This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11institutetext: J. Dick 22institutetext: School of Mathematics and Statistics, University of New South Wales, Sydney NSW 2052, Australia
22email: [email protected]
33institutetext: T. Goda 44institutetext: H. Murata 55institutetext: School of Engineering, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
55email: [email protected]
55email: [email protected]

Toeplitz Monte Carlo

Josef Dick    Takashi Goda    Hiroya Murata
(Received: date / Accepted: date)
Abstract

Motivated mainly by applications to partial differential equations with random coefficients, we introduce a new class of Monte Carlo estimators, called Toeplitz Monte Carlo (TMC) estimator for approximating the integral of a multivariate function with respect to the direct product of an identical univariate probability measure. The TMC estimator generates a sequence x1,x2,x_{1},x_{2},\ldots of i.i.d. samples for one random variable, and then uses (xn+s1,xn+s2,xn)(x_{n+s-1},x_{n+s-2}\ldots,x_{n}) with n=1,2,n=1,2,\ldots as quadrature points, where ss denotes the dimension. Although consecutive points have some dependency, the concatenation of all quadrature nodes is represented by a Toeplitz matrix, which allows for a fast matrix-vector multiplication. In this paper we study the variance of the TMC estimator and its dependence on the dimension ss. Numerical experiments confirm the considerable efficiency improvement over the standard Monte Carlo estimator for applications to partial differential equations with random coefficients, particularly when the dimension ss is large.

Keywords:
Monte Carlo Toeplitz matrix fast matrix-vector multiplication high-dimensional integration PDEs with random coefficients
MSC:
MSC 65C05

1 Introduction

The motivation of this research mainly comes from applications to uncertainty quantification for ordinary or partial differential equations with random coefficients. The problem we are interested in is to estimate an expectation (integral)

Iρs(f):=Ωsf(𝒙)ρs(𝒙)d𝒙withρs(𝒙)=j=1sρ(xj),I_{\rho_{s}}(f):=\int_{\Omega^{s}}f(\boldsymbol{x})\rho_{s}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}\quad\text{with}\quad\rho_{s}(\boldsymbol{x})=\prod_{j=1}^{s}\rho(x_{j}),

for large ss with ρ\rho being the univariate probability density function defined over Ω\Omega\subseteq\mathbb{R}. In some applications, the integrand is of the form

f(𝒙)=g(𝒙A),f(\boldsymbol{x})=g(\boldsymbol{x}A),

for a matrix As×tA\in\mathbb{R}^{s\times t} and a function g:tg\colon\mathbb{R}^{t}\to\mathbb{R}, see Dick et al. (2015). Here we note that 𝒙\boldsymbol{x} is defined as a row vector. Typically, ρ\rho is given by the uniform distribution on the unit interval Ω=[0,1]\Omega=[0,1], or by the standard normal distribution on the real line Ω=\Omega=\mathbb{R}.

The standard Monte Carlo method approximates Iρs(f)I_{\rho_{s}}(f) as follows: we first generate a sequence of i.i.d. samples of the random variables 𝒙ρs\boldsymbol{x}\sim\rho_{s}:

𝒙1=(x1,1,,xs,1),𝒙2=(x1,2,,xs,2),,\boldsymbol{x}_{1}=(x_{1,1},\ldots,x_{s,1}),\boldsymbol{x}_{2}=(x_{1,2},\ldots,x_{s,2}),\ldots,

and then approximate Iρs(f)I_{\rho_{s}}(f) by

IρsMC(f;N)=1Nn=1Nf(𝒙n)=1Nn=1Ng(𝒙nA).I^{\mathrm{MC}}_{\rho_{s}}(f;N)=\frac{1}{N}\sum_{n=1}^{N}f(\boldsymbol{x}_{n})=\frac{1}{N}\sum_{n=1}^{N}g(\boldsymbol{x}_{n}A). (1)

It is well known that

𝔼[IρsMC(f;N)]=Iρs(f)\mathbb{E}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]=I_{\rho_{s}}(f)

and

𝕍[IρsMC(f;N)]=Iρs(f2)(Iρs(f))2N,\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]=\frac{I_{\rho_{s}}(f^{2})-(I_{\rho_{s}}(f))^{2}}{N}, (2)

which ensures the canonical “one over square root of NN” convergence.

Now let us consider a situation where computing 𝒙nA\boldsymbol{x}_{n}A for n=1,,Nn=1,\ldots,N takes a significant amount of time in the computation of IρsMC(f;N)I^{\mathrm{MC}}_{\rho_{s}}(f;N). In general, if the matrix AA does not have any special structure such as circulant, Hankel, Toeplitz, or Vandermonde, then fast matrix-vector multiplication is not available and the computation of IρsMC(f;N)I^{\mathrm{MC}}_{\rho_{s}}(f;N) requires O(Nst)O(Nst) arithmetic operations. Some examples where a fast matrix-vector multiplication has been established are the following: In Feischl et al. (2018) the authors use HH-matrices to obtain an approximation of a covariance matrix which also permits a fast matrix vector multiplication; In Giles et al. (2008) the authors show how a (partially) fast matrix vector product can be implemented for multi-asset pricing in finance; Brownian bridge and principle component analysis factorizations of the covariance matrix in finance also permit a fast matrix vector multiplication (Giles et al. 2008). Here we consider the case where either a fast matrix-vector product is not available, or one wants to avoid HH-matrices and particular covariance factorizations, since we do not impose any restrictions on AA.

In order to reduce this computational cost, we propose an alternative, novel Monte Carlo estimator in this paper. Instead of generating a sequence of i.i.d. samples of the vector 𝒙\boldsymbol{x}, we generate a sequence of i.i.d. samples of a single random variable, denoted by x1,x2,x_{1},x_{2},\ldots, and then approximates Iρs(f)I_{\rho_{s}}(f) by

IρsTMC(f;N)=1Nn=1Nf(𝒙~n)I^{\mathrm{TMC}}_{\rho_{s}}(f;N)=\frac{1}{N}\sum_{n=1}^{N}f(\tilde{\boldsymbol{x}}_{n}) (3)

with

𝒙~n=(xn+s1,,xn).\tilde{\boldsymbol{x}}_{n}=(x_{n+s-1},\ldots,x_{n}).

The computation of IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N) can be done as follows:

Algorithm 1

For N>0N\in\mathbb{Z}_{>0}, let x1,x2,,xN+s1x_{1},x_{2},\ldots,x_{N+s-1} be N+s1N+s-1 i.i.d. samples of a random variable following ρ\rho.

  1. 1.

    Define XN×sX\in\mathbb{R}^{N\times s} by

    X=(xsxs1xs2x1xs+1xsxs1x2xs+2xs+1xsx3xN+s2xN+s3xN+s4xN1xN+s1xN+s2xN+s3xN).X=\begin{pmatrix}x_{s}&x_{s-1}&x_{s-2}&\cdots&x_{1}\\ x_{s+1}&x_{s}&x_{s-1}&\cdots&x_{2}\\ x_{s+2}&x_{s+1}&x_{s}&\cdots&x_{3}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ x_{N+s-2}&x_{N+s-3}&x_{N+s-4}&\cdots&x_{N-1}\\ x_{N+s-1}&x_{N+s-2}&x_{N+s-3}&\cdots&x_{N}\\ \end{pmatrix}.

    Note that XX is a Toeplitz matrix.

  2. 2.

    Compute

    XA=Y=(𝒚1𝒚N)N×t.XA=Y=\begin{pmatrix}\boldsymbol{y}_{1}\\ \vdots\\ \boldsymbol{y}_{N}\end{pmatrix}\in\mathbb{R}^{N\times t}.
  3. 3.

    Then IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N) is given by

    IρsTMC(f;N)=1Nn=1Ng(𝒚n).I^{\mathrm{TMC}}_{\rho_{s}}(f;N)=\frac{1}{N}\sum_{n=1}^{N}g(\boldsymbol{y}_{n}).

The idea behind introducing this algorithm comes from a recent paper by Dick et al. (2015) who consider replacing the point set used in the standard Monte Carlo estimator (1) with a special class of quasi-Monte Carlo point sets which permit a fast matrix-vector multiplication 𝒙nA\boldsymbol{x}_{n}A for n=1,,Nn=1,\ldots,N. This paper considers a sampling scheme different from Dick et al. (2015) while still allowing for a fast matrix-vector multiplication.

When ss is quite large, say thousands or million, NN has to be set significantly smaller than 2s2^{s}. Throughout this paper we consider the case where NsκN\approx s^{\kappa} for some κ>0\kappa>0. Since the matrix-vector multiplication between a Toeplitz matrix XX and each column vector of AA can be done with O(Nlogs)O(N\log s) arithmetic operations by using the fast Fourier transform (Frigo and Johnson 2005), the matrix-matrix multiplication XAXA appearing in the second item of Algorithm 1 can be done with O(tNlogs)O(tN\log s) arithmetic operations. This way the necessary computational cost can be reduced from O(Nst)O(Nst) to O(tNlogs)O(tN\log s), which is the major advantage of using IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N).

Remark 1

Here we give some comments about memory requirements and parallel implementation of the estimators. For the standard Monte Carlo estimator, since each sample 𝒙n\boldsymbol{x}_{n} is generated independently, the corresponding function values f(𝒙n)=g(𝒙nA)f(\boldsymbol{x}_{n})=g(\boldsymbol{x}_{n}A) can be computed in parallel. A required size to keep one vector 𝒙nA\boldsymbol{x}_{n}A in memory until evaluating g(𝒙nA)g(\boldsymbol{x}_{n}A) is only of order tt. Regarding the TMC estimator, let us assume that NN is a multiple of ss, that is, N=LsN=Ls with some L>0L\in\mathbb{Z}_{>0}. For each {1,,L}\ell\in\{1,\ldots,L\}, define a Toeplitz submatrix

X=(xsxs1x(1)s+1xs+1xsx(1)s+2xs+s1xs+s2xs)s×s.X_{\ell}=\begin{pmatrix}x_{\ell s}&x_{\ell s-1}&\cdots&x_{(\ell-1)s+1}\\ x_{\ell s+1}&x_{\ell s}&\cdots&x_{(\ell-1)s+2}\\ \vdots&\vdots&\ddots&\vdots\\ x_{\ell s+s-1}&x_{\ell s+s-2}&\cdots&x_{\ell s}\\ \end{pmatrix}\in\mathbb{R}^{s\times s}.

In fact, it is easy to see that X=(X1,,XL)X=\left(X_{1}^{\top},\ldots,X_{L}^{\top}\right)^{\top} and

Y=((X1A),,(XLA))N×t.Y=\left((X_{1}A)^{\top},\ldots,(X_{L}A)^{\top}\right)^{\top}\in\mathbb{R}^{N\times t}.

This clearly shows that each XAX_{\ell}A can be computed in parallel by using the fast Fourier transform. Given that the matrix XAs×tX_{\ell}A\in\mathbb{R}^{s\times t} has to be kept in memory until evaluating g(𝒚n)g(\boldsymbol{y}_{n}) for n=(1)s+1,,sn=(\ell-1)s+1,\ldots,\ell s, the required memory size of the TMC estimator is of order stst.

