examples on a unit
square domain and corresponding 8 × 8 window used in the LFA. L-s. = L-shaped.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1397
Table 1
Asymptotic two-grid convergence factors predicted by LFA and the corresponding computed
average multigrid convergence factors (in parentheses) using two-grid cycles and different pre- and
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

postsmoothing steps for the five examples. L-s. = L-shaped; GS = Gauss–Seidel.

Pre-, postsmoothing steps


(1, 0) (1, 1) (2, 2)
(CP,CR) 0.42(0.42) 0.18(0.19) 0.04(0.03)
GS
(WP,CR) 0.41(0.37) 0.19(0.16) 0.07(0.11)
Vertical jump
(CP,CR) 0.65(0.65) 0.43(0.42) 0.19(0.19)
Jacobi
(WP,CR) 0.63(0.59) 0.40(0.35) 0.19(0.19)
(CP,CR) 0.42(0.37) 0.15(0.12) 0.04(0.03)
GS
(WP,CR) 0.40(0.39) 0.16(0.16) 0.09(0.09)
Four corner problem
(CP,CR) 0.63(0.61) 0.40(0.39) 0.16(0.16)
Jacobi
(WP,CR) 0.62(0.62) 0.40(0.40) 0.19(0.18)
(CP,CR) 0.45(0.44) 0.21(0.19) 0.04(0.04)
GS
(WP,CR) 0.41(0.40) 0.18(0.17) 0.11(0.11)
Sq. inclusion (k = 104 )
(CP,CR) 0.60(0.65) 0.36(0.42) 0.13(0.19)
Jacobi
(WP,CR) 0.60(0.62) 0.38(0.37) 0.22(0.21)
(CP,CR) 0.46(0.45) 0.21(0.20) 0.05(0.05)
GS
(WP,CR) 0.41(0.40) 0.19(0.19) 0.12(0.12)
Sq. inclusion (k = 10−4 )
(CP,CR) 0.61(0.65) 0.38(0.42) 0.15(0.19)
Jacobi
(WP,CR) 0.61(0.59) 0.39(0.39) 0.23(0.23)
(CP,CR) 0.64(0.61) 0.43(0.42) 0.41(0.41)
GS
(WP,CR) 0.62(0.61) 0.43(0.41) 0.41(0.40)
Periodic Sq. inclusions
(CP,CR) 0.81(0.78) 0.66(0.65) 0.44(0.46)
Jacobi
(WP,CR) 0.81(0.80) 0.66(0.65) 0.44(0.43)
(CP,CR) 0.50(0.50) 0.32(0.26) 0.21(0.21)
GS
(WP,CR) 0.54(0.53) 0.40(0.40) 0.30(0.30)
Periodic L-s. inclusions
(CP,CR) 0.71(0.63) 0.54(0.48) 0.36(0.35)
Jacobi
(WP,CR) 0.71(0.71) 0.56(0.56) 0.42(0.42)

achieve excellent predictions in all considered benchmark problems. For example, for
the vertical jump test, the two-grid analysis considering four smoothing steps of the
Gauss–Seidel smoother and the combination (CP,CR) of intergrid transfer operators
provides a factor of 0.11 when a 2 × 2 window is used, a factor of 0.06 for a 4 × 4
window, and a factor of 0.04, which matches perfectly the real convergence, when the
8 × 8 window is considered.
We remark that with the current multigrid approach, the quality of the coarse-grid
discretization may not be satisfactory, for example, for the chessboard or L-shaped in-
clusion example. For such cases, we recommend either using more powerful smoothers
such as the ILU smoother or adapting the coarsening. Furthermore, for PDEs with
strong local variations in the coefficient fields, homogenization techniques [22, 29, 30]
may also be used to obtain the coarser representation of the fine-grid problem. Com-
paring the results of the two combinations of intergrid transfer operators, we observe
a very similar performance for all five test cases studied here. In the rest of the
paper, we therefore choose the strategy (CP,CR) because of its simplicity and low
computational cost.

5. LFA results for PDEs with random coefficients. Here, we consider the
stochastic PDE (5) defined on a unit square domain D = (0, 1)2 with homogeneous
Dirichlet boundary conditions. Two different types of diffusion coefficients based on
random jumps and lognormal random fields are studied. The randomly jumping
coefficient problem can be seen as a transition from the deterministic to a stochastic
setting.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1398 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Fig. 4. An example of random realization of U = ln k with m = 5 on a unit square domain.

5.1. Randomly jumping coefficients. To simulate random jumps, the domain


D is subdivided into square-blocks of size [ 81 × 18 ] and the value of the coefficients on
each of the blocks is sampled as
(22) k = eU with U ∼ U{−m, m} and m ∈ Z.
In other words, U is an independent identically distributed (i.i.d.) integer sampled
from a discrete uniform distribution U{−m, m}. Here, the integer m defines the order
of magnitude of the jumps; an example for m = 5 is shown in Figure 4. Notice that
for this choice of m, we may encounter interfaces with maximum jumps of magnitude
equal to e10 .
For each random realization of the jumping coefficient field, we compare the LFA
two-grid convergence factors with the computed asymptotic convergence factors of the
multigrid method by using W-cycles. To perform the LFA, we again use a window of
size 8 × 8 capturing the heterogeneity of the whole domain. Furthermore, the LFA
equivalent of the randomly jumping coefficient problem is similar to the four corner
problem in Figure 3(d) with the magnitude of each block given by (22) and the cross-
point exactly at the center of the LFA block. Regarding the multigrid components,
a lexicographic Gauss–Seidel iteration is employed as the smoother, and the simplest
combination (CP,CR) of intergrid transfer operators is chosen. Also, we use a 4×4 grid
as the coarsest in the multigrid hierarchy. To determine the asymptotic convergence
factors of the multigrid method, the right-hand side is again set to zero.
The experimental convergence factor of the multigrid method for the ith realiza-
tion of the random field is then computed, as follows:
 1/ki
||reski ||∞
(23) ρi = for i = 1, 2, . . . , NM G ,
||res0 ||∞
where ||res0 ||∞ is the infinity norm of the residual obtained from an initial solution
and ||reski ||∞ is the residual after ki iterations of the multigrid cycle. We use these
quantities to calculate the average and the standard deviation of the asymptotic con-
vergence factors:
v
NM G u N
1 X u 1 XMG

σM G = t
2
(24) hρiM G = ρi (ρi − hρiM G ) ,
NM G i=1 (NM G − 1) i=1

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1399
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Fig. 5. Comparison of the mean and the standard deviation of the LFA (dashed line) and
multigrid (solid line) convergence factors for different W -cycling strategies for randomly jumping
coefficients with m = 2 (left) and m = 5 (right).

