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

Analyticity and sparsity in uncertainty quantification
for PDEs with Gaussian random field inputs

Dinh Dũng Information Technology Institute, Vietnam National University, Hanoi
144 Xuan Thuy, Cau Giay, Hanoi, Vietnam
Email: [email protected]
Van Kien Nguyen Department of Mathematical Analysis, University of Transport and Communications
No.3 Cau Giay Street, Lang Thuong Ward, Dong Da District, Hanoi, Vietnam
Email: [email protected]
Christoph Schwab Seminar for Applied Mathematics, ETH Zürich, 8092 Zürich, Switzerland
Email: [email protected]
Jakob Zech Interdisziplinäres Zentrum für wissenschaftliches Rechnen,
Universität Heidelberg, 69120 Heidelberg, Germany
Email: [email protected]
(February 24, 2025)
Abstract

We establish sparsity and summability results for coefficient sequences of Wiener-Hermite polynomial chaos expansions of countably-parametric solutions of linear elliptic and parabolic divergence-form partial differential equations with Gaussian random field inputs.

The novel proof technique developed here is based on analytic continuation of parametric solutions into the complex domain. It differs from previous works that used bootstrap arguments and induction on the differentiation order of solution derivatives with respect to the parameters. The present holomorphy-based argument allows a unified, “differentiation-free” proof of sparsity (expressed in terms of p\ell^{p}-summability or weighted 2\ell^{2}-summability) of sequences of Wiener-Hermite coefficients in polynomial chaos expansions in various scales of function spaces. The analysis also implies corresponding analyticity and sparsity results for posterior densities in Bayesian inverse problems subject to Gaussian priors on uncertain inputs from function spaces.

Our results furthermore yield dimension-independent convergence rates of various constructive high-dimensional deterministic numerical approximation schemes such as single-level and multi-level versions of Hermite-Smolyak anisotropic sparse-grid interpolation and quadrature in both forward and inverse computational uncertainty quantification.

1 Introduction

Gaussian random fields (GRFs for short) play a fundamental role in the modelling of spatio-temporal phenomena subject to uncertainty. In several broad research areas, particularly, in spatial statistics, data assimilation, climate modelling and meteorology to name but a few, GRFs play a pivotal role in mathematical models of physical phenomena with distributed, uncertain input data. Accordingly, there is an extensive literature devoted to mathematical, statistical and computational aspects of GRFs. We mention only [87, 75, 3] and the references there for mathematical foundations, and [89, 56] and the references there for a statistical perspective on GRFs.

In recent years, the area of computational uncertainty quantification (UQ for short) has emerged at the interface of the fields of applied mathematics, numerical analysis, scientific computing, computational statistics and data assimilation. Here, a key topic is the mathematical and numerical analysis of partial differential equations (PDEs for short) with random field inputs, and in particular with GRF inputs. The mathematical analysis of PDEs with GRF inputs addresses questions of well-posedness, pathwise and LpL^{p}-integrability and regularity in scales of Sobolev and Besov spaces of random solution ensembles of such PDEs. The numerical analysis focuses on questions of efficient numerical simulation methods of GRF inputs (see, e.g., [89, 50, 58, 28, 29, 12, 14, 100] and the references there), and the numerical approximation of corresponding PDE solution ensembles, which arise for GRF inputs. This concerns in particular the efficient representation of such solution ensembles (see [72, 9, 8, 52, 43]), and the numerical quadrature of corresponding solution fields (see, e.g., [72, 83, 51, 59, 68, 67, 98, 31, 43] and the references there). Applications include for instance subsurface flow models (see, e.g., [50, 57]) but also other PDE models for media with uncertain properties (see, e.g., [76] for electromagnetics). The careful analysis of efficient computational sampling of solution families of PDEs subject to GRF inputs is also a key ingredient in numerical data assimilation, e.g., in Bayesian inverse problems (BIPs for short); we refer to the surveys [48, 47] and the references therein for a mathematical formulation of BIPs for PDEs subject to Gaussian prior measures and function space inputs.