In this paper we call IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N) a Toeplitz Monte Carlo (TMC) estimator of Iρs(f)I_{\rho_{s}}(f) as we rely on the Toeplitz structure of XX to achieve a faster computation.111If the sample nodes are given by 𝒙~n=(xn,,xn+s1)\tilde{\boldsymbol{x}}_{n}=(x_{n},\ldots,x_{n+s-1}) instead of 𝒙~n=(xn+s1,,xn)\tilde{\boldsymbol{x}}_{n}=(x_{n+s-1},\ldots,x_{n}), the matrix XX becomes a Hankel matrix, which also allows for a fast matrix-vector multiplication. Therefore we can call our proposal a Hankel Monte Carlo (HMC) estimator instead. However, in the context of Monte Carlo methods, HMC often refers to the Hamiltonian Monte Carlo algorithm, and we would like to avoid duplication of the abbreviations by coining the name Toeplitz Monte Carlo. In the remainder of this paper, we study the variance of the TMC estimator and its dependence on the dimension ss, and also see practical efficiency of the TMC estimator by carrying out numerical experiments for applications from ordinary/partial differential equations with random coefficients.

2 Theoretical results

2.1 Variance analysis

In order to study the variance of IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N), we introduce the concept of the analysis-of-variance (ANOVA) decomposition of multivariate functions (Hoeffding 1948, Kuo et al. 2010, Sobol’ 1993). In what follows, for simplicity of notation, we write [1:s]={1,,s}[1:s]=\{1,\ldots,s\}. For a subset u[1:s]u\subseteq[1:s], we write u:=[1:s]u-u:=[1:s]\setminus u and denote the cardinality of uu by |u||u|. Let ff be a square-integrable function, i.e., Iρs(f2)<I_{\rho_{s}}(f^{2})<\infty. Then ff can be decomposed into

f(𝒙)=u[1:s]fu(𝒙u),f(\boldsymbol{x})=\sum_{u\subseteq[1:s]}f_{u}(\boldsymbol{x}_{u}),

where we write 𝒙u=(xj)ju\boldsymbol{x}_{u}=(x_{j})_{j\in u} and each summand is defined recursively by f=Iρs(f)f_{\emptyset}=I_{\rho_{s}}(f) and

fu(𝒙u)=Ωs|u|f(𝒙)ρs|u|(𝒙u)d𝒙uvufv(𝒙v)f_{u}(\boldsymbol{x}_{u})=\int_{\Omega^{s-|u|}}f(\boldsymbol{x})\rho_{s-|u|}(\boldsymbol{x}_{-u})\,\mathrm{d}\boldsymbol{x}_{-u}-\sum_{v\subset u}f_{v}(\boldsymbol{x}_{v})

for u[1:s]\emptyset\neq u\subseteq[1:s]. Regarding this decomposition of multivariate functions, the following properties hold. We refer to Lemmas A.1 & A.3 of Owen (2019) for the proof of the case where ρ\rho is the uniform distribution over the unit interval Ω=[0,1]\Omega=[0,1].

Lemma 1

With the notation above, we have:

  1. 1.

    For any non-empty u[1:s]u\subseteq[1:s] and juj\in u,

    Ωfu(𝒙u)ρ(xj)dxj=0.\int_{\Omega}f_{u}(\boldsymbol{x}_{u})\rho(x_{j})\,\mathrm{d}x_{j}=0.
  2. 2.

    For any u,v[1:s]u,v\subseteq[1:s],

    Iρs(fufv)\displaystyle I_{\rho_{s}}(f_{u}f_{v}) =Ωsfu(𝒙u)fv(𝒙v)ρs(𝒙)d𝒙\displaystyle=\int_{\Omega^{s}}f_{u}(\boldsymbol{x}_{u})f_{v}(\boldsymbol{x}_{v})\rho_{s}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}
    ={Iρs(fu2)if u=v,0otherwise.\displaystyle=\begin{cases}I_{\rho_{s}}(f_{u}^{2})&\text{if $u=v$,}\\ 0&\text{otherwise.}\end{cases}

It follows from the second assertion of Lemma 1 that

Iρs(f2)\displaystyle I_{\rho_{s}}(f^{2}) =Iρs(u,v[1:s]fufv)\displaystyle=I_{\rho_{s}}\left(\sum_{u,v\subseteq[1:s]}f_{u}f_{v}\right)
=u,v[1:s]Iρs(fufv)=u[1:s]Iρs(fu2).\displaystyle=\sum_{u,v\subseteq[1:s]}I_{\rho_{s}}(f_{u}f_{v})=\sum_{u\subseteq[1:s]}I_{\rho_{s}}(f_{u}^{2}).

This equality means that the variance of ff can be expressed as a sum of the variances of the lower-dimensional functions:

Iρs(f2)(Iρs(f))2=u[1:s]Iρs(fu2).I_{\rho_{s}}(f^{2})-(I_{\rho_{s}}(f))^{2}=\sum_{\emptyset\neq u\subseteq[1:s]}I_{\rho_{s}}(f_{u}^{2}). (4)

Using these facts, the variance of the TMC estimator IρsTMC(f;N)I^{\mathrm{TMC}}_{\rho_{s}}(f;N) can be analyzed as follows:

Theorem 2.1

Let N,s2N,s\in\mathbb{Z}_{\geq 2}. Then we have

𝔼[IρsTMC(f;N)]=Iρs(f)\mathbb{E}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]=I_{\rho_{s}}(f)

and

𝕍[IρsTMC(f;N)]=𝕍[IρsMC(f;N)]\displaystyle\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]=\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]
+2N2=1min(s,N)1(N)u[1:s]Iρ|u|(fufu+),\displaystyle\quad+\frac{2}{N^{2}}\sum_{\ell=1}^{\min(s,N)-1}(N-\ell)\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell}),

where we write u+={j+:ju}u+\ell=\{j+\ell\colon j\in u\}.

Note that, in the theorem, we write

Iρ|u|(fufu+)=Ω|u|fu(𝒙u)fu+(𝒙u)ρ|u|(𝒙u)d𝒙u.I_{\rho_{|u|}}(f_{u}f_{u+\ell})=\int_{\Omega^{|u|}}f_{u}(\boldsymbol{x}_{u})f_{u+\ell}(\boldsymbol{x}_{u})\rho_{|u|}(\boldsymbol{x}_{u})\,\mathrm{d}\boldsymbol{x}_{u}.

The readers should not be confused with

Iρs(fufu+)=Ωsfu(𝒙u)fu+(𝒙u+)ρs(𝒙)d𝒙=0.I_{\rho_{s}}(f_{u}f_{u+\ell})=\int_{\Omega^{s}}f_{u}(\boldsymbol{x}_{u})f_{u+\ell}(\boldsymbol{x}_{u+\ell})\rho_{s}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}=0.
Proof

The first assertion follows immediately from the linearity of expectation and the trivial equality 𝔼[f(𝒙~n)]=Iρs(f)\mathbb{E}[f(\tilde{\boldsymbol{x}}_{n})]=I_{\rho_{s}}(f). For the second assertion, by using the ANOVA decomposition of ff, we have

𝕍[IρsTMC(f;N)]\displaystyle\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]
=𝔼[(IρsTMC(f;N))2](𝔼[IρsTMC(f;N)])2\displaystyle=\mathbb{E}\left[(I^{\mathrm{TMC}}_{\rho_{s}}(f;N))^{2}\right]-\left(\mathbb{E}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]\right)^{2}
=𝔼[(IρsTMC(f;N))2](Iρs(f))2\displaystyle=\mathbb{E}\left[(I^{\mathrm{TMC}}_{\rho_{s}}(f;N))^{2}\right]-\left(I_{\rho_{s}}(f)\right)^{2}
=1N2m,n=1N𝔼[f(𝒙~m)f(𝒙~n)](Iρs(f))2\displaystyle=\frac{1}{N^{2}}\sum_{m,n=1}^{N}\mathbb{E}[f(\tilde{\boldsymbol{x}}_{m})f(\tilde{\boldsymbol{x}}_{n})]-\left(I_{\rho_{s}}(f)\right)^{2}
=1N2n=1N𝔼[(f(𝒙~n))2]+2N2m,n=1m>nN𝔼[f(𝒙~m)f(𝒙~n)]\displaystyle=\frac{1}{N^{2}}\sum_{n=1}^{N}\mathbb{E}[(f(\tilde{\boldsymbol{x}}_{n}))^{2}]+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\mathbb{E}[f(\tilde{\boldsymbol{x}}_{m})f(\tilde{\boldsymbol{x}}_{n})]
(Iρs(f))2\displaystyle\quad-\left(I_{\rho_{s}}(f)\right)^{2}
=Iρs(f2)N+2N2m,n=1m>nNu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle=\frac{I_{\rho_{s}}(f^{2})}{N}+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
(Iρs(f))2.\displaystyle\quad-\left(I_{\rho_{s}}(f)\right)^{2}.

It follows from the first assertion of Lemma 1 that the second term on the right-most side above becomes

2N2m,n=1m>nNu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=2N2m,n=1m>nNf2+2N2m,n=1m>nNu[1:s]f𝔼[fu(𝒙~m,u)]\displaystyle=\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}f_{\emptyset}^{2}+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{\emptyset\neq u\subseteq[1:s]}f_{\emptyset}\,\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})]
+2N2m,n=1m>nNv[1:s]f𝔼[fv(𝒙~n,v)]\displaystyle\quad+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{\emptyset\neq v\subseteq[1:s]}f_{\emptyset}\,\mathbb{E}[f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
+2N2m,n=1m>nNu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\quad+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=N1N(Iρs(f))2\displaystyle=\frac{N-1}{N}\left(I_{\rho_{s}}(f)\right)^{2}
+2N2m,n=1m>nNu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\quad+\frac{2}{N^{2}}\sum_{\begin{subarray}{c}m,n=1\\ m>n\end{subarray}}^{N}\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=N1N(Iρs(f))2\displaystyle=\frac{N-1}{N}\left(I_{\rho_{s}}(f)\right)^{2}
+2N2=1N1m,n=1mn=Nu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)],\displaystyle\quad+\frac{2}{N^{2}}\sum_{\ell=1}^{N-1}\sum_{\begin{subarray}{c}m,n=1\\ m-n=\ell\end{subarray}}^{N}\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})],

where we reordered the sum over mm and nn with respect to the difference mnm-n in the last equality.

If mnsm-n\geq s, there is no overlapping of the components between 𝒙~m\tilde{\boldsymbol{x}}_{m} and 𝒙~n\tilde{\boldsymbol{x}}_{n}. Because of the independence of samples, it follows from the first assertion of Lemma 1 that the inner sum over uu and vv above is given by

u,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=u,v[1:s]𝔼[fu(𝒙~m,u)]𝔼[fv(𝒙~n,v)]=0.\displaystyle=\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})]\mathbb{E}[f_{v}(\tilde{\boldsymbol{x}}_{n,v})]=0.

If =mn<s\ell=m-n<s, on the other hand, we have x~n,j=x~m,j+\tilde{x}_{n,j}=\tilde{x}_{m,j+\ell} for any j=1,,sj=1,\ldots,s-\ell. With this equality and the first assertion of Lemma 1, the inner sum over uu and vv becomes

u,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=u[1:s]𝔼[fu+(𝒙~m,u+)fu(𝒙~n,u)]\displaystyle=\sum_{\emptyset\neq u\subseteq[1:s-\ell]}\mathbb{E}[f_{u+\ell}(\tilde{\boldsymbol{x}}_{m,u+\ell})f_{u}(\tilde{\boldsymbol{x}}_{n,u})]
=u[1:s]𝔼[fu+(𝒙~n,u)fu(𝒙~n,u)]\displaystyle=\sum_{\emptyset\neq u\subseteq[1:s-\ell]}\mathbb{E}[f_{u+\ell}(\tilde{\boldsymbol{x}}_{n,u})f_{u}(\tilde{\boldsymbol{x}}_{n,u})]
=u[1:s]Iρ|u|(fu+fu).\displaystyle=\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u+\ell}f_{u}).