respectively. These averaged quantities are defined similarly for the LFA results (based
on LFA two-grid factors) and are denoted, respectively, by hρiLF A and σLF A .
In Figure 5, we show the comparison, the mean ± standard deviation, of the
LFA prediction and the multigrid convergence for jump parameter m = 2 (left) and
m = 5 (right) computed using NLF A = NM G = 100. For clarity, the LFA predicted
mean and standard deviation are plotted at an arbitrary location (h = 1/200) and
the mean is extended throughout the x-axis. Overall, a good match between the LFA
and multigrid convergence is seen up to one decimal place. We also observe that for
this specific jumping coefficient problem, there is no further improvement with an
increase in the number of smoothing steps after the W (2, 2)−cycle.
5.2. Lognormal random fields. Next, we test the LFA prediction capability
for a more realistic lognormal diffusion problem. Lognormality leads to positive per-
meability throughout the domain. The logarithm of the permeability field, Z = log k,
is modeled by a zero-mean Gaussian random field Z : Ω × D → R. A simplification is
the use of a homogeneous covariance function CΦ : R → R, so that

(25) Cov(Z(x, ·), Z(y, ·)) = CΦ (r) with r = ||x − y||2 .

By the so-called Matérn family of covariance functions, random coefficient fields with
different degrees of smoothness can be generated. The Matérn covariance function
[32] is characterized by a parameter set Φ = (νc , λc , σc2 ), as follows:
1−νc
 νc  
22 √ r √ r
(26) CΦ (r) = σc 2 νc K ν 2 νc .
Γ(νc ) λc λc

Here, Γ is the gamma function and Kν the modified Bessel function of the second
kind. The different parameters have different roles regarding the field’s randomness.
Parameter νc defines the field’s smoothness, σc2 represents its variance, and λc is
the correlation length of the covariance function. Moreover, parameters λc and σc2
prescribe the number of peaks and the amplitude of the random field, respectively.
When the smoothness parameter is νc = 1/2, the Matérn function corresponds to an
exponential model, whereas when νc → ∞ it represents a Gaussian model.
Realizations of the random field are “almost surely” Hölder continuous, Z, k ∈
C η (D), with 0 < η < νc (see, e.g., [33, 34]). For (5) regularity results have been
obtained by taking into account the regularity of the lognormal coefficient field; see
[35, 36, 34].

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1400 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

Remark. The random fields are generated using the circulant embedding method
[37, 38]. In Appendix A, we briefly describe this sampling algorithm. This method
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

employs the fast Fourier transform (FFT) for the covariance matrix decomposition
and, therefore, requires O(h−2 log h−2 ) operations to generate one sample of the ran-
dom field on a uniform grid mesh size h. Other techniques, e.g., the Karhunen–Loéve
(KL) expansion, can also be utilized to sample these fields. The benefit of the circu-
lant embedding technique is that it yields an exact representation of the random field
on the sampling mesh for all Φ, whereas the KL expansion gives a low-dimensional
representation of the field introducing a bias (due to truncation after a finite number
of eigenmodes). The number of terms in the KL expansion is dependent on Φ [39, 40]
and can be large for stochastic processes with small correlation lengths.
To illustrate the performance of LFA for PDEs with random parameters, we
consider four Matérn reference parameter sets Φ with increasing order of complexity,
listed in Table 2. Random fields generated with these Matérn parameter sets, sampled
on a uniform mesh, are presented in Figure 6.

Table 2
Different combinations of the Matérn reference parameters Φ = (νc , λc , σc2 ) with increasing
complexity from left to right.

Φ1 Φ2 Φ3 Φ4
(1.5,0.3,1) (0.5,0.3,1) (1.5,0.1,3) (0.5,0.1,3)

1 1 2
2
0.8 0.8 1
1
0.6 0.6 0
y
y

0
0.4 0.4 -1

-1
0.2 0.2 -2

-2 -3
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x
(a) Φ1 = (1.5, 0.3, 1) (b) Φ2 = (0.5, 0.3, 1)
1 1
6
6

0.8 0.8 4
4

0.6 2 0.6 2
y
y

0 0
0.4 0.4

-2 -2
0.2 0.2
-4 -4
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x
(c) Φ3 = (1.5, 0.1, 3) (d) Φ4 = (0.5, 0.1, 3)
Fig. 6. Logarithm of the permeability field, log10 k, generated using four reference parameter
sets Φ = (νc , λc , σc2 ).

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1401
Table 3
Comparison of the average and the standard deviation of the LFA and multigrid convergence
factors. Different numbers of smoothing steps, h = 1/64, and the four reference parameter sets in
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Table 2 are considered. The coarsest grid considered in the multigrid method is a 4x4 mesh.

Φ1 Φ2 Φ3 Φ4
MG LFA MG LFA MG LFA MG LFA
hρi 0.20 0.20 0.20 0.20 0.19 0.20 0.23 0.21
W (1, 1)
σ 0.004 0.002 0.004 0.005 0.005 0.01 0.04 0.02
hρi 0.04 0.04 0.04 0.04 0.07 0.06 0.17 0.08
W (2, 2)
σ 0.001 0.001 0.002 0.003 0.03 0.03 0.04 0.03
hρi 0.01 0.02 0.02 0.02 0.05 0.03 0.13 0.06
W (3, 3)
σ 0.001 0.002 0.002 0.004 0.02 0.02 0.03 0.03

For a fixed Φ and mesh size h, we generate NM G realizations of the permeability


field. For each of these random fields, we can compare the LFA two-grid convergence
factors with the computed asymptotic convergence factors of the multigrid method by
using the same components that were utilized for randomly jumping coefficient fields.
The comparison is shown in Table 3, where different numbers of smoothing steps are
considered and the comparison is done for the four reference parameter sets, described
in Table 2. For all experiments, we set h = 1/64 and NM G = NLF A = 100. In general,
we observe a good agreement between the experimental and the LFA quantities. For
the first three sets of parameters excellent convergence factors are obtained already
with two pre- and two postsmoothing steps. When the more difficult set of parameters
is considered, however, more smoothing steps may be necessary to obtain a good
convergence factor. It is also pointed out that the performance of W(1,1)-cycle is well
predicted by LFA for all Φi , i = 1, 4. A slight discrepancy is observed for the W(2,2)-
and W(3,3)-cycles in the case of Φ4 .