In the past few years there have been considerable developments in the analysis and numerical simulation of PDEs with random field input subject to Gaussian measures (GMs for short). The method of choice in many applications for the numerical treatment of GMs is Monte-Carlo (MC for short) sampling. The (mean-square) convergence rate 1/21/2 in terms of the number of MC samples is assured under rather mild conditions (existence of MC samples, and of finite second moments). We refer to, e.g., [36, 30, 106, 70] and the references there for a discussion of MC methods in this context. Given the high cost of MC sampling, recent years have seen the advent of numerical techniques which afford higher convergence orders than 1/21/2, also on infinite-dimensional integration domains. Like MC, these techniques are not prone to the so-called curse of dimensionality. Among them are Hermite-Smolyak sparse-grid interpolation (also referred to as “stochastic collocation”), see e.g. [52, 43, 45], and sparse-grid quadrature [31, 52, 63, 43, 45], and quasi-Monte Carlo (QMC for short) integration as developed in [59, 96, 83, 78, 68, 67] and the reference there.

The key condition which emerged as governing the convergence rates of numerical integration and interpolation methods for a function is a sparsity of the coefficients of its Wiener-Hermite polynomial chaos (PC for short) expansion, see, e.g., [72, 12]. Rather than counting the ratio of nonzero coefficients, the sparsity is quantified by p\ell^{p}-summability and/or weighted 2\ell^{2}-summability of these coefficients. This observation forms the foundation for the current text.

1.1 An example

To indicate some of the mathematical issues which are considered in this book, consider in the interval D=(0,1){D}=(0,1) and in a probability space (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}), a GRF g:Ω×Dg:\Omega\times{D}\to\mathbb{R} which takes values in L(D)L^{\infty}({D}). That is to say, that the map ωg(ω,)\omega\mapsto g(\omega,\cdot) is an element of the Banach space L(D)L^{\infty}({D}). Formally, at this stage, we represent realizations of the random element gL(D)g\in L^{\infty}({D}) with a representation system (ψj)j=1JL(D)(\psi_{j})_{j=1}^{J}\subset L^{\infty}({D}) in affine-parametric form

g(ω,x)=j=1Jyj(ω)ψj(x);,g(\omega,x)=\sum_{j=1}^{J}y_{j}(\omega)\psi_{j}(x)\,;, (1.1)

where the coefficients (yj)j=1J(y_{j})_{j=1}^{J} are assumed to be i.i.d. standard normal random variables (RVs for short) and JJ may be a finite number or infinity. Representations such as (1.1) are widely used both in the analysis and in the numerical simulation of random elements gg taking values in a function space. The coefficients yj(ω)y_{j}(\omega) being standard normal RVs, the sum j=1Jyjψj(x)\sum_{j=1}^{J}y_{j}\psi_{j}(x) may be considered as a parametric deterministic map g:JL(D)g:\mathbb{R}^{J}\to L^{\infty}({D}). The random element g(ω,x)g(\omega,x) in (1.1) can then be obtained by evaluating this deterministic map in random coordinates, i.e., by sampling it in Gaussian random vectors (yj(ω))j=1JJ(y_{j}(\omega))_{j=1}^{J}\in\mathbb{R}^{J}.

Gaussian random elements as inputs for PDEs appear in particular, in coefficients of diffusion equations. Consider, for illustration, in D{D}, and for given fL2(D)f\in L^{2}({D}), the boundary value problem: find a random function u:ΩVu:\Omega\to V with V:={wH1(D):w(0)=0}V:=\{w\in H^{1}({D}):w(0)=0\} such that

f(x)+ddx(a(x,ω)ddxu(x,ω))=0inD,a(1,ω)u(1,ω)=f¯.f(x)+\frac{\,\mathrm{d}}{\,\mathrm{d}x}\left(a(x,\omega)\frac{\,\mathrm{d}}{\,\mathrm{d}x}u(x,\omega)\right)=0\quad\mbox{in}\quad{D}\;,\quad a(1,\omega)u^{\prime}(1,\omega)=\bar{f}\;. (1.2)

Here, a(x,ω)=exp(g(x,ω))a(x,\omega)=\exp(g(x,\omega)) with GRF g:ΩL(D)g:\Omega\to L^{\infty}({D}), and f¯:=F(1)\bar{f}:=F(1) with

F(x):=0xf(ξ)dξV,xD.F(x):=\int_{0}^{x}f(\xi)\,\mathrm{d}\xi\in V,\quad x\in{D}.