Altogether we obtain

𝕍[IρsTMC(f;N)]\displaystyle\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]
=Iρs(f2)N+N1N(Iρs(f))2(Iρs(f))2\displaystyle=\frac{I_{\rho_{s}}(f^{2})}{N}+\frac{N-1}{N}\left(I_{\rho_{s}}(f)\right)^{2}-\left(I_{\rho_{s}}(f)\right)^{2}
+2N2=1N1m,n=1mn=Nu,v[1:s]𝔼[fu(𝒙~m,u)fv(𝒙~n,v)]\displaystyle\quad+\frac{2}{N^{2}}\sum_{\ell=1}^{N-1}\sum_{\begin{subarray}{c}m,n=1\\ m-n=\ell\end{subarray}}^{N}\sum_{\emptyset\neq u,v\subseteq[1:s]}\mathbb{E}[f_{u}(\tilde{\boldsymbol{x}}_{m,u})f_{v}(\tilde{\boldsymbol{x}}_{n,v})]
=Iρs(f2)(Iρs(f))2N\displaystyle=\frac{I_{\rho_{s}}(f^{2})-\left(I_{\rho_{s}}(f)\right)^{2}}{N}
+2N2=1min(s,N)1u[1:s]Iρ|u|(fufu+)m,n=1mn=N1\displaystyle\quad+\frac{2}{N^{2}}\sum_{\ell=1}^{\min(s,N)-1}\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell})\sum_{\begin{subarray}{c}m,n=1\\ m-n=\ell\end{subarray}}^{N}1
=𝕍[IρsMC(f;N)]\displaystyle=\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]
+2N2=1min(s,N)1(N)u[1:s]Iρ|u|(fufu+).\displaystyle\quad+\frac{2}{N^{2}}\sum_{\ell=1}^{\min(s,N)-1}(N-\ell)\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell}).

Thus we are done.∎

Remark 2

We now consider a parallel implementation of the TMC method with LL CPUs. Each CPU independently generates a Toeplitz matrix of NN random numbers and CPU vv computes an estimator IρsTMC(f;N;v)I^{\mathrm{TMC}}_{\rho_{s}}(f;N;v) of the form (3), v=1,2,,Lv=1,2,\ldots,L. In this case, Theorem 2.1 then applies to the output of each CPU in the parallel implementation. The average of these LL independent estimators is again unbiased and satisfies

𝕍[1Lv=1LIρsTMC(f;N;v)]=1L𝕍[IρsMC(f;N)]\displaystyle\mathbb{V}\left[\frac{1}{L}\sum_{v=1}^{L}I^{\mathrm{TMC}}_{\rho_{s}}(f;N;v)\right]=\frac{1}{L}\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]
+2LN2=1min(s,N)1(N)u[1:s]Iρ|u|(fufu+).\displaystyle\quad+\frac{2}{LN^{2}}\sum_{\ell=1}^{\min(s,N)-1}(N-\ell)\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell}).

Notice that we have LNLN points altogether.

If L=1L=1 then we get the result from Theorem 2.1 and if N=1N=1 we get (2) (where, in this case, LL is the number of samples), as the sum =10\sum_{\ell=1}^{0}\ldots is equal to 0 in this instance.

As is clear from Theorem 2.1, the TMC estimator is unbiased and maintains the canonical “one over square root of NN” convergence. Moreover, the TMC estimator can be regarded as a variance reduction technique since the second term on the variance 𝕍[IρsTMC(f;N)]\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)] can be negative, depending on the function.

Example 1

To illustrate the last comment, let us consider a simple test function f:3f\colon\mathbb{R}^{3}\to\mathbb{R} given by

f(x,y,z)=xyz+xyxzyz,f(x,y,z)=x-y-z+xy-xz-yz,

and let x,y,zx,y,z be normally distributed independent random variables with mean 0 and variance 1. It is easy to see that

f{1}(x)=x,f{2}(y)=y,f{3}(z)=z,\displaystyle f_{\{1\}}(x)=x,\quad f_{\{2\}}(y)=-y,\quad f_{\{3\}}(z)=-z,
f{1,2}(x,y)=xy,f{1,3}(x,z)=xz,\displaystyle f_{\{1,2\}}(x,y)=xy,\quad f_{\{1,3\}}(x,z)=-xz,
f{2,3}(y,z)=yz,f{1,2,3}(x,y,z)=0.\displaystyle f_{\{2,3\}}(y,z)=-yz,\quad f_{\{1,2,3\}}(x,y,z)=0.

Then it follows that

𝕍[IρsMC(f;N)]=1Nu[1:3]Iρs(fu2)=6N,\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]=\frac{1}{N}\sum_{\emptyset\neq u\subseteq[1:3]}I_{\rho_{s}}(f_{u}^{2})=\frac{6}{N},

whereas, for N3N\geq 3, we have

𝕍[IρsTMC(f;N)]\displaystyle\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]
=6N+2(N2)N2Iρ1(f{1}f{3})+2(N1)N2\displaystyle=\frac{6}{N}+\frac{2(N-2)}{N^{2}}I_{\rho_{1}}(f_{\{1\}}f_{\{3\}})+\frac{2(N-1)}{N^{2}}
×[Iρ1(f{1}f{2})+Iρ1(f{2}f{3})+Iρ2(f{1,2}f{2,3})]\displaystyle\quad\times\left[I_{\rho_{1}}(f_{\{1\}}f_{\{2\}})+I_{\rho_{1}}(f_{\{2\}}f_{\{3\}})+I_{\rho_{2}}(f_{\{1,2\}}f_{\{2,3\}})\right]
=6N2(N2)N22(N1)N2=2N+6N2.\displaystyle=\frac{6}{N}-\frac{2(N-2)}{N^{2}}-\frac{2(N-1)}{N^{2}}=\frac{2}{N}+\frac{6}{N^{2}}.

Therefore the variance of the TMC estimator is almost one-third of the variance of the standard Monte Carlo estimator.

It is also possible that the variance of the TMC estimator increases compared to standard Monte Carlo, however, we show below that this increase is bounded.

2.2 Weighted L2L_{2} space and tractability

Here we study the dependence of the variance 𝕍[IρsTMC(f;N)]\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)] on the dimension ss. For this purpose, we first give a bound on 𝕍[IρsTMC(f;N)]\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]. For 1s1\leq\ell\leq s, let

α(f):=(u[1:s]minjuj=Iρ|u|(fu2))1/2.\alpha_{\ell}(f):=\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}.

Then it follows from the decomposition (4) that

Iρs(f2)(Iρs(f))2==1su[1:s]minjuj=Iρs(fu2)==1s(α(f))2,I_{\rho_{s}}(f^{2})-(I_{\rho_{s}}(f))^{2}=\sum_{\ell=1}^{s}\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}I_{\rho_{s}}(f_{u}^{2})=\sum_{\ell=1}^{s}(\alpha_{\ell}(f))^{2},

resulting in an equality

𝕍[IρsMC(f;N)]=1N=1s(α(f))2.\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]=\frac{1}{N}\sum_{\ell=1}^{s}(\alpha_{\ell}(f))^{2}.

Using Theorem 2.1, the variance 𝕍[IρsTMC(f;N)]\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)] is bounded above as follows.

Corollary 1

We have

𝕍[IρsTMC(f;N)]1N(=1sα(f))2.\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]\leq\frac{1}{N}\left(\sum_{\ell=1}^{s}\alpha_{\ell}(f)\right)^{2}.
Proof

For any =1,,s1\ell=1,\ldots,s-1, the Cauchy-Schwarz inequality leads to

u[1:s]Iρ|u|(fufu+)\displaystyle\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell})
u[1:s](Iρ|u|(fu2))1/2(Iρ|u|(fu+2))1/2\displaystyle\leq\sum_{\emptyset\neq u\subseteq[1:s-\ell]}\left(I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}\left(I_{\rho_{|u|}}(f_{u+\ell}^{2})\right)^{1/2}
=v=1su[1:s]minjuj=v(Iρ|u|(fu2))1/2(Iρ|u|(fu+2))1/2\displaystyle=\sum_{v=1}^{s-\ell}\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s-\ell]\\ \min_{j\in u}j=v\end{subarray}}\left(I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}\left(I_{\rho_{|u|}}(f_{u+\ell}^{2})\right)^{1/2}
v=1s(u[1:s]minjuj=vIρ|u|(fu2))1/2\displaystyle\leq\sum_{v=1}^{s-\ell}\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s-\ell]\\ \min_{j\in u}j=v\end{subarray}}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}
×(u[1:s]minjuj=vIρ|u|(fu+2))1/2\displaystyle\qquad\times\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s-\ell]\\ \min_{j\in u}j=v\end{subarray}}I_{\rho_{|u|}}(f_{u+\ell}^{2})\right)^{1/2}
=v=1s(u[1:s]minjuj=vIρ|u|(fu2))1/2\displaystyle=\sum_{v=1}^{s-\ell}\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s-\ell]\\ \min_{j\in u}j=v\end{subarray}}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}
×(u[+1:s]minjuj=+vIρ|u|(fu2))1/2\displaystyle\qquad\times\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[\ell+1:s]\\ \min_{j\in u}j=\ell+v\end{subarray}}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}
v=1sαv(f)α+v(f).\displaystyle\leq\sum_{v=1}^{s-\ell}\alpha_{v}(f)\alpha_{\ell+v}(f).

Applying this bound to the second assertion of Theorem 2.1, we obtain

𝕍[IρsTMC(f;N)]\displaystyle\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]
1N=1s(α(f))2+2N2=1s1(N)v=1sαv(f)α+v(f)\displaystyle\leq\frac{1}{N}\sum_{\ell=1}^{s}(\alpha_{\ell}(f))^{2}+\frac{2}{N^{2}}\sum_{\ell=1}^{s-1}(N-\ell)\sum_{v=1}^{s-\ell}\alpha_{v}(f)\alpha_{\ell+v}(f)
1N=1s(α(f))2+2Nv=1s1αv(f)=v+1sα(f)\displaystyle\leq\frac{1}{N}\sum_{\ell=1}^{s}(\alpha_{\ell}(f))^{2}+\frac{2}{N}\sum_{v=1}^{s-1}\alpha_{v}(f)\sum_{\ell=v+1}^{s}\alpha_{\ell}(f)
=1N(=1sα(f))2.\displaystyle=\frac{1}{N}\left(\sum_{\ell=1}^{s}\alpha_{\ell}(f)\right)^{2}.

Using this result, we have

𝕍[IρsTMC(f;N)]𝕍[IρsMC(f;N)](α1(f)++αs(f))2α12(f)++αs2(f)s,\frac{\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]}{\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]}\leq\frac{(\alpha_{1}(f)+\cdots+\alpha_{s}(f))^{2}}{\alpha^{2}_{1}(f)+\cdots+\alpha^{2}_{s}(f)}\leq s,

wherein, for the second inequality, the equality is attained if and only if α1(f)==αs(f)\alpha_{1}(f)=\cdots=\alpha_{s}(f). Therefore, when we fix the number of samples, the variance of the TMC estimator can at most be ss times larger than the variance of the standard Monte Carlo estimator.