5.3. Mesh dependency. Next, we wish to study the influence of the size of
the sampling mesh on the multigrid convergence. For this purpose, we choose two
representative parameter sets describing a smooth and a highly oscillating random
field that are generated by the parameter sets Φ2 and Φ4 , respectively. Figure 7 shows
the average convergence factors for Φ2 (left side) and Φ4 (right side), predicted by LFA
(top) and experimentally observed multigrid convergence (bottom), for different mesh
sizes and different numbers of smoothing steps. For Φ2 , the average reduction factor is
roughly the same, independent of the size of the mesh. These predictions coincide well
with the experimentally observed factors. For the results corresponding to parameter
set Φ4 , we observe robustness of the method when the mesh is sufficiently fine, which
is also confirmed by the multigrid experiments. In the same figure, the standard
deviation is presented, which decreases when h → 0. The LFA predictions on the
coarser grids are less reliable when compared to the multigrid convergence rates. We
would like to mention that a three-grid LFA [41] is not helpful here. On coarse grids,
boundary conditions have an impact on the method’s convergence, but they are not
taken into account in the analysis. Moreover, to analyze the robustness of the proposed
multigrid method, in the next experiment we fix the field’s smoothness parameter as
νc = 0.5 and vary the other parameters σc2 ∈ [0.5, 5] and λc ∈ [0.05, 0.5]. In Figure 8
we show the average LFA two-grid convergence factors when two smoothing steps are
considered for h = 1/32 (left) and h = 1/64 (right). The multigrid convergence is
very satisfactory for all combinations of the parameters and increases slightly when
λc tends to be small and σc2 becomes large. This case, however, represents a rather
extreme situation in which the jumps in the permeability field are of more than

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1402 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

Φ2 Φ4
0.4
W (1, 1) W (1, 1)
0.4
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

W (2, 2) W (2, 2)
0.3 W (3, 3) W (3, 3)
hρiLF A ± σLF A

hρiLF A ± σLF A
0.3

0.2
0.2

0.1 0.1

0 0
0 50 100 150 200 250 0 50 100 150 200 250
1/h 1/h
0.4
W (1, 1) W (1, 1)
0.4
W (2, 2) W (2, 2)
0.3 W (3, 3) W (3, 3)
0.3
hρiM G ± σM G

hρiM G ± σM G

0.2
0.2

0.1 0.1

0 0
0 50 100 150 200 250 0 50 100 150 200 250
1/h 1/h

Fig. 7. Average reduction LFA two-grid factors (top) and asymptotic multigrid convergence
(bottom) along with the standard deviation for two reference parameter sets Φ2 (left column) and
Φ4 (right column); NM G = NLF A = 100.

Fig. 8. Contour of average LFA two-grid convergence factors, hρiLF A , with two smoothing
steps for different λc and σc2 , and for fixed νc = 0.5 with h = 1/32 (left) and h = 1/64 (right).

15 orders of magnitude. Again, we see an improvement in the convergence rate with


grid refinement. Note that with the considered range of parameters we cover all
realistic cases.
6. Multilevel Monte Carlo computations. In the preceding section, we
demonstrated that LFA performs well in predicting the mean convergence rate of the
multigrid method for a given Matérn parameter set. In this section, we wish to utilize

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1403

LFA-based information to select the multigrid cycle for MLMC simulations. Within
MLMC methods, a hierarchy of grids, {Dh` }L `=0 , characterized by h0 > · · · > hL is
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

defined to accurately estimate the quantity of interest Q, based on the linearity of the
expectation operator, i.e.,
L
X
(27) E[QhL ] = E[Qh0 ] + E[Qh` − Qh`−1 ].
`=1
The advantage of using the above telescopic decomposition is that on the coarsest grid,
for ` = 0, large numbers of samples to accurately determine E[Qh0 ] can be computed
at a small computational effort. For larger values of ` the numerical solution is
comparatively expensive; however, fewer samples are required on these finer levels, as
the variance of the correction terms, V[Qh` − Qh`−1 ], is significantly smaller compared
to the variance V[Qh` ]. A multilevel estimator for E[QhL ] can be defined in terms of
the standard MC estimator as
N0 L N
!
ML 1 X X 1 X̀
(28) EL [QhL ] := Qh (ωi ) + (Qh` (ωi ) − Qh`−1 (ωi )) ,
N0 i=1 0 N` i=1
`=1
where the level dependent samples N` ∈ N form a decreasing sequence for increasing
`-values. We will use the same geometric sequence of meshes in the MLMC estimator
as employed for the multigrid method described in subsection 2.2, based on uniform
coarsening such that h` = h`−1 /2 = 2−` h0 , with h0 the coarsest mesh size. As each
of the estimators in (28) is independent, the variance of the multilevel estimator is
the sum of the variances of individual estimators, i.e.,
L
X V`
(29) V[ELM L [QhL ]] = ,
N`
`=0

where the level dependent sample variance, V` = V[Qh` − Qh`−1 ] = O(2−β` ), β > 0,
and Qh−1 = 0. Assume that the mean finite volume error for mesh width h` satisfies
(30) |E[Qh` − Q]| = O(hα
`)

with α > 0. The mean square error of the MLMC estimator can be expressed as a
sum of bias and sampling errors,
E[(ELM L [QhL ] − E[Q])2 ] = (E[QhL − Q])2 + V[ELM L [QhL ]],
L
X V`
(31) = c0 h2α
L + ,
N`
`=0
with c0 a constant independent of hL and α. The level dependent sample size N`
can be chosen such that the sampling errors of each of the MC estimators in (28)
reduces to the size of the discretization error. This is achieved by taking N` such
−2αL
that V` /N` = O(h2α
L ) = O(2 ). Moreover, by fixing the number of the finest level
samples NL , the number of coarse-grid samples can be simply expressed as
(32) N` = dNL 2β(L−`) e.
In practice the value NL is small ∼ O(1) and can be chosen heuristically [42, 43]. The
sampling error on the coarsest level does not depend on β, and therefore, by using (32),
the balance V0 /N0 = O(h2αL ) holds only asymptotically for large L. Alternatively, one
can also solve an optimization problem, as in [8, 9, 10], which minimizes the total cost
of the MLMC estimator for a given tolerance ε.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1404 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