In order to dispense with summability and measurability issues, let us temporarily assume that the sum in (1.1) is finite, with JJ\in\mathbb{N} terms. We find that a random solution uu of the problem must satisfy

u(x,ω)=exp(g(x,ω))F(x),xD,ωΩ.u^{\prime}(x,\omega)=-\exp(-g(x,\omega))F(x),\quad x\in{D},\omega\in\Omega\;.

Inserting (1.1), this is equivalent to the parametric, deterministic family of solutions u(x,𝒚):D×Ju(x,{\boldsymbol{y}}):{D}\times\mathbb{R}^{J}\to\mathbb{R} given by

u(x,𝒚)=exp(g(x,𝒚))F(x),xD,𝒚J.u^{\prime}(x,{\boldsymbol{y}})=-\exp(-g(x,{\boldsymbol{y}}))F(x),\quad x\in{D},{\boldsymbol{y}}\in\mathbb{R}^{J}\;. (1.3)

Hence

u(,𝒚)L2(D)=exp(g(,𝒚))FL2(D),𝒚J,\|u^{\prime}(\cdot,{\boldsymbol{y}})\|_{L^{2}({D})}=\|\exp(-g(\cdot,{\boldsymbol{y}}))F\|_{L^{2}({D})}\;,\quad{\boldsymbol{y}}\in\mathbb{R}^{J}\;,

which implies the (sharp) bounds

u(,𝒚)L2(D){exp(g(,𝒚)L(D))FL2(D)exp(g(,𝒚)L(D))FL2(D).\|u^{\prime}(\cdot,{\boldsymbol{y}})\|_{L^{2}({D})}\left\{\begin{array}[]{l}\geq\exp(-\|g(\cdot,{\boldsymbol{y}})\|_{L^{\infty}({D})})\|F\|_{L^{2}({D})}\\ \leq\exp(\|g(\cdot,{\boldsymbol{y}})\|_{L^{\infty}({D})})\|F\|_{L^{2}({D})}.\end{array}\right.

Due to the homogeneous Dirichlet condition at x=0x=0, up to an absolute constant the same bounds also hold for u(,𝒚)V\|u(\cdot,{\boldsymbol{y}})\|_{V}.

It is evident from the explicit expression (1.3) and the upper and lower bounds, that for every parameter 𝒚J{\boldsymbol{y}}\in\mathbb{R}^{J}, the solution uVu\in V exists. However, we can not, in general, expect uniform w.r.t. 𝒚J{\boldsymbol{y}}\in\mathbb{R}^{J} a-priori estimates, also of the higher derivatives, for smoother functions xg(x,𝒚)x\mapsto g(x,{\boldsymbol{y}}) and xf(x)x\mapsto f(x). Therefore, the parametric problem (1.2) is nonuniformly elliptic, [28, 70]. In particular, also a-priori error bounds for various discretization schemes will contain this uniformity w.r.t. 𝒚{\boldsymbol{y}}. The random solution will be recovered from (1.3) by inserting for the coordinates yjy_{j} samples of i.i.d. standard normal RVs.

This book focuses on developing a regularity theory for countably-parametric solution families u(,𝒚):𝒚J{u(\cdot,{\boldsymbol{y}}):{\boldsymbol{y}}\in\mathbb{R}^{J}} with a particular emphasis on the case J=J=\infty. This allows for arbitrary Gaussian random fields g(,ω)g(\cdot,\omega) in (1.2). Naturally, our results also cover the finite-parametric setting where the number JJ of random parameters is finite, but may be very large. Then, all constants in our error estimates are either independent of the parameter dimension JJ or their dependence of JJ is explicitly indicated. Previous works [8, 9, 59] addressed the p\ell^{p}-summability of the Wiener-Hermite PC expansion coefficients of solution families {u(,𝒚):𝒚}V\{u(\cdot,{\boldsymbol{y}}):{\boldsymbol{y}}\in\mathbb{R}^{\infty}\}\subset V for the forward problem, based on moment bounds of derivatives of parametric solutions w.r.t. GM. Estimates for these coefficients and, in particular, for the summability, were obtained in [72, 8, 9, 59, 71]. In these references, all arguments were based on real-variable, bootstrapping arguments with respect to 𝒚{\boldsymbol{y}}.