Now let us consider the case s=ts=t and assume, as discussed in the first section, that the computational time for the standard Monte Carlo estimator is proportional to Ns2Ns^{2}, whereas the computational time for the TMC estimator is proportional to NslogsNs\log s (assuming that the main cost in evaluating f(𝒙)=g(𝒙A)f(\boldsymbol{x})=g(\boldsymbol{x}A) lies in the computation of 𝒙A\boldsymbol{x}A). When we fix the cost instead of the number of samples, we have

NMCs2NTMCslogs,N_{\mathrm{MC}}s^{2}\asymp N_{\mathrm{TMC}}s\log s,

where \asymp indicates that the terms should be of the same order, and so

𝕍[IρsTMC(f;NTMC)]𝕍[IρsMC(f;NMC)]\displaystyle\frac{\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N_{\mathrm{TMC}})]}{\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N_{\mathrm{MC}})]} NMCNTMC(α1(f)++αs(f))2α12(f)++αs2(f)\displaystyle\leq\frac{N_{\mathrm{MC}}}{N_{\mathrm{TMC}}}\cdot\frac{(\alpha_{1}(f)+\cdots+\alpha_{s}(f))^{2}}{\alpha^{2}_{1}(f)+\cdots+\alpha^{2}_{s}(f)}
logs.\displaystyle\propto\log s.

Thus, the variance of the TMC estimator for a given cost is at most logs\log s times as large as the standard Monte Carlo estimator (up to some constant factor). On the other hand, if there is some decay of the importance of the ANOVA terms as the index of the variable increases, for instance, if the first few terms in α1(f)++αs(f)\alpha_{1}(f)+\cdots+\alpha_{s}(f) dominate the sum, then the ratio (α1(f)++αs(f))2α12(f)++αs2(f)\frac{(\alpha_{1}(f)+\cdots+\alpha_{s}(f))^{2}}{\alpha^{2}_{1}(f)+\cdots+\alpha^{2}_{s}(f)} can be bounded independently of ss, leading to a gain in the efficiency of the TMC estimator. We observe such a behaviour in our numerical experiments below.

Following the idea from Sloan and Woźniakowski (1998), we now introduce the notion of a weighted L2L_{2} space. Let (γu)u(\gamma_{u})_{u\subset\mathbb{N}} be a sequence of the non-negative real numbers called weights. Then the weighted L2L_{2} space is defined by

s,𝜸={f:Ωsfs,𝜸},\mathcal{F}_{s,\boldsymbol{\gamma}}=\left\{f\colon\Omega^{s}\to\mathbb{R}\mid\|f\|_{s,\boldsymbol{\gamma}}\leq\infty\right\},

where

fs,𝜸:=(u[1:s]γu1Iρs(fu2))1/2.\|f\|_{s,\boldsymbol{\gamma}}:=\left(\sum_{u\subseteq[1:s]}\gamma_{u}^{-1}I_{\rho_{s}}(f_{u}^{2})\right)^{1/2}.

For any subset uu with γu=0\gamma_{u}=0, we assume that the corresponding ANOVA term fuf_{u} is 0 and we formally set 0/0=00/0=0.

For a randomized algorithm using NN function evaluations of ff to estimate Iρs(f)I_{\rho_{s}}(f), which we denote by Alg(f;N)\mathrm{Alg}(f;N), let us consider the minimal cost to estimate Iρs(f)I_{\rho_{s}}(f) with mean square error ε2\varepsilon^{2} for any ff in the unit ball of s,𝜸\mathcal{F}_{s,\boldsymbol{\gamma}}:

N(ε,Alg):=min{N>0e2(Alg,s,𝜸)ε2},N(\varepsilon,\mathrm{Alg}):=\min\left\{N\in\mathbb{Z}_{>0}\mid e^{2}(\mathrm{Alg},\mathcal{F}_{s,\boldsymbol{\gamma}})\leq\varepsilon^{2}\right\},

where

e2(Alg,s,𝜸):=supfs,𝜸fs,𝜸1𝔼[(Alg(f;N)Iρs(f))2].e^{2}(\mathrm{Alg},\mathcal{F}_{s,\boldsymbol{\gamma}}):=\sup_{\begin{subarray}{c}f\in\mathcal{F}_{s,\boldsymbol{\gamma}}\\ \|f\|_{s,\boldsymbol{\gamma}}\leq 1\end{subarray}}\mathbb{E}\left[\left(\mathrm{Alg}(f;N)-I_{\rho_{s}}(f)\right)^{2}\right].

We say that the algorithm Alg\mathrm{Alg} is

  • a weakly tractable algorithm if

    limε1+slnN(ε,Alg)ε1+s=0,\lim_{\varepsilon^{-1}+s\to\infty}\frac{\ln N(\varepsilon,\mathrm{Alg})}{\varepsilon^{-1}+s}=0,
  • a polynomially tractable algorithm if there exist non-negative constants C,p,qC,p,q, such that

    N(ε,Alg)CεpsqN(\varepsilon,\mathrm{Alg})\leq C\varepsilon^{-p}s^{q}

    holds for all s=1,2,s=1,2,\ldots, where pp and qq are called the ε1\varepsilon^{-1}-exponent and the ss-exponent, respectively,

  • a strongly polynomially tractable algorithm if Alg\mathrm{Alg} is a polynomially tractable algorithm with the ss-exponent 0.

We refer to Novak and Woźniakowski (2010) for more information on the notion of tractability.

For instance, the standard Monte Carlo estimator IρsMC(f;N)I^{\mathrm{MC}}_{\rho_{s}}(f;N) is a strongly polynomially tractable algorithm with ε1\varepsilon^{-1}-exponent 2 if

supuγu<\sup_{u\subset\mathbb{N}}\gamma_{u}<\infty (5)

holds. This claim can be proven as follows:

𝔼[(IρsMC(f;N)Iρs(f))2]\displaystyle\mathbb{E}\left[\left(I^{\mathrm{MC}}_{\rho_{s}}(f;N)-I_{\rho_{s}}(f)\right)^{2}\right]
=𝕍[IρsMC(f;N)]=1Nu[1:s]Iρs(fu2)\displaystyle=\mathbb{V}[I^{\mathrm{MC}}_{\rho_{s}}(f;N)]=\frac{1}{N}\sum_{\emptyset\neq u\subseteq[1:s]}I_{\rho_{s}}(f_{u}^{2})
1N(maxu[1:s]γu)(u[1:s]Iρs(fu2)γu)\displaystyle\leq\frac{1}{N}\left(\max_{u\subseteq[1:s]}\gamma_{u}\right)\left(\sum_{u\subseteq[1:s]}\frac{I_{\rho_{s}}(f_{u}^{2})}{\gamma_{u}}\right)
=fs,𝜸2Nmaxu[1:s]γu.\displaystyle=\frac{\|f\|_{s,\boldsymbol{\gamma}}^{2}}{N}\max_{u\subseteq[1:s]}\gamma_{u}.

It follows that, in order to have

𝔼[(IρsMC(f;N)Iρs(f))2]ε2\mathbb{E}\left[\left(I^{\mathrm{MC}}_{\rho_{s}}(f;N)-I_{\rho_{s}}(f)\right)^{2}\right]\leq\varepsilon^{2}

for any fs,𝜸f\in\mathcal{F}_{s,\boldsymbol{\gamma}} with fs,𝜸1\|f\|_{s,\boldsymbol{\gamma}}\leq 1, we need Nε2maxu[1:s]γuN\geq\varepsilon^{-2}\max_{u\subseteq[1:s]}\gamma_{u}. Thus the minimal cost is bounded above by

N(ε,IρsMC)ε2maxu[1:s]γu.N(\varepsilon,I^{\mathrm{MC}}_{\rho_{s}})\leq\varepsilon^{-2}\max_{u\subseteq[1:s]}\gamma_{u}.

Given the condition (5), we see that N(ε,IρsMC)N(\varepsilon,I^{\mathrm{MC}}_{\rho_{s}}) is bounded independently of the dimension ss and the algorithm IρsMCI^{\mathrm{MC}}_{\rho_{s}} is strongly polynomially tractable with the ε1\varepsilon^{-1}-exponent 2.

The following theorem gives the necessary conditions on the weights (γu)u(\gamma_{u})_{u\subset\mathbb{N}} for the TMC estimator to be a weakly tractable algorithm, a polynomially tractable algorithm, or a strongly polynomially tractable algorithm.

Theorem 2.2

The TMC estimator is

  • a weakly tractable algorithm if

    lims1sln(=1smaxu[1:s]minjuj=γu)=0,\lim_{s\to\infty}\frac{1}{s}\ln\left(\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}u\subset[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}\right)=0,
  • a polynomially tractable algorithm with the ε1\varepsilon^{-1}-exponent 2 if there exists q>0q>0 such that

    sups=1,2,1sq=1smaxu[1:s]minjuj=γu<,\sup_{s=1,2,\ldots}\frac{1}{s^{q}}\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}u\subset[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}<\infty,
  • a strongly polynomially tractable algorithm with the ε1\varepsilon^{-1}-exponent 2 if

    =1supuminjuj=γu<.\sum_{\ell=1}^{\infty}\sup_{\begin{subarray}{c}u\subset\mathbb{N}\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}<\infty.
Proof

It follows from Corollary 1 and Hölder’s inequality for sums that

𝔼[(IρsTMC(f;N)Iρs(f))2]\displaystyle\mathbb{E}\left[\left(I^{\mathrm{TMC}}_{\rho_{s}}(f;N)-I_{\rho_{s}}(f)\right)^{2}\right]
=𝕍[IρsTMC(f;N)]1N(=1sα(f))2\displaystyle=\mathbb{V}[I^{\mathrm{TMC}}_{\rho_{s}}(f;N)]\leq\frac{1}{N}\left(\sum_{\ell=1}^{s}\alpha_{\ell}(f)\right)^{2}
=1N(=1s(u[1:s]minjuj=Iρ|u|(fu2))1/2)2\displaystyle=\frac{1}{N}\left(\sum_{\ell=1}^{s}\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}\right)^{2}
1N(=1s(maxu[1:s]minjuj=γu)1/2\displaystyle\leq\frac{1}{N}\left(\sum_{\ell=1}^{s}\left(\max_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}\right)^{1/2}\right.
×(u[1:s]minjuj=γu1Iρ|u|(fu2))1/2)2\displaystyle\quad\times\left.\left(\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}^{-1}I_{\rho_{|u|}}(f_{u}^{2})\right)^{1/2}\right)^{2}
1N(=1smaxu[1:s]minjuj=γu)(=1su[1:s]minjuj=γu1Iρ|u|(fu2))\displaystyle\leq\frac{1}{N}\left(\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}\right)\left(\sum_{\ell=1}^{s}\sum_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}^{-1}I_{\rho_{|u|}}(f_{u}^{2})\right)
fs,𝜸2N=1smaxu[1:s]minjuj=γu.\displaystyle\leq\frac{\|f\|_{s,\boldsymbol{\gamma}}^{2}}{N}\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}.

Thus, the minimal cost to have

𝔼[(IρsTMC(f;N)Iρs(f))2]ε2\mathbb{E}\left[\left(I^{\mathrm{TMC}}_{\rho_{s}}(f;N)-I_{\rho_{s}}(f)\right)^{2}\right]\leq\varepsilon^{2}

for any fs,𝜸f\in\mathcal{F}_{s,\boldsymbol{\gamma}} with fs,𝜸1\|f\|_{s,\boldsymbol{\gamma}}\leq 1 is bounded above by

N(ε,IρsTMC)ε2=1smaxu[1:s]minjuj=γu.N(\varepsilon,I^{\mathrm{TMC}}_{\rho_{s}})\leq\varepsilon^{-2}\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}.