The computational cost of the MLMC estimator is given by


L
X
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

(33) WLM L = N` · (W` + W`−1 ),


`=0

where W` represents total number of arithmetic operations to obtain a sample of the


quantity of interest Q on grid level `. This typically includes the cost of generating the
random permeability field, assembling and solving the linear system of equations, and
postprocessing. Of these, the cost of solving the linear system is usually dominant,
and, therefore, W` will only represent the cost of the multigrid method. For the
current combination of multigrid components, it is straightforward to show that the
total amount of computational work for one multigrid cycle is proportional to the
number of grid points on the finest multigrid level [28]. The cost of the multigrid
solver can thus be expressed as
(34) W` = c̄νk` h−2
` ,

where c̄ is a constant, k` ∈ N is the number of multigrid iterations, ν is the number of


smoothing steps, and h−2 ` represents the number of unknowns on grid level `. The cost
associated with the other components, such as computations of the defect, transfer
to coarser grids, or interpolation of corrections, is accounted for by the constant c̄, to
facilitate the comparison of W-cycles with different numbers of smoothing steps.
The product N` · (W` + W`−1 ) in (33) gives the individual cost contributions at
any level `. For the proposed multigrid solver, we have (W` + W`−1 ) = O(h−2 ` ). For
simplicity, let’s assume that the cost of one sample on level ` is given by O(h−γ ` ),
γ ≥ d, where d is the number of spatial dimensions. The computational cost (33)
PL
can be conveniently expressed as WLM L = O( `=0 2(γ−β)` ), leading to the following
three cases. When the rate of decay of N` is faster than the growth in computational
cost (β > γ), the dominating cost comes from the coarsest level. When the sample
decay rate is similar to the growth in computational cost (β = γ), all levels equally
contribute to the total cost. If the sample decay rate is slower than the growth in
computational cost (β < γ), the dominant cost is on the finest level. This scenario is
typically encountered in many practical settings involving three-dimensional problems
with low regularity permeability models, resulting in a slow variance decay (or a
small β).
Note that in many cases the rates α, β, γ are not available. A common practice is
to use a few “warm-up samples and levels” to obtain an estimate for these quantities.
Having a reliable LFA tool can be further helpful in efficiently optimizing the MLMC
simulation, especially to assess the performance of a multigrid solver for different
combinations of transfer operators and smoothers. This way, one can select the most
suitable combination for the considered random input data. We can also use the
LFA to obtain solver statistics on fine grids where computing warm-up samples is not
feasible due to high computational cost.
6.1. MLMC numerical experiments. In this section we analyze the perfor-
mance of the proposed multigrid MLMC method. We consider PDE (5) on domain
D ∈ (0, 1)2 with mixed Dirichlet–Neumann boundary conditions,
(35) u(0, y, ω) = 1, u(1, y, ω) = 0, and
∂u ∂u
k(x, ω) = 0, k(x, ω) = 0,
∂x y=0 ∂x y=1

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1405
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Fig. 9. (Left) Convergence of finite volume error. (Right) Decay of level dependent variance
for the outflow, Qh` , with mesh refinement for different Matérn parameters.

respectively. For all tests the source term is set to zero, i.e., f = 0. As the quantity
of interest, the outflow through the boundary x = 1, y = (0, 1), also referred to as
effective permeability, is considered:
Z 1
∂u
(36) Q(u) = − k(x, ω) (x, ω) dy.
0 ∂x x=1

In Figure 9, we show the convergence of the FV bias, |E[Qh` − Qh`−1 ]|, and the level
dependent variance V` for the considered quantity of interest (36). We observe that
the rate of decay of the finite volume bias depends on the smoothness parameter νc
of the random fields. In the case of parameter sets Φ1 and Φ3 , we see a second-
order convergence, whereas a first-order convergence is observed for Φ2 and Φ4 . The
correlation length λc and variance σc2 only affect the proportionality constant. Also,
for all four cases, the level dependent variance shows a quadratic decay.
To compute samples of Qh` (ωi ) − Qh`−1 (ωi ), we invoke the multigrid solver twice
on level ` and ` − 1 with the permeability field k(x` , ωi ) and an upscaled version
of permeability k(x`−1 , ωi ), respectively. We define k(x`−1 , ωi ) using the covariance
upscaling technique from [44] (see Appendix B). It is important to emphasize that
homogenization techniques, e.g., [30, 31] for obtaining the coarse-scale solution, can-
not be used to compute Qh`−1 (ωi ). This is due to the fact that the homogenized
permeabilities are typically obtained by averaging procedures that can modify the
spatial covariance as well as the marginal distribution defined for the permeability
field. This will lead to a violation of the telescopic identity (27). However, homog-
enization methods can still be used within a multigrid solver to obtain coarse-grid
discretization operators.
Next, we conduct MLMC experiments for Matérn parameter sets Φ2 and Φ4 ,
using W(1,1), W(2,2), and W(3,3)-cycles. The multigrid components described in
section 2.2 are used with stopping criterion εM G = 10−10 . We consider α = 1 and
β = 2 for both parameter sets Φ2 and Φ4 . Recall that the MLMC sampling strategy
(32) requires the total number of MLMC levels, the number of finest level samples
NL , along with the rate β. We set NL = 64 within MLMC for both parameter sets
resulting in a sample sequence N` = NL 4(L−`) . We also tested smaller values for
NL , where we still observed high statistical fluctuation in the estimates of E[QhL ],
especially for parameter set Φ4 . As mentioned earlier, the number of samples on all
levels can also be determined by solving an optimization problem where the total cost
of MLMC for a given tolerance is minimized [8, 9].

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1406 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

Rel. error in mean of Q Computational


Rel. work of Q
error in variance
10−1 106−1
10
slope=-1 W (1, 1) slope=-2
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

105 W (2, 2)
W (3, 3)

L [Q ]
hL ] − Eref [Qh∗ ]

h∗
slope=2

(sec)
104−2
10
ML

hL ] − Vref
MCPU-time M
10−2 103
M L [Q

L [Q
102−3
10

VL
EL

101

10−3 100−4
10
1 1.5 2 2.5 3 1 1 1.5 1.5 2 2 2.5 2.5 3 3
log 10(h−1
L ) loglog
10(h −1−1
10(h
L L) )

M L [Q
Fig. 10. (Left) Convergence of error in EL hL ] with increasing number of MLMC levels for
parameter Φ2 . The reference solution is based on mesh size h∗ = 1/512. (Right) Mean CPU-times
versus accuracy for different W-cycles. Computational cost proportional to O(h−2L ) is observed for
all cycling strategies.

Remark. The experiments are performed on a parallel machine with a homoge-


neous set of cores. The MLMC algorithm allows for three degrees of parallelization:
over samples, over MC levels, and within the multigrid solver. Recent advancements
on parallelization can be found in [43, 45, 46]. We only consider parallelization over
the MLMC samples and levels. For the considered MLMC sampling strategy with
known number of samples and levels, a greedy algorithm based approach can be fol-
lowed where cores are distributed between levels such that the sum of idle times for
all cores is minimized.