Let us consider the first assertion of the theorem. If the weights satisfy

lims1sln(=1smaxu[1:s]minjuj=γu)=0,\lim_{s\to\infty}\frac{1}{s}\ln\left(\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}u\subset[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}\right)=0,

then we have

limε1+slnN(ε,IρsTMC)ε1+s\displaystyle\lim_{\varepsilon^{-1}+s\to\infty}\frac{\ln N(\varepsilon,I^{\mathrm{TMC}}_{\rho_{s}})}{\varepsilon^{-1}+s}
limε1+s(lnε2ε1+1sln(=1smaxu[1:s]minjuj=γu))=0,\displaystyle\leq\lim_{\varepsilon^{-1}+s\to\infty}\left(\frac{\ln\varepsilon^{-2}}{\varepsilon^{-1}}+\frac{1}{s}\ln\left(\sum_{\ell=1}^{s}\max_{\begin{subarray}{c}\emptyset\neq u\subseteq[1:s]\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}\right)\right)=0,

meaning that IρsTMCI^{\mathrm{TMC}}_{\rho_{s}} is a weakly tractable algorithm. Since the second and third assertions can be shown similarly, we omit the proof.∎

For instance, if the weights satisfy γuγv\gamma_{u}\geq\gamma_{v} whenever uvu\subset v, we always have

supuminjuj=γu=γ{}\sup_{\begin{subarray}{c}u\subset\mathbb{N}\\ \min_{j\in u}j=\ell\end{subarray}}\gamma_{u}=\gamma_{\{\ell\}}

for any >0\ell\in\mathbb{Z}_{>0}. Therefore, the necessary condition for the TMC estimator to be strongly polynomially tractable reduces to a simple summability:

=1γ{}<.\sum_{\ell=1}^{\infty}\gamma_{\{\ell\}}<\infty.

It is obvious to see that the necessary condition for the TMC estimator to be weakly tractable is stronger than that for the standard Monte Carlo estimator to be strongly tractable. Whether we can weaken the necessary conditions for the TMC estimator given in Theorem 2.2 or not is an open question.

3 Numerical experiments

In order to see the practical performance of the TMC estimator, we conduct four kinds of numerical experiments.222The C codes used in our experiments are available from https://github.com/takashigoda/Toeplitz-Monte-Carlo. The first test case follows Section 4.1 of Dick et al. (2015), which considers generating quadrature points from a multivariate normal distribution with a general covariance matrix. The second test case, taken from Section 4.2 of Dick et al. (2015), deals with approximating linear functionals of solutions of one-dimensional PDE with “uniform” random coefficients. The third test case is an extension of the second test case to a one-dimensional PDE with “log-normal” random coefficients. Finally, in the fourth test case we consider another possible extension of the second test case, namely a two-dimensional PDE with uniform random coefficients. All computations are performed on a laptop with 1.6 GHz Intel Core i5 CPU and 8 GB memory.

For every test case, we carry out numerical experiments with various values of NN and ss using both the standard Monte Carlo estimator and the TMC estimator. For each pair of NN and ss, we repeat computations R=25R=25 times independently and calculate the average computational time. For the latter three test cases, the variances of these estimators are measured by

1R(R1)r=1R(Iρs,(r)(f;N)Iρs(f;N)¯)2\frac{1}{R(R-1)}\sum_{r=1}^{R}\left(I^{\bullet,(r)}_{\rho_{s}}(f;N)-\overline{I^{\bullet}_{\rho_{s}}(f;N)}\right)^{2}

with

Iρs(f;N)¯=1Rr=1RIρs,(r)(f;N),\overline{I^{\bullet}_{\rho_{s}}(f;N)}=\frac{1}{R}\sum_{r=1}^{R}I^{\bullet,(r)}_{\rho_{s}}(f;N),

for {MC,TMC}\bullet\in\{\mathrm{MC},\mathrm{TMC}\}, where Iρs,(r)(f;N)I^{\bullet,(r)}_{\rho_{s}}(f;N) denotes the rr-th realization of the estimator Iρs(f;N)I^{\bullet}_{\rho_{s}}(f;N).

3.1 Generating points from multivariate Gaussian

Table 1: Average times (in seconds) to generate normally distributed points with a random covariance matrix for various values of NN and ss using the standard Monte Carlo method and the TMC method.
NN s=128s=128 s=256s=256 s=512s=512 s=1024s=1024 s=2048s=2048
stdMC 1024 0.041 0.192 1.369 9.613
TMC 0.165 0.239 0.403 0.872
saving 0.250 0.800 3.399 11.021
stdMC 2048 0.083 0.367 2.077 20.154 217.162
TMC 0.322 0.465 0.923 1.794 3.599
saving 0.259 0.790 2.251 11.234 60.342
stdMC 4096 0.154 0.740 5.551 45.287 446.974
TMC 0.586 0.981 1.909 3.407 7.112
saving 0.263 0.754 2.908 13.291 62.851
stdMC 8192 0.288 1.669 10.332 87.988 830.648
TMC 1.169 1.868 3.110 6.204 14.561
saving 0.247 0.893 3.322 14.183 57.047
stdMC 16384 0.660 3.255 19.764 159.792 1820.426
TMC 2.434 3.750 7.088 14.073 28.082
saving 0.271 0.868 2.788 11.355 64.826
stdMC 32768 1.078 5.627 33.011 331.747 3648.755
TMC 5.486 9.098 15.164 25.950 57.171
saving 0.197 0.618 2.177 12.784 63.822

Generating quadrature points from the multivariate normal distribution 𝒩(𝝁,Σ)\mathcal{N}(\boldsymbol{\mu},\Sigma) with mean vector 𝝁s\boldsymbol{\mu}\in\mathbb{R}^{s} and covariance matrix Σs×s\Sigma\in\mathbb{R}^{s\times s} is ubiquitous in scientific computation. The standard procedure is as follows (Devroye 1986, Chapter XI.2): Let As×sA\in\mathbb{R}^{s\times s} be a matrix which satisfies AA=ΣA^{\top}A=\Sigma. For instance, the Cholesky decomposition gives such AA in an upper triangular form for any symmetric positive-definite matrix Σ\Sigma. Using this decomposition, we can generate a point 𝒚𝒩(𝝁,Σ)\boldsymbol{y}\sim\mathcal{N}(\boldsymbol{\mu},\Sigma) by first generating 𝒙=(x1,,xs)\boldsymbol{x}=(x_{1},\ldots,x_{s}) with xj𝒩(0,1)x_{j}\sim\mathcal{N}(0,1) and then transforming 𝒙\boldsymbol{x} by

𝒚=𝝁+𝒙A.\boldsymbol{y}=\boldsymbol{\mu}+\boldsymbol{x}A.

Even if the matrix AA does not have any further structure, a set of quadrature points can be generated in a fast way by following Algorithm 1.

For our experiments, we fix 𝝁=(0,,0)\boldsymbol{\mu}=(0,\ldots,0) and choose AA randomly such that AA is a random upper triangular matrix with positive diagonal entries. Table 1 shows the average computational times for various values of NN and ss. As the theory predicts, the computational time for the standard Monte Carlo (stdMC) estimator scales as Ns2Ns^{2}, whereas that for the TMC estimator it approximately scales as NslogsNs\log s. For low-dimensional cases up to s=256s=256, the stdMC estimator is faster to compute than the TMC estimator. However, as the dimension ss increases, the relative speed of the TMC estimator also increases. For the case s=2048s=2048, the TMC is approximately 60 times faster than the stdMC. This result indicates that the TMC estimator is useful in high-dimensional settings for generating normally distributed quadrature points for a general covariance matrix.

3.2 1D differential equation with uniform random coefficients

Let us consider the ODE

ddx(a(x,𝒚)ddxu(x,𝒚))=g(x)1\displaystyle-\frac{\mathrm{d}}{\mathrm{d}x}\left(a(x,\boldsymbol{y})\frac{\mathrm{d}}{\mathrm{d}x}u(x,\boldsymbol{y})\right)=g(x)\equiv 1
for x(0,1)x\in(0,1) and 𝒚[12,12]\boldsymbol{y}\in\left[-\frac{1}{2},\frac{1}{2}\right]^{\mathbb{N}}
u(x,𝒚)=0for x=0,1\displaystyle u(x,\boldsymbol{y})=0\quad\text{for $x=0,1$}
a(x,𝒚)=2+j=1yjsin(2πjx)j3/2.\displaystyle a(x,\boldsymbol{y})=2+\sum_{j=1}^{\infty}y_{j}\frac{\sin(2\pi jx)}{j^{3/2}}.

In order to solve this ODE approximately with the finite element method, we consider a system of hat functions

ϕm(x)={(xxm1)Mif xm1xxm,(xm+1x)Mif xmxxm+1,0otherwise,\phi_{m}(x)=\begin{cases}(x-x_{m-1})M&\text{if $x_{m-1}\leq x\leq x_{m}$,}\\ (x_{m+1}-x)M&\text{if $x_{m}\leq x\leq x_{m+1}$,}\\ 0&\text{otherwise,}\end{cases}

for m=1,,M1m=1,\ldots,M-1 over M+1M+1 equi-distributed nodes xm=m/Mx_{m}=m/M for m=0,1,,Mm=0,1,\ldots,M, and truncate the infinite sum appearing in the random field aa by the first ss terms. Therefore, we have three different parameters N,MN,M and ss.

Given the boundary condition on uu, the approximate solution uMu_{M} of the ODE for y1,,ysU[1/2,1/2]y_{1},\ldots,y_{s}\sim U[-1/2,1/2] is given by

uM=m=1M1u^mϕm(x),u_{M}=\sum_{m=1}^{M-1}\hat{u}_{m}\phi_{m}(x),

with

(u^1,,u^M1)B(y1,,ys)=(g^1,,g^M1),\left(\hat{u}_{1},\ldots,\hat{u}_{M-1}\right)\cdot B(y_{1},\ldots,y_{s})=(\hat{g}_{1},\ldots,\hat{g}_{M-1}), (6)

for the symmetric stiffness matrix BB depending on y1,,ysy_{1},\ldots,y_{s} and the forcing vector with entries g^m=01g(x)ϕm(x)dx=1/M\hat{g}_{m}=\int_{0}^{1}g(x)\phi_{m}(x)\,\mathrm{d}x=1/M. The entries of the matrix B=(bk,)k,(M1)×(M1)B=(b_{k,\ell})_{k,\ell}\in\mathbb{R}^{(M-1)\times(M-1)} are given by

bk,\displaystyle b_{k,\ell} =01a(x,(y1,,ys,0,0,))ϕk(x)ϕ(x)dx\displaystyle=\int_{0}^{1}a(x,(y_{1},\ldots,y_{s},0,0,\ldots)){\phi}^{\prime}_{k}(x){\phi}^{\prime}_{\ell}(x)\,\mathrm{d}x
=201ϕk(x)ϕ(x)dx\displaystyle=2\int_{0}^{1}{\phi}^{\prime}_{k}(x){\phi}^{\prime}_{\ell}(x)\,\mathrm{d}x
+j=1syj01sin(2πjx)j3/2ϕk(x)ϕ(x)dx\displaystyle\quad+\sum_{j=1}^{s}y_{j}\int_{0}^{1}\frac{\sin(2\pi jx)}{j^{3/2}}{\phi}^{\prime}_{k}(x){\phi}^{\prime}_{\ell}(x)\,\mathrm{d}x
=:ak,(0)+j=1syjak,(j).\displaystyle=:a_{k,\ell}^{(0)}+\sum_{j=1}^{s}y_{j}a_{k,\ell}^{(j)}.

Hence, by letting A(j)=(ak,(j))k,(M1)×(M1)A^{(j)}=(a_{k,\ell}^{(j)})_{k,\ell}\in\mathbb{R}^{(M-1)\times(M-1)} for j=0,1,,sj=0,1,\ldots,s, we have

B=A(0)+j=1syjA(j).B=A^{(0)}+\sum_{j=1}^{s}y_{j}A^{(j)}.

Here we note that every entry of A(j)A^{(j)} can be explicitly calculated as

ak,(0)={4Mif k=,2Mif |k|=1,0otherwise,a_{k,\ell}^{(0)}=\begin{cases}4M&\text{if $k=\ell$,}\\ -2M&\text{if $|k-\ell|=1$,}\\ 0&\text{otherwise,}\end{cases}

and

ak,(j)={M2πj5/2sin(2πjM)sin(2πjkM)if k=,M2πj5/2sin(πjM)sin(πj(k+)M)if |k|=1,0otherwise,a_{k,\ell}^{(j)}=\begin{cases}\frac{M^{2}}{\pi j^{5/2}}\sin\left(\frac{2\pi j}{M}\right)\sin\left(\frac{2\pi jk}{M}\right)&\text{if $k=\ell$,}\\ -\frac{M^{2}}{\pi j^{5/2}}\sin\left(\frac{\pi j}{M}\right)\sin\left(\frac{\pi j(k+\ell)}{M}\right)&\text{if $|k-\ell|=1$,}\\ 0&\text{otherwise,}\end{cases}

for 1js1\leq j\leq s. Since each matrix A(j)A^{(j)} is tridiagonal, the LU decomposition requires only O(M)O(M) computational time to solve the system of linear equations (6). This way, it is clear that computing the matrix BB for NN Monte Carlo samples on (y1,,ys)(y_{1},\ldots,y_{s}) is computationally dominant for the whole process, and as shown in Section 3.2 of Dick et al. (2015), the standard Monte Carlo method requires O(NMs)O(NMs) arithmetic operations, whereas Algorithm 1 can reduce them to O(NMlogs)O(NM\log s). It is expected that the TMC estimator brings substantial computational cost savings, particularly for large ss. For our experiments, we estimate the expectation of u(1/2,)u(1/2,\cdot).

Table 2 shows the results for various values of N,MN,M and ss. In general, in the area of PDEs with random coefficients, both MM and ss grow with NN so that the following three errors are balanced (Dick et al. 2015): the finite element discretization error, the truncation error on the random field aa, the Monte Carlo error. The optimal balancing depends on the decay rates of these errors. However, in our numerical experiments, we are only interested in how the computation times change for different relations between MM, ss, and NN, and so we test different relations between those parameters (irrespective of what the optimal choice actually is).

Since computations are repeated 25 times independently for each pair, here we report the average of the estimation values and the variance of the estimator for two methods. By comparing the mean values computed by two methods, we can confirm that the TMC estimator is also unbiased (just as the stdMC estimator is). The variances for both of the estimators decay with the rate O(1/N)O(1/N), whereas the magnitude for the TMC estimator is approximately 2–5 larger than the stdMC estimator. On the other hand, the computational time for the stdMC estimator increases with NN (equivalently, with ss) significantly faster than the TMC estimator. This increment behavior of the computation times indicates that computation of the stiffness matrix is the most computationally dominant part in this computation, and so the TMC estimator is quite effective in reducing the computation time.

Table 2: Estimating the expectation of u(1/2,)u(1/2,\cdot) with various values of N,MN,M and ss using the standard Monte Carlo method and the TMC method for the uniform case. The average estimate, the variance of the estimator and the average computational time (in seconds) are shown for each method. The efficiency is defined by the ratio of the product of the variance and the computational time between two methods.
stdMC TMC
mean variance time mean variance time efficiency
NN N=M=sN=M=s
64 0.066 2.00108\cdot 10^{-8} 0.002 0.066 4.39108\cdot 10^{-8} 0.023 0.046
128 0.065 4.20109\cdot 10^{-9} 0.020 0.065 1.47108\cdot 10^{-8} 0.071 0.081
256 0.064 2.00109\cdot 10^{-9} 0.189 0.064 7.11109\cdot 10^{-9} 0.221 0.240
512 0.064 1.01109\cdot 10^{-9} 1.455 0.064 2.27109\cdot 10^{-9} 0.823 0.781
1024 0.064 5.751010\cdot 10^{-10} 19.720 0.064 8.451010\cdot 10^{-10} 3.202 4.190
2048 0.064 1.961010\cdot 10^{-10} 217.303 0.064 5.511010\cdot 10^{-10} 12.069 6.405
4096 0.064 9.501011\cdot 10^{-11} 2367.836 0.064 2.361010\cdot 10^{-10} 57.402 16.612
NN N=M2=sN=M^{2}=s
256 0.076 3.34108\cdot 10^{-8} 0.005 0.076 1.17107\cdot 10^{-7} 0.013 0.125
512 0.072 7.88109\cdot 10^{-9} 0.042 0.072 2.44108\cdot 10^{-8} 0.031 0.441
1024 0.070 2.77109\cdot 10^{-9} 0.298 0.070 4.98109\cdot 10^{-9} 0.087 1.892
2048 0.068 4.781010\cdot 10^{-10} 1.871 0.068 2.07109\cdot 10^{-9} 0.235 1.838
4096 0.066 3.661010\cdot 10^{-10} 10.965 0.066 6.481010\cdot 10^{-10} 0.766 8.088
8192 0.066 6.701011\cdot 10^{-11} 75.863 0.066 2.281010\cdot 10^{-10} 2.390 9.329
16384 0.065 2.201011\cdot 10^{-11} 963.569 0.065 9.701011\cdot 10^{-11} 7.033 31.073
NN N=2M=2sN=2M=2s
64 0.069 3.06108\cdot 10^{-8} 0.001 0.070 9.98108\cdot 10^{-8} 0.017 0.013
128 0.066 6.16109\cdot 10^{-9} 0.004 0.066 2.36108\cdot 10^{-8} 0.044 0.026
256 0.065 2.15109\cdot 10^{-9} 0.041 0.065 9.24109\cdot 10^{-9} 0.135 0.070
512 0.064 7.611010\cdot 10^{-10} 0.326 0.064 2.58109\cdot 10^{-9} 0.449 0.214
1024 0.064 3.861010\cdot 10^{-10} 2.681 0.064 8.951010\cdot 10^{-10} 1.413 0.818
2048 0.064 1.721010\cdot 10^{-10} 35.731 0.064 5.671010\cdot 10^{-10} 5.648 1.919
4096 0.064 7.841011\cdot 10^{-11} 444.156 0.064 2.411010\cdot 10^{-10} 23.480 6.144
NN N=2M2=2sN=2M^{2}=2s
512 0.076 1.39108\cdot 10^{-8} 0.011 0.076 4.37108\cdot 10^{-8} 0.026 0.137
1024 0.072 3.25109\cdot 10^{-9} 0.064 0.072 9.13109\cdot 10^{-9} 0.062 0.364
2048 0.070 7.201010\cdot 10^{-10} 0.534 0.070 3.26109\cdot 10^{-9} 0.183 0.645
4096 0.068 2.231010\cdot 10^{-10} 3.204 0.068 9.551010\cdot 10^{-10} 0.524 1.427
8192 0.066 1.411010\cdot 10^{-10} 20.888 0.066 3.011010\cdot 10^{-10} 1.596 6.133
16384 0.066 3.601011\cdot 10^{-11} 159.880 0.066 1.221010\cdot 10^{-10} 4.618 10.216
32768 0.065 1.021011\cdot 10^{-11} 2023.179 0.065 5.801011\cdot 10^{-11} 13.586 26.144

As is standard (see for instance Chapter 8 of Owen (2019)), we measure the relative efficiency of the TMC estimator compared to the stdMC estimator by the ratio

TMCσMC2TTMCσTMC2,\frac{T_{\mathrm{MC}}\sigma^{2}_{\mathrm{MC}}}{T_{\mathrm{TMC}}\sigma^{2}_{\mathrm{TMC}}},

where TT_{\bullet} and σ2\sigma^{2}_{\bullet} denote the computational time spent and the variance, respectively, for the estimators {MC,TMC}\bullet\in\{\mathrm{MC},\mathrm{TMC}\}. As shown in the rightmost column of Table 2, the relative efficiency is smaller than 1 for low-dimensional cases, which means that we do not gain any benefit from using the TMC estimator. However, because of the substantial computational time savings, the efficiency increases significantly for large ss where it goes well beyond 1.

3.3 1D differential equation in the log-normal case

Let us move on to an ODE with the log-normal random coefficients:

ddx(a(x,𝒚)ddxu(x,𝒚))=g(x)1\displaystyle-\frac{\mathrm{d}}{\mathrm{d}x}\left(a(x,\boldsymbol{y})\frac{\mathrm{d}}{\mathrm{d}x}u(x,\boldsymbol{y})\right)=g(x)\equiv 1
for x(0,1)x\in(0,1) and yji.i.d. N(0,1)y_{j}\sim\text{i.i.d.\ }N(0,1)
u(x,𝒚)=0for x=0,1\displaystyle u(x,\boldsymbol{y})=0\quad\text{for $x=0,1$}
a(x,𝒚)=exp(j=1yjsin(2πjx)j2).\displaystyle a(x,\boldsymbol{y})=\exp\left(\sum_{j=1}^{\infty}y_{j}\frac{\sin(2\pi jx)}{j^{2}}\right).

Similarly to the uniform case, we truncate the infinite sum appearing in the random field aa by the first ss terms. A similar test case was also used in Section 4.3 of Dick et al. (2015).333We point out that in Section 4.3 of Dick et al. (2015) the authors replaced the normal distribution by a uniform distribution. However, a normal distribution could have been used in (Dick et al. 2015, Section 4.3) as well, for instance by shifting the QMC points in [0,1]s[0,1]^{s} in the quadrature rule by (1/(2N),,1/(2N))(1/(2N),\ldots,1/(2N)) and applying the inverse CDF to the shifted points. Doing so avoids the point (0,,0)[0,1]s(0,\ldots,0)\in[0,1]^{s} which would get transformed to (,,)(-\infty,\ldots,-\infty). Note that this shift does not effect the fast QMC matrix vector product, which can still be applied in the usual way.

Now, as the entries of the stiffness matrix B=(bk,)k,(M1)×(M1)B=(b_{k,\ell})_{k,\ell}\in\mathbb{R}^{(M-1)\times(M-1)} cannot be expressed simply as a linear sum of y1,y2,y_{1},y_{2},\ldots, we need to approximate the integral by using some quadrature formulas, except for the case |k|2|k-\ell|\geq 2 where we just have bk,=0b_{k,\ell}=0. Denoting the quadrature nodes and the corresponding weights by x1,k,,,xI,k,x_{1,k,\ell},\ldots,x_{I,k,\ell} and ω1,k,,,ωI,k,\omega_{1,k,\ell},\ldots,\omega_{I,k,\ell}, the entry bk,b_{k,\ell} is approximated by

bk,b^k,=i=1Iωi,k,exp(θi,k,)ϕk(xi,k,)ϕ(xi,k,),b_{k,\ell}\approx\hat{b}_{k,\ell}=\sum_{i=1}^{I}\omega_{i,k,\ell}\exp(\theta_{i,k,\ell}){\phi}^{\prime}_{k}(x_{i,k,\ell}){\phi}^{\prime}_{\ell}(x_{i,k,\ell}),

where

θi,k,=j=1syjsin(2πjxi,k,)j2,\theta_{i,k,\ell}=\sum_{j=1}^{s}y_{j}\frac{\sin(2\pi jx_{i,k,\ell})}{j^{2}},

for the case |k|1|k-\ell|\leq 1. As stated in Section 3.2 of Dick et al. (2015), computing θi,k,\theta_{i,k,\ell} for all 1iI1\leq i\leq I and all NN Monte Carlo samples on (y1,,ys)(y_{1},\ldots,y_{s}) can be done in a fast way by using the TMC estimator, requiring O(INMlogs)O(INM\log s) arithmetic operations. On the other hand, we need O(INMs)O(INMs) arithmetic operations when using the standard Monte Carlo estimator. This way, the log-normal case poses a further computational challenge compared to the uniform case, so that it would be interesting to see whether the TMC is still effective. In this paper we apply the 3-point closed Newton-Cotes formula with nodes at xk1,xk,xk+1x_{k-1},x_{k},x_{k+1} if k=k=\ell and the 2-point closed one with nodes at x(k+1)/2,x(k++1)/2x_{(k+\ell-1)/2},x_{(k+\ell+1)/2} if |k|=1|k-\ell|=1 . Again we estimate the expectation of u(1/2,)u(1/2,\cdot).