6.1.1. MLMC results for Φ2 . We first analyze the error convergence of the
expected value and the variance of the quantity of interest. In Figure 10 (left), relative
errors with respect to a reference solution computed using the MLMC estimator with
the finest level h∗ = 1/512 are shown. We see the error |ELM L [QhL ] − Eref ML
[Qh∗ ]|
converges with O(hL ). Next, we analyze the performance of the MLMC estimator
for different multigrid cycling strategies. In Figure 10 (right), we present the CPU-
times for the MLMC simulations with an increasing number of MLMC levels. These
CPU-times are derived by summing up the run-times from the multigrid solves for all
samples over all levels. Here, we do not include the cost of generation of the random
field and the postprocessing costs as they are the same for different cycles. The cost
scales as O(h−2L ) coinciding with the theoretical MLMC complexity when β = γ. No
substantial difference in the CPU-times from the three cycles is observed. We also
show the levelwise CPU-times in Table 4 for the three cycles with the same number
of MLMC samples per level. In general, the computational cost improves with levels
for all cycles. We observe that the W(2,2)-cycle slightly outperforms the other two
variants.
We also show the distribution of the number of multigrid iterations for the W(2,2)-
cycle on different grid levels along with the number of iterations predicted by the
LFA, k̃` := dlog(εM G )/ log(hρiLF A,` )e, in Figure 11. The cost is more heterogeneous
on coarser levels with larger variance in the number of iterations and is homogeneous
for h` = 1/32 and onward. A high variance in the multigrid convergence rate was also
predicted by the LFA experiments in the previous section. Notice that if the dominant
cost of the MLMC simulation comes from the coarsest level, the solution time for the

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1407
Table 4
Comparison of the three W-cycles in terms of the levelwise CPU-times for parameter Φ2 .
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Levelwise CPU-time (sec.)


h` N` W(1,1) W(2,2) W(3,3)
1/8 262144 489 331 323
1/16 65536 528 391 390
1/32 16384 454 362 363
1/64 4096 407 336 348
1/128 1024 376 320 342
1/256 256 357 308 338
1/512 64 350 300 333

e
k3 = 7
104 h1 = 1/16 h2 = 1/32 h3 = 1/64
103

103 102
e
k1 = 9 102
count

count

count
102
e
k2 = 8 101
101
101

100 100 100


6 7 8 9 10 6 7 8 6 7 8
Iteration Iteration Iteration

Fig. 11. Distribution of number of multigrid iterations of W (2, 2)-cycles for different mesh
sizes for parameter Φ2 to reach εM G < 10−10 and the LFA predicted number of iteration e k` .

MLMC technique will be varying significantly when the proposed multigrid solver is
utilized on the coarse grids. In such cases, it is advised to use a sparse direct solver
on these coarser levels.
In this work, the considered multigrid stopping criteria of 10−10 is quite conser-
vative. For many engineering applications a residual reduction of 10−6 may already
be sufficient to reach a converged solution, therefore reducing the computational cost
roughly by a factor of two.

6.1.2. MLMC results for Φ4 . Now we describe MLMC results for the chal-
lenging parameter set Φ4 for the same quantity of interest. In Figure 12 (left), we
display the convergence of the relative error in the expected value of the quantity of
interest using a reference solution obtained with finest resolution h∗ = 1/1024. Due
to a large variance σc and a small correlation length λc of the random field, the error
is larger, as compared to the results of parameter set Φ2 for same mesh size hL , and
the expected convergence rates are visible on relatively fine grids. A similar trend
is observed for the computational cost with scaling as O(h−2 L ) in Figure 12 (right).
Again, in Table 5, the levelwise CPU-times for the three cycles are listed. Similar to
the results for set Φ2 , the cost improves with levels but the gain from W(2,2)-cycles
is more prominent.
Last, the distribution of the number of iterations for W(2,2)-cycles is depicted
in Figure 13. High variability in the number of multigrid iterations persists until a
relatively fine grid (h = 1/128). However, the average number of iterations predicted
by LFA coincides very well with the mode of these distributions. Again, a direct
solver can be applied on these coarser levels to get a more reliable estimate of the
cost. Here, we point out that the W(2,2)-cycle takes roughly seven iterations similar
to parameter Φ2 to reduce the residual by 10−10 , indicating robustness with respect
to the Matérn parameter.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1408 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

Rel. error in mean of Q Rel.


Computational
error in variance
work of Q
100 6 1
1010
slope=-1 W (1, 1) slope=-2
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

5 W (2, 2)
10 0
W (3, 3)

L [Q ]
hL ] − Eref [Qh∗ ]

10

h∗
slope=2

(sec)
−1
10
ML

hL ] − Vref
104

MCPU-time M
10−1
103
M L [Q

L [Q
10−2

VL
EL

10−2
102

10−3 101−3
10
1 1.5 2 2.5 3 1 1 1.5 1.5 2 2 2.5 2.5 3 3
log 10(h−1
L ) loglog
10(h−1−1
10(h
L L) )

M L [Q
Fig. 12. (Left) Convergence of error in EL hL ] with increasing number of MLMC levels for
parameter Φ4 . The reference solution is based on mesh size h∗ = 1/1024. (Right) Mean CPU-times
versus accuracy for different W-cycles. Computational cost proportional to O(h−2L ) is observed for
all cycling strategies.

Table 5
Comparison of the three W-cycles in terms of the levelwise CPU-times for parameter Φ4 .

Levelwise CPU-time (sec.)


h` N` W(1,1) W(2,2) W(3,3)
1/8 1048576 2.23[+3] 2.10[+3] 2.18[+2]
1/16 262144 2.42[+3] 2.52[+3] 2.80[+3]
1/32 65536 1.85[+3] 2.10[+3] 2.45[+3]
1/64 16384 1.56[+3] 1.59[+3] 1.98[+3]
1/128 4096 1.45[+3] 1.33[+3] 1.60[+3]
1/256 1024 1.40[+3] 1.23[+3] 1.40[+3]
1/512 256 1.38[+3] 1.23[+3] 1.34[+3]
1/1024 64 1.36[+3] 1.21[+3] 1.31[+3]

104 104 103


h1 = 1/16 h2 = 1/32 e h3 = 1/64
k3 = 9
e
k1 = 12 e
k2 = 10
103 103
102
count

count

count

102 102

101
101 101

100 100 100


10 15 20 25 30 35 8 10 12 14 16 18 20 22 8 10 12 14
Iteration Iteration Iteration
3 3
10 10
e h4 = 1/128 h5 = 1/256 h6 = 1/512
k4 = 8
e
k5 = 7 e
k6 = 7
2
10
102 102
count

count

count

101
101 101

100 100 100


7 8 9 10 11 6 7 8 6 7 8
Iteration Iteration Iteration

Fig. 13. Distribution of the number of multigrid iterations of W (2, 2)-cycles for different mesh
sizes for parameter Φ4 to reach εM G < 10−10 and the LFA predicted number of iteration e k` .