Table 3 shows the results for various values of N,MN,M and ss. Similarly to the uniform case, we see that the TMC estimator is unbiased as the mean values agree well with the results for the stdMC estimator. In this case, however, the variances for both of the estimators do not necessarily decay with the rate O(1/N)O(1/N). This is possible because we increase MM and ss simultaneously with NN, which may lead to an increment of the variance of uM(1/2,)u_{M}(1/2,\cdot) in a non-asymptotic range of NN. As NN increases further, it is expected that the variance of uM(1/2,)u_{M}(1/2,\cdot) stays almost the same and that the variances for both of the estimators tend to decay with the rate O(1/N)O(1/N). Moreover, it can be seen from the table that the magnitude of the variance for the TMC estimator is comparable to that of the stdMC estimator for many choices of N,M,sN,M,s. As expected, the computational time for the stdMC estimator increases with ss significantly faster than the TMC estimator, and it is clear that computation of the stiffness matrix takes most of the computational time, even for the log-normal case. The relative efficiency of the TMC estimator over the stdMC estimator gets larger as NN (or, equivalently ss) increases.

Table 3: Estimating the expectation of u(1/2,)u(1/2,\cdot) with various values of N,MN,M and ss using the standard Monte Carlo method and the TMC method for the log-normal case.
stdMC TMC
mean variance time mean variance time efficiency
NN N=M=sN=M=s
64 0.018 8.34108\cdot 10^{-8} 0.073 0.018 2.97107\cdot 10^{-7} 0.079 0.260
128 0.018 1.17108\cdot 10^{-8} 0.580 0.018 1.76107\cdot 10^{-7} 0.171 0.227
256 0.018 1.55108\cdot 10^{-8} 4.650 0.018 7.74108\cdot 10^{-8} 0.549 1.694
512 0.018 2.11108\cdot 10^{-8} 36.640 0.018 3.35108\cdot 10^{-8} 2.260 10.230
1024 0.018 2.65108\cdot 10^{-8} 310.415 0.018 6.63108\cdot 10^{-8} 10.385 11.956
2048 0.018 1.23108\cdot 10^{-8} 2391.495 0.018 3.06108\cdot 10^{-8} 41.504 23.262
4096 0.018 2.33108\cdot 10^{-8} 20739.823 0.018 2.03108\cdot 10^{-8} 182.282 131.012
NN N=M2=sN=M^{2}=s
256 0.017 8.08109\cdot 10^{-9} 0.301 0.017 1.24108\cdot 10^{-8} 0.032 6.069
512 0.018 1.44108\cdot 10^{-8} 1.680 0.018 9.75109\cdot 10^{-9} 0.093 26.813
1024 0.018 2.68108\cdot 10^{-8} 9.762 0.018 1.24108\cdot 10^{-8} 0.277 76.387
2048 0.018 8.59109\cdot 10^{-9} 56.480 0.018 5.37108\cdot 10^{-8} 0.615 14.695
4096 0.018 1.89108\cdot 10^{-8} 317.294 0.018 2.98108\cdot 10^{-8} 2.212 91.011
8192 0.018 6.15108\cdot 10^{-8} 1865.110 0.018 3.27108\cdot 10^{-8} 7.094 494.909
16384 0.018 1.91108\cdot 10^{-8} 9964.289 0.018 2.35108\cdot 10^{-8} 16.548 473.083
NN N=2M=2sN=2M=2s
64 0.018 1.16107\cdot 10^{-7} 0.021 0.018 1.83107\cdot 10^{-7} 0.046 0.288
128 0.018 9.58108\cdot 10^{-8} 0.163 0.018 1.46107\cdot 10^{-7} 0.126 0.844
256 0.018 5.09108\cdot 10^{-8} 1.279 0.018 7.30108\cdot 10^{-8} 0.394 2.264
512 0.018 2.40108\cdot 10^{-8} 9.997 0.018 3.33108\cdot 10^{-8} 1.089 6.616
1024 0.018 9.53108\cdot 10^{-8} 82.822 0.018 6.00108\cdot 10^{-8} 4.914 26.758
2048 0.018 1.18108\cdot 10^{-8} 644.228 0.018 2.97108\cdot 10^{-8} 19.265 13.243
4096 0.018 1.45108\cdot 10^{-8} 5127.910 0.018 1.72108\cdot 10^{-8} 73.837 58.506
NN N=2M2=2sN=2M^{2}=2s
512 0.017 1.81108\cdot 10^{-8} 0.597 0.017 5.08109\cdot 10^{-9} 0.060 35.776
1024 0.018 5.07109\cdot 10^{-9} 3.370 0.018 6.70109\cdot 10^{-9} 0.151 16.834
2048 0.018 9.32109\cdot 10^{-9} 22.059 0.018 7.58109\cdot 10^{-9} 0.393 68.996
4096 0.018 8.18109\cdot 10^{-9} 123.556 0.018 5.70109\cdot 10^{-9} 1.187 149.507
8192 0.018 4.32109\cdot 10^{-9} 646.786 0.018 4.16109\cdot 10^{-9} 4.738 141.992
16384 0.018 2.47109\cdot 10^{-9} 3546.950 0.018 5.11109\cdot 10^{-9} 11.673 146.925
32768 0.018 1.81109\cdot 10^{-9} 20018.280 0.018 3.97109\cdot 10^{-9} 33.323 274.032

3.4 2D differential equation with random coefficients

Following Section 4 of Dick et al. (2016), our last example considers the following two-dimensional ODE with the uniform random coefficients:

(a(𝒙,𝒚)u(𝒙,𝒚))=g(𝒙)100x1\displaystyle-\nabla\cdot\left(a(\boldsymbol{x},\boldsymbol{y})\nabla u(\boldsymbol{x},\boldsymbol{y})\right)=g(\boldsymbol{x})\equiv 100x_{1}
for 𝒙[0,1]2\boldsymbol{x}\in[0,1]^{2} and 𝒚[12,12]\boldsymbol{y}\in\left[-\frac{1}{2},\frac{1}{2}\right]^{\mathbb{N}}
u(𝒙,𝒚)=0for 𝒙([0,1]2)\displaystyle u(\boldsymbol{x},\boldsymbol{y})=0\quad\text{for $\boldsymbol{x}\in\partial\left([0,1]^{2}\right)$}
a(𝒙,𝒚)=1+j=1yjsin(πkj,1x1)sin(πkj,2x2)(kj,12+kj,22)2.\displaystyle a(\boldsymbol{x},\boldsymbol{y})=1+\sum_{j=1}^{\infty}y_{j}\frac{\sin(\pi k_{j,1}x_{1})\sin(\pi k_{j,2}x_{2})}{(k_{j,1}^{2}+k_{j,2}^{2})^{2}}.

Here the elements (kj,1,kj,2)j(k_{j,1},k_{j,2})_{j} are ordered in such a way that {(kj,1,kj,2)j}=×\{(k_{j,1},k_{j,2})\mid j\in\mathbb{N}\}=\mathbb{N}\times\mathbb{N} and

1(kj,12+kj,22)21(kj+1,12+kj+1,22)2for all j.\frac{1}{(k_{j,1}^{2}+k_{j,2}^{2})^{2}}\geq\frac{1}{(k_{j+1,1}^{2}+k_{j+1,2}^{2})^{2}}\quad\text{for all $j\in\mathbb{N}$.}

In cases where equality holds, the ordering is arbitrary. We solve this ODE by a finite element discretization. Given the boundary condition on uu, we exclude the basis functions along the boundary and use a set of local piecewise linear hat functions {ϕp,q:0<p,q<M}\{\phi_{p,q}:0<p,q<M\} as the system of basis functions. The basis function ϕp,q\phi_{p,q} has center at (p/M,q/M)(p/M,q/M) and the support is given by the following hexagon:

(pM,qM)\left(\frac{p}{M},\frac{q}{M}\right)(p+1M,qM)\left(\frac{p+1}{M},\frac{q}{M}\right)(p1M,qM)\left(\frac{p-1}{M},\frac{q}{M}\right)(pM,q+1M)\left(\frac{p}{M},\frac{q+1}{M}\right)(pM,q1M)\left(\frac{p}{M},\frac{q-1}{M}\right)(p+1M,q+1M)\left(\frac{p+1}{M},\frac{q+1}{M}\right)(p1M,q1M)\left(\frac{p-1}{M},\frac{q-1}{M}\right)

The approximate solution uMu_{M} of the ODE in this setting is given by

uM=p,q=1M1u^p,qϕp,q(x1,x2),u_{M}=\sum_{p,q=1}^{M-1}\hat{u}_{p,q}\phi_{p,q}(x_{1},x_{2}),

with

(u^1,1,,u^1,M1,,u^M1,1,,u^M1,M1)\displaystyle\left(\hat{u}_{1,1},\ldots,\hat{u}_{1,M-1},\ldots,\hat{u}_{M-1,1},\ldots,\hat{u}_{M-1,M-1}\right)
B(y1,,ys)\displaystyle\quad\cdot B(y_{1},\ldots,y_{s})
=(g^1,1,,g^1,M1,,g^M1,1,,g^M1,M1),\displaystyle=(\hat{g}_{1,1},\ldots,\hat{g}_{1,M-1},\ldots,\hat{g}_{M-1,1},\ldots,\hat{g}_{M-1,M-1}), (7)

for the stiffness matrix BB depending on y1,,ysy_{1},\ldots,y_{s} and the forcing vector with entries g^p,q=[0,1]2g(𝒙)ϕp,q(𝒙)d𝒙\hat{g}_{p,q}=\int_{[0,1]^{2}}g(\boldsymbol{x})\phi_{p,q}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}, which is explicitly computable. By truncating the infinite sum appearing in the random field aa to the first ss terms, the entries of the matrix B=(b(p,q),(p,q))p,q,p,qB=(b_{(p,q),(p^{\prime},q^{\prime})})_{p,q,p^{\prime},q^{\prime}} are given by

b(p,q),(p,q)\displaystyle b_{(p,q),(p^{\prime},q^{\prime})}
=[0,1]2a(𝒙,(y1,,ys,0,))ϕp,q(𝒙)ϕp,q(𝒙)d𝒙\displaystyle=\int_{[0,1]^{2}}a(\boldsymbol{x},(y_{1},\ldots,y_{s},0,\ldots))\nabla\phi_{p,q}(\boldsymbol{x})\nabla\phi_{p^{\prime},q^{\prime}}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}
=[0,1]2ϕp,q(𝒙)ϕp,q(𝒙)d𝒙\displaystyle=\int_{[0,1]^{2}}\nabla\phi_{p,q}(\boldsymbol{x})\nabla\phi_{p^{\prime},q^{\prime}}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}
+j=1syj[0,1]2sin(πkj,1x1)sin(πkj,2x2)(kj,12+kj,22)2\displaystyle\quad+\sum_{j=1}^{s}y_{j}\int_{[0,1]^{2}}\frac{\sin(\pi k_{j,1}x_{1})\sin(\pi k_{j,2}x_{2})}{(k_{j,1}^{2}+k_{j,2}^{2})^{2}}
×ϕp,q(𝒙)ϕp,q(𝒙)d𝒙\displaystyle\qquad\qquad\qquad\times\nabla\phi_{p,q}(\boldsymbol{x})\nabla\phi_{p^{\prime},q^{\prime}}(\boldsymbol{x})\,\mathrm{d}\boldsymbol{x}
=:a(p,q),(p,q)(0)+j=1syja(p,q),(p,q)(j).\displaystyle=:a_{(p,q),(p^{\prime},q^{\prime})}^{(0)}+\sum_{j=1}^{s}y_{j}a_{(p,q),(p^{\prime},q^{\prime})}^{(j)}.