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1409

7. Conclusions. A novel, generalized local Fourier analysis which can be em-


ployed for quantitative assessment of multigrid methods for PDEs involving jumping
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

and random coefficients has been presented. In particular, a cell-centered multigrid


algorithm for solving a model problem based on a single-phase flow problem in a ran-
dom porous medium has been used to show the accuracy of the proposed analysis.
This technique, however, is appropriate for the prediction of the performance of a
wider class of multigrid methods for solving PDEs with random fields. For instance,
this approach can be extended to higher-order cell-centered and vertex-centered dis-
cretizations and corresponding multigrid methods, or even to multigrid methods for
structured grid discretizations of coupled PDE systems, such as the Darcy–Stokes
problem; see [47, 48]. Moreover, it should also be possible to apply it to triangular
grids and finite element discretizations [49, 50]. Further, the novel local Fourier anal-
ysis could be used to estimate a priori the time needed for solving certain uncertainty
quantification problems by using a multigrid multilevel Monte Carlo method.
Appendix A. Sampling Gaussian random fields. A fast algorithm for gen-
erating Gaussian random fields is critical for obtaining an efficient multigrid MLMC
estimator. When dealing with stationary covariance functions, one can utilize spec-
tral generators, e.g., in [37, 38, 51], that exploit the efficiency of the FFT algorithm
to achieve fast sampling of these random fields on a uniform mesh. In this work,
we use the fast Fourier transform moving average (FFT-MA) technique from [51] to
decompose the covariance matrix CΦ (r) (recall (26)). Although this sampling method
is similar to the Cholesky factorization technique, the key idea is to make the com-
putational domain periodic. Thus, the resulting covariance operator is also periodic,
which can now be decomposed as a convolutional product. The samples of random
fields are computed using a cheaper vector-vector product compared to the expensive
matrix-vector operation required when using Cholesky factorization. As a periodic
covariance function sampled on a uniform grid results in a circulant covariance matrix,
these methods are sometimes also referred to as the circulant embedding technique.
In the following, we provide a brief description of FFT-MA method from [51].
Samples of correlated Gaussian random vectors z` (ω) can be obtained using a
Cholesky decomposition of the covariance matrix C`Φ on mesh D` as

(37) C`Φ = L` LT` and use z` = L` y` ,

where y` is a vector of i.i.d. samples from the standard normal distribution. The
FFT-MA method is based on a decomposition of the covariance function CΦ (r) as
0 0
a convolutional product of some function SΦ (r) and its transpose SΦ (r) (SΦ (r) =
SΦ (−r)). In a discrete setting, we can express this decomposition as

(38) c` = s` ∗ s0` ,

where c` , s` are vectors obtained by evaluating CΦ (r) and SΦ (r), respectively, at grid
points of the mesh D` . A correlated random vector z` can now be synthesized by
using the convolution product

(39) z` = s` ∗ y` .

The FFT-MA method performs the above computations in the frequency domain.
First the vector c` is transformed into a periodic signal, which is also real, positive,
and symmetric. For more details on the practical aspects of this transformation, see
[37, 38, 52, 44]. Note that the resulting vector s` is also real, positive, and symmetric

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1410 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

and s` = s0` . As the convolution product in the spatial domain is equivalent to the
componentwise product in the frequency domain, we can rewrite (38) as
p
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

(40) F(c` ) = F(s` ) · F(s` ) =⇒ F(s` ) = F(c` ),


where F denotes the discrete FFT and · denotes componentwise multiplication. It is
pointed out that the componentwise square-root operation does not pose any problems
as the power spectrum F(c` ) is real, positive. Next, we express the convolution
product in (39) as a vector-vector product in the frequency domain as
(41) F(z` ) = F(s` ∗ y` ) = F(s` ) · F(y` ).
In the final step, an inverse FFT is applied to obtain the samples for Gaussian random
fields
(42) z` = F −1 (F(s` ) · F(y` )).
Due to the periodicity in the covariance vector c` , the resulting random field z` is also
periodic. Thus, we only retain the part of the vector that corresponds to the physical
domain. We also remark that it takes two FFT evaluations to obtain one sample of z`
(ignoring the FFT operation in (40) that is performed just once). Therefore, in terms
of the number of floating point operations, the sampling cost is significantly smaller
compared to one multigrid solve for the mesh sizes considered.
Appendix B. Upscaling Gaussian random fields. While estimating the cor-
rection term E[Qh` − Qh`−1 ] in the telescopic sum (27), the approximations Qh` (ωi )
and Qh`−1 (ωi ) need to be positively correlated such that the variance V[Qh` − Qh`−1 ]
is small. This is typically achieved by first sampling the fine-grid permeability vec-
tor, k(x` , ωi ), to compute Qh` (ωi ) and using an upscaled version, k(x`−1 , ωi ), for
Qh`−1 (ωi ). While performing such upscaling of random fields, it is important to en-
sure that the telescopic sum (27) is not violated. In other words, the expectation of
the random variable Qh` when estimating E[Qh` − Qh`−1 ] and E[Qh`+1 − Qh` ] should
be the same, i.e.,
(43) E[Qh` ](coarse) = E[Qh` ](f ine) for ` = {0, 1, . . . , L − 1}.
There are a number of ways to generate an upscaled version of a random field. For in-
stance, the coarse-grid permeability can be computed by means of homogenization and
variational coarsening with operator-dependent interpolation (regarding more general
flow equations with a full tensor permeability and additional closure terms) [30, 31].
Upscaling algorithms based on homogenization techniques work well in the context of
deterministic PDEs; however, these homogenization procedures may result in a mod-
ified covariance structure on the coarser levels for PDEs with random coefficients,
violating (43). They may, however, be applicable for quantities of interest that do not
depend on the permeability field. In general, the same covariance structure can be
maintained using the covariance upscaling [44] that employs the spectral generator
on two consecutive grids using the same normally distributed vector y` . Furthermore,
in the case of the FFT-MA algorithm, the vector y` is associated with respective grid
points, and coarser realizations of the fine-grid Gaussian random field z` (log k(x` ))
can be obtained by using multidimensional averaging of the vector y` . For instance,
in two dimensions for the cell-centred mesh,
i,j 1  2i−1,2j−1 
(44) y`−1 = y` + y`2i−1,2j + y`2i,2j−1 + y`2i,2j ,
2

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1411
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

Fig. 14. Comparison of cross section of the pressure fields u at x = 0.5 (left) and y = 0.5
(right) obtained from a random permeability field (based on Φ4 ) on a 256 × 256 grid and from the
corresponding upscaled permeabilities at 128 × 128 and 64 × 64 grids.

where (i, j) is the cell index for the mesh D`−1 . The scaling by a factor of 2 is needed
i,j
to obtain a standard normal distribution for the averaged quantity y`−1 . The coarse
random field can now be simply assembled as

(45) z`−1 = F −1 (F(s`−1 ) · F(y`−1 )).

Although this process can be recursively applied to generate upscaled random fields
on all coarser scales, for the MLMC estimator, we just need this upscaling procedure
for the next coarser grid. As the averaging in (44) smooths out high frequencies, the
upscaled version z`−1 will also be smoother compared to z` . In Figure 14, we compare
the cross sections of pressure fields obtained from a permeability field sampled (using
parameter set Φ4 ) on a 256 × 256 grid and from the corresponding upscaled versions
on 128 × 128 and 64 × 64 grids. The fine-scale properties are well preserved on the
coarser levels.

REFERENCES

[1] M. Bolten and H. Rittich, Fourier analysis of periodic stencils in multigrid methods, SIAM
J. Sci. Comput., 40 (2018), pp. A1642–A1668, doi:10.1137/16M1073959.
[2] J. P. Delhomme, Spatial variability and uncertainty in groundwater flow parameters:
A geostatistical approach, Water Resources Research, 15 (1979), pp. 269–280, doi:10.1029/
WR015i002p00269.
[3] R. A. Freeze, A stochastic-conceptual analysis of one-dimensional groundwater flow in
nonuniform homogeneous media, Water Resources Research, 11 (1975), pp. 725–741,
doi:10.1029/WR011i005p00725.
[4] R. J. Hoeksema and P. K. Kitanidis, Analysis of the spatial structure of properties
of selected aquifers, Water Resources Research, 21 (1985), pp. 563–572, doi:10.1029/
WR021i004p00563.
[5] I. Babuška, F. Nobile, and R. Tempone, A stochastic collocation method for elliptic partial
differential equations with random input data, SIAM J. Numer. Anal., 45 (2007), pp.
1005–1034, doi:10.1137/050645142.
[6] D. Xiu, Numerical Methods for Stochastic Computations: A Spectral Method Approach, Prince-
ton University Press, Princeton, NJ, 2010.
[7] B. Seynaeve, E. Rosseel, B. Nicola, and S. Vandewalle, Fourier mode analysis of multigrid
methods for partial differential equations with random coefficients, J. Comput. Phys., 224
(2007), pp. 132–149, doi:10.1016/j.jcp.2006.12.011.
[8] K. Cliffe, M. B. Giles, R. Scheichl, and A. L. Teckentrup, Multilevel Monte Carlo
methods and applications to elliptic PDEs with random coefficients, Comput. Vis. Sci.,
14 (2011), pp. 3–15, doi:10.1007/s00791-011-0160-x.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


A1412 KUMAR, RODRIGO, GASPAR, AND OOSTERLEE

[9] M. B. Giles, Multilevel Monte Carlo path simulation, Oper. Res., 56 (2008), pp. 981–986,
doi:10.1287/opre.1070.0496.
[10] M. B. Giles, Multilevel Monte Carlo methods, Acta Numer., 24 (2015), pp. 259–328,
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

doi:10.1017/S096249291500001X.
[11] A. Brandt and O. E. Livne, Multigrid Techniques: 1984 Guide with Applications to Fluid
Dynamics, Classics in Appl. Math. 67, SIAM, Philadelphia, 2011.
[12] J. Molenaar, A simple cell-centered multigrid method for 3D interface problems, Comput.
Math. Appl., 31 (1996), pp. 25–33, doi:10.1016/0898-1221(96)00039-9.
[13] I. Yavneh, Coarse-grid correction for nonelliptic and singular perturbation problems, SIAM J.
Sci. Comput., 19 (1998), pp. 1682–1699, doi:10.1137/ S1064827596310998.
[14] W. Hackbusch, Multigrid Methods and Applications, Springer, New York, 1985.
[15] J. H. Bramble, R. E. Ewing, J. E. Pasciak, and J. Shen, The analysis of multigrid algorithms
for cell centered finite difference methods, Adv. Comput. Math., 5 (1996), pp. 15–29,
doi:10.1007/BF02124733.
[16] J. H. Bramble, J. E. Pasciak, and J. Xu, The analysis of multigrid algorithms with
non-nested spaces or non-inherited quadratic forms, Math. Comp., 56 (1991), pp. 1–34,
doi:10.2307/2008527.
[17] M. Brezina, R. Falgout, S. MacLachlan, T. Manteuffel, S. McCormick, and J. Ruge,
Adaptive smoothed aggregation (αSA) multigrid, SIAM Rev., 47 (2005), pp. 317–346,
doi:10.1137/S1064827502418598.
[18] K. Stüben, Appendix A: An introduction to algebraic multigrid, in Multigrid, U. Trotten-
berg, C. W. Oosterlee, and A. Schüller, eds., Academic Press, San Diego, CA, 2001, pp.
413–532.
[19] P. Vaněk, M. Brezina, and J. Mandel, Convergence of algebraic multigrid based on
smoothed aggregation, Numer. Math., 88 (2001), pp. 559–579, doi:10.1007/s211-001-8015-y.
[20] D. Braess, Towards algebraic multigrid for elliptic problems of second order, Computing, 55
(1995), pp. 379–393, doi:10.1007/BF02238488.
[21] P. Vaněk, J. Mandel, and M. Brezina, Algebraic multigrid by smoothed aggregation for
second and fourth order elliptic problems, Computing, 56 (1996), pp. 179–196, doi:10.1007/
BF02238511.
[22] R. E. Alcouffe, A. Brandt, J. E. Dendy, Jr., and J. W. Painter, The multigrid methods
for the diffusion equation with strongly discontinuous coefficients, SIAM J. Sci. Statist.
Comput., 2 (1981), pp. 430–454, doi:10.1137/0902035.
[23] J. Dendy, Black box multigrid, J. Comput. Phys., 48 (1982), pp. 366–386, doi:10.1016/0021-
9991(82)90057-2.
[24] J. Dendy, Black box multigrid for non-symmetric problems, Appl. Math. Comput., 13 (1983),
pp. 261–283, doi:10.1016/0096-3003(83)90016-4.
[25] M. Khalil and P. Wesseling, Vertex-centered and cell-centered multigrid for interface prob-
lems, J. Comput. Phys., 98 (1992), pp. 1–10, doi:10.1016/0021-9991(92)90168-X.
[26] P. Wesseling, Cell-centered multigrid for interface problems, J. Comput. Phys., 79 (1988),
pp. 85–91, doi:10.1016/0021-9991(88)90005-8.
[27] J. E. Dendy and J. D. Moulton, Black box multigrid with coarsening by a factor of three,
Numer. Linear Algebra Appl., 17, pp. 577–598, doi:10.1002/nla.705.
[28] U. Trottenberg, C. W. Oosterlee, and A. Schuller, Multigrid, Academic Press, San
Diego, CA, 2000.
[29] S. Knapek, Matrix-dependent multigrid homogenization for diffusion problems, SIAM J. Sci.
Comput., 20 (1998), pp. 515–533, doi:10.1137/S1064827596304848.
[30] J. Moulton, J. E. Dendy, and J. M. Hyman, The black box multigrid numerical homogeniza-
tion algorithm, J. Comput. Phys., 142 (1998), pp. 80–108, doi:10.1006/jcph.1998.5911.
[31] S. P. MacLachlan and J. D. Moulton, Multilevel upscaling through variational coarsening,
Water Resources Research, 42 (2006), doi:10.1029/2005WR003940.
[32] M. S. Handcock and J. R. Wallis, An approach to statistical spatial-temporal modeling of
meteorological fields (with discussion), J. Amer. Statist. Assoc., 89 (1994), pp. 368–390,
doi:10.2307/2290832.
[33] R. J. Adler, The Geometry of Random Fields, Classics in Math. Appl. 62, SIAM, Philadelphia,
2010, doi:10.1137/1.9780898718980.
[34] F. Nobile and F. Tesei, A multilevel Monte Carlo method with control variate for elliptic
PDEs with log-normal coefficients, Stoch. Partial Differ. Equ. Anal. Comput., 3 (2015),
pp. 398–444, doi:10.1007/s40072-015-0055-9.
[35] J. Charrier, R. Scheichl, and A. L. Teckentrup, Finite element error analysis of elliptic
PDEs with random coefficients and its application to multilevel Monte Carlo methods,
SIAM J. Numer. Anal., 51 (2013), pp. 322–352, doi:10.1137/110853054.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LFA FOR PDEs WITH JUMPING AND RANDOM COEFFICIENTS A1413