Similarly to the one-dimensional case, every term a(p,q),(p,q)(j)a_{(p,q),(p^{\prime},q^{\prime})}^{(j)} is explicitly calculable. Hence, by letting A(j)=(a(p,q),(p,q)(j))p,q,p,qA^{(j)}=(a_{(p,q),(p^{\prime},q^{\prime})}^{(j)})_{p,q,p^{\prime},q^{\prime}} for j=0,1,,sj=0,1,\ldots,s, we have

B=A(0)+j=1syjA(j).B=A^{(0)}+\sum_{j=1}^{s}y_{j}A^{(j)}.

This time, unlike the one-dimensional uniform case, each matrix A(j)A^{(j)} is no longer tridiagonal but has a bandwidth of O(M)O(M). Instead of the LU decomposition, we use the BiCGSTAB method without preconditioner to solve the system of linear equations (7). We refer to Saad (2003) for detailed information on the BiCGSTAB method. Since this is an iterative method, the resulting solution uMu_{M} is not precise and the required computational time depends on the stopping criterion we use. Moreover, such an additional computational burden makes it a priori unclear whether TMC can significantly reduce the computational time as a whole. However, in all our tests, computing the matrix BB for NN Monte Carlo samples on (y1,,ys)(y_{1},\ldots,y_{s}) remained the computationally dominant part, and the relative efficiency of the TMC estimator did not strongly depend on the stopping criterion. For our experiments, we estimate the expectation of u((1/2,1/2),)u((1/2,1/2),\cdot).

Table 4 shows the results for various values of N,MN,M and ss. The stopping criterion of the BiCGSTAB method is set such that the relative 2-norm of the residual is less than 10510^{-5}. We see that the TMC estimator is unbiased, and that the variances for both of the estimators approximately decay with the rate O(1/N)O(1/N). Similarly to the one-dimensional log-normal case, the magnitude for the TMC estimator is comparable to that of the stdMC estimator. The computational time for the stdMC estimator increases with ss much faster than the TMC estimator, resulting in a substantial relative efficiency of the TMC estimator over the stdMC estimator for larger ss.

Table 4: Estimating the expectation of u((1/2,1/2),)u((1/2,1/2),\cdot) with various values of N,MN,M and ss using the standard Monte Carlo method and the TMC method for the two-dimensional uniform case.
stdMC TMC
mean variance time mean variance time efficiency
NN N=M2=sN=M^{2}=s
16 3.516 1.09103\cdot 10^{-3} 0.0002 3.514 7.17104\cdot 10^{-4} 0.0020 0.119
64 3.647 2.16104\cdot 10^{-4} 0.009 3.643 2.43104\cdot 10^{-4} 0.027 0.302
256 3.676 8.03105\cdot 10^{-5} 1.248 3.678 8.84105\cdot 10^{-5} 0.406 2.792
1024 3.685 1.89105\cdot 10^{-5} 127.366 3.686 1.25105\cdot 10^{-5} 8.369 22.925
4096 3.688 3.62106\cdot 10^{-6} 14773.594 3.688 3.74106\cdot 10^{-6} 228.490 62.562
NN N=2M2=2sN=2M^{2}=2s
32 3.517 3.97104\cdot 10^{-4} 0.0004 3.515 3.20104\cdot 10^{-4} 0.0035 0.132
128 3.646 6.81105\cdot 10^{-5} 0.019 3.646 1.26104\cdot 10^{-4} 0.057 0.174
512 3.676 3.41105\cdot 10^{-5} 2.773 3.677 3.35105\cdot 10^{-5} 0.769 3.670
2048 3.686 5.67106\cdot 10^{-6} 260.868 3.686 7.59106\cdot 10^{-6} 16.375 11.893
8192 3.688 1.55106\cdot 10^{-6} 47914.800 3.688 1.66106\cdot 10^{-6} 494.336 90.562
NN 2N=M2=2s2N=M^{2}=2s
8 3.517 1.83103\cdot 10^{-3} 0.0001 3.513 1.89103\cdot 10^{-3} 0.0021 0.046
32 3.646 4.06104\cdot 10^{-4} 0.003 3.639 3.19104\cdot 10^{-4} 0.020 0.194
128 3.679 7.52105\cdot 10^{-5} 0.291 3.679 1.26104\cdot 10^{-4} 0.771 0.771
512 3.685 2.61105\cdot 10^{-5} 21.702 3.686 3.35105\cdot 10^{-5} 4.135 4.087
2048 3.687 3.63106\cdot 10^{-6} 2385.625 3.688 7.59106\cdot 10^{-6} 105.021 10.862
NN 2N=M2=s2N=M^{2}=s
32 3.522 4.11104\cdot 10^{-4} 0.0005 3.515 3.21104\cdot 10^{-4} 0.0035 0.184
128 3.646 7.54105\cdot 10^{-5} 0.034 3.646 1.27104\cdot 10^{-4} 0.041 0.490
512 3.676 2.61105\cdot 10^{-5} 5.259 3.677 3.35105\cdot 10^{-5} 0.749 5.468
2048 3.685 3.63106\cdot 10^{-6} 742.863 3.686 7.59106\cdot 10^{-6} 16.304 21.790
8192 3.688 1.96106\cdot 10^{-6} 115052.392 3.688 1.66106\cdot 10^{-6} 510.611 266.587

4 Conclusion

Motivated by applications to partial differential equations with random coefficients, we introduced the Toeplitz Monte Carlo estimator in this paper. The theoretical analysis of the TMC estimator shows that it is unbiased and the variance converges with the canonical 1/N1/N rate. From the viewpoint of tractability in the weighted L2L_{2} space, the TMC estimator requires a stronger condition on the weights than the standard Monte Carlo estimator to achieve strong polynomial tractability. Through a series of numerical experiments for PDEs with random coefficients, we observed that the TMC estimator is quite effective in reducing necessary computational times and the relative efficiency over the standard Monte Carlo estimator is substantial, particularly for high-dimensional settings.

We leave the following topics open for future research.

  • Combination with variance reduction techniques: In our numerical experiments for the one-dimensional uniform case, the variance of the TMC estimator tends to be much larger than the standard Monte Carlo estimator. To address this issue, it would be reasonable to consider applying some variance reduction techniques to the TMC estimator such that the resulting algorithm still allows for a fast matrix-vector multiplication. In particular, it would be interesting to design a variance reduction technique which reduces the term

    =1min(s,N)1(N)u[1:s]Iρ|u|(fufu+).\sum_{\ell=1}^{\min(s,N)-1}(N-\ell)\sum_{\emptyset\neq u\subseteq[1:s-\ell]}I_{\rho_{|u|}}(f_{u}f_{u+\ell}).
  • Multilevel Toeplitz Monte Carlo (MLTMC): Recently, multilevel Monte Carlo (MLMC) methods from Giles (2008) have been studied intensively in the context of PDEs with random coefficients, see for instance Cliffe et al. (2011) and Teckentrup et al. (2013). By combining the TMC estimator with MLMC, the dependence of the total computational complexity not only on the truncation dimension ss but also on the discretization parameter MM can be possibly weakened.

  • Applications to different areas: Although this work has been originally motivated by PDEs with random coefficients, the TMC estimator itself is more general and can be applied in different contexts as well. Since generating points from multivariate normal distribution is quite common, for instance, in financial engineering, operations research and machine learning, one may apply the TMC estimator also to those areas.

Acknowledgements.
Josef Dick is partly supported by the Australian Research Council Discovery Project DP190101197. Takashi Goda is partly supported by JSPS KAKENHI Grant Number 20K03744. J.D. would like to thank T.G. for his hospitality during his visit at U. Tokyo.

References

  • Cliffe et al. (2011) Cliffe, K. A., Giles, M. B., Scheichl, R., Teckentrup, A. L.: Multilevel Monte Carlo methods and applications to elliptic PDEs with random coefficients. Comput. Vis. Sci. 14, 3–15 (2011)
  • Devroye (1986) Devroye, L.: Non-Uniform Random Variate Generation. Springer, New York (1986)
  • Dick et al. (2015) Dick, J., Kuo, F. Y., Le Gia, Q. T., Schwab, C.: Fast QMC matrix-vector multiplication. SIAM J. Sci. Comput. 37, A1436–A1450 (2015)
  • Dick et al. (2016) Dick, J., Kuo, F. Y., Le Gia, Q. T., Schwab, C.: Multilevel higher order QMC Petrov-Galerkin discretization for affine parametric operator equations. SIAM J. Numer. Anal. 54, 2541–2568 (2016)
  • Feischl et al. (2018) Feischl, M., Kuo, F. Y., Sloan, I. H.: Fast random field generation with HH-matrices. Numer. Math. 140, 639–676 (2018)
  • Frigo and Johnson (2005) Frigo, M., Johnson, S. G.: The design and implementation of FFTW3. Proc. IEEE 93, 216–231 (2005)
  • Giles (2008) Giles, M. B.: Multilevel Monte Carlo path simulation. Oper. Res. 56, 607–617 (2008)
  • Giles et al. (2008) Giles, M. B., Kuo, F. Y., Sloan, I. H., Waterhouse, B. J.: Quasi-Monte Carlo for finance applications. ANZIAM J. 50, C308–C323 (2008)
  • Hoeffding (1948) Hoeffding, W.: A class of statistics with asymptotically normal distribution. Ann. Math. Statist. 19, 293–325 (1948)
  • Kuo et al. (2010) Kuo, F. Y., Sloan, I. H., Wasilkowski, G., Woźniakowski, H.: On decompositions of multivariate functions. Math. Comp. 79, 953–966 (2010)
  • Novak and Woźniakowski (2010) Novak, E., Woźniakowski, H.: Tractability of Multivariate Problems. Volume II: Standard Information for Functionals. EMS Tracts in Mathematics, 12. European Mathematical Society, Zürich (2010)
  • Owen (2019) Owen, A. B.: Monte Carlo Theory, Methods and Examples. http://statweb.stanford.edu/~owen/mc/ (2019) [Last accessed 4 March 2020]
  • Saad (2003) Saad, Y.: Iterative Methods for Sparse Linear Systems. 2nd Ed., SIAM, Philadelphia (2003)
  • Sloan and Woźniakowski (1998) Sloan, I. H., Woźniakowski, H.: When are quasi-Monte Carlo algorithms efficient for high-dimensional integrals? J. Complexity 14, 1–33 (1998)
  • Sobol’ (1993) Sobol’, I. M.: Sensitivity estimates for nonlinear mathematical models. Math. Mod. Comp. Exp. 1, 407–414 (1993)
  • Teckentrup et al. (2013) Teckentrup, A. L., Scheichl, R., Giles, M. B., Ullmann, E.: Further analysis of multilevel Monte Carlo methods for elliptic PDEs with random coefficients. Numer. Math. 125, 569–600 (2013)