[36] J. Galvis and M. Sarkis, Approximating infinity-dimensional stochastic Darcy’s equa-


tions without uniform ellipticity, SIAM J. Numer. Anal., 47 (2009), pp. 3624–3651,
doi:10.1137/080717924.
Downloaded 01/21/24 to 46.232.120.201 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy

[37] C. R. Dietrich and G. N. Newsam, Fast and exact simulation of stationary Gaussian pro-
cesses through circulant embedding of the covariance matrix, SIAM J. Sci. Comput., 18
(1997), pp. 1088–1107, doi:10.1137/S1064827592240555.
[38] A. T. A. Wood and G. Chan, Simulation of stationary Gaussian processes in [0, 1]d , J.
Comput. Graph. Statist., 3 (1994), pp. 409–432, doi:10.2307/1390903.
[39] R. G. Ghanem and P. D. Spanos, Stochastic Finite Elements: A Spectral Approach, Springer,
New York, 1991, doi:10.1007/978-1-4612-3094-6.
[40] Ch. Schwab and R. A. Todor, Karhunen-Loeve approximation of random fields by
generalized fast multipole methods, J. Comput. Phys., 217 (2006), pp. 100–122,
doi:10.1016/j.jcp.2006.01.048.
[41] R. Wienands and C. W. Oosterlee, On three-grid Fourier analysis for multigrid, SIAM J.
Sci. Comput., 23 (2001), pp. 651–671, doi:10.1137/S106482750037367X.
[42] S. Mishra and Ch. Schwab, Sparse tensor multi-level Monte Carlo finite volume methods
for hyperbolic conservation laws with random initial data, Math. Comp., 81 (2012), pp.
1979–2018, doi:10.1090/S0025-5718-2012-02574-9.
[43] S. Mishra, Ch. Schwab, and J. Šukys, Multi-level Monte Carlo finite volume methods for
nonlinear systems of conservation laws in multi-dimensions, J. Comput. Phys., 231 (2012),
pp. 3365–3388, doi:10.1016/j.jcp.2012.01.011.
[44] S. Mishra, Ch. Schwab, and J. Šukys, Multi-level Monte Carlo finite volume methods for
uncertainty quantification of acoustic wave propagation in random heterogeneous layered
medium, J. Comput. Phys., 312 (2016), pp. 192–217, doi:10.1016/j.jcp.2016.02.014.
[45] D. Drzisga, B. Gmeiner, U. Rüde, R. Scheichl, and B. Wohlmuth, Scheduling massively
parallel multigrid for multilevel Monte Carlo methods, SIAM J. Sci. Comput., 39 (2017),
pp. S873–S897, doi:10.1137/16M1083591.
[46] J. Šukys, S. Mishra, and Ch. Schwab, Static load balancing for multilevel Monte Carlo finite
volume solvers, in Parallel Processing and Applied Mathematics, Springer, Berlin, 2012,
pp. 245–254, doi:10.1007/978-3-642-31464-3 25.
[47] P. Luo, C. Rodrigo, F. J. Gaspar, and C. W. Oosterlee, Uzawa smoother in multigrid for
the coupled porous medium and Stokes flow system, SIAM J. Sci. Comput., 39 (2017), pp.
S633–S661, doi:10.1137/16M1076514.
[48] P. Kumar, P. Luo, F. J. Gaspar, and C. W. Oosterlee, A multigrid multilevel Monte
Carlo method for transport in the Darcy-Stokes system, J. Comput. Phys., 371 (2018), pp.
382–408, doi:10.1016/j.jcp.2018.05.046.
[49] C. Rodrigo, P. Salinas, F. J. Gaspar, and F. J. Lisbona, Local Fourier analysis for cell-
centered multigrid methods on triangular grids, J. Comput. Appl. Math., 259 (2014), pp.
35–47, doi:10.1016/j.cam.2013.03.040.
[50] C. Rodrigo, F. J. Gaspar, X. Hu, and L. T. Zikatanov, Stability and monotonicity for some
discretizations of the Biot’s consolidation model, Comput. Methods Appl. Mech. Engrg.,
298 (2016), pp. 183–204, doi:10.1016/j.cma.2015.09.019.
[51] M. L. Ravalec, B. Noetinger, and L. Y. Hu, The FFT Moving Average (FFT-MA) genera-
tor: An efficient numerical method for generating and conditioning Gaussian simulations,
Math. Geol., 32 (2000), pp. 701–723, doi:10.1023/A:1007542406333.
[52] I. G. Graham, F. Y. Kuo, D. Nuyens, R. Scheichl, and I. H. Sloan, Quasi-Monte Carlo
methods for elliptic PDEs with random coefficients and applications, J. Comput. Phys.,
230 (2011), pp. 3668–3694, doi:10.1016/j.jcp.2011.01.023.

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

You might also like