1.2 Contributions

We make the following contributions to the area computational UQ for PDEs with GRF inputs. First, we provide novel proofs of some of the sparsity results in [72, 9, 8] of the infinite-dimensional parametric forward solution map to PDEs with GRF inputs. The presently developed proof technique is based on holomorphic continuation and complex variable arguments in order to bound derivatives of parametric solutions, and their coefficients in Wiener-Hermite PC expansions. This is in line with similar arguments in the so-called “uniform case” in [39, 32]. There, the random parameters in the representation of the input random fields range in compact subsets of {\mathbb{R}}. Unlike in these references, in the present text due to the Gaussian setup the parameter domain {\mathbb{R}}^{\infty} is not compact. This entails significant modifications of mathematical arguments as compared to those in [39, 32].

Contrary to the analysis in [8, 9, 59], where parametric regularity results were obtained by real-variable arguments combined with induction-based bootstrapping with respect to the derivative order, the present text develops derivative-free, complex variable arguments which allow directly to obtain bounds of the Wiener-Hermite PC expansion coefficients of the parametric solutions in scales of Sobolev and Besov spaces in the physical domain D{D} in which the parametric PDE is posed. They also allow to treat in a unified manner parametric regularity of the solution map in several scales of Sobolev and Kondrat’ev spaces in the physical domain D{D} which is the topic of Section 3.8, resulting in novel sparsity results for the solution operators to linear elliptic and parabolic PDEs with GRF inputs in scales of Sobolev and Besov spaces. We apply the quantified holomorphy of parametric solution families to PDEs with GRF inputs and preservation of holomorphy under composition, to problems of Bayesian PDE inversion conditional on noise observation data in Section 5, establishing in particular quantified parametric holomorphy of the corresponding Bayesian posterior.

We construct deterministic sparse-grid interpolation and quadrature methods for the parametric solution with convergence rate bounds that are free from the curse of dimensionality, and that afford possibly high convergence rates, given a sufficient sparsity in the Wiener-Hermite PC expansion of the parametric solutions. For sampling strategies in deterministic numerical quadrature, our findings show improved convergence rates, as compared to previous results in this area. Additionally, our novel sparsity results provided in scales of function spaces of varying spatial regularity enable us to construct apriori multilevel versions of sparse-grid interpolation and quadrature, with corresponding approximation rate bounds which are free from the curse of dimensionality, and explicit in terms of the overall number of degrees of freedom. Lastly, and in contrast to previous works, leveraging the preservation of holomorphy under compositions with holomorphic maps, our holomorphy-based arguments enable us to establish that our algorithms and bounds are applicable to posterior distributions in Bayesian inference problems involving GRF or Besov priors, as developed in [48, 94] and the references there.

1.3 Scope of results

We prove quantified holomorphy of countably-parametric solution families of linear elliptic and parabolic PDEs. The parameter range equals {\mathbb{R}}^{\infty}, corresponding to countably-parametric representations of GRF input data, taking values in a separable locally convex space, in particular, Hilbert or Banach space of uncertain input data, endowed for example with a Gaussian product measure γ\gamma on {\mathbb{R}}^{\infty}.

The results established in this text and the related bounds on partial derivatives w.r.t. the parameters in Karhunen-Loève or Lévy-Cieselsky expansions of uncertain GRF inputs imply convergence rate bounds for several families of computational methods to numerically access these parametric solution maps. Importantly, we prove that in terms of n1n\geq 1, an integer measure of work and memory, an approximation accuracy O(na)O(n^{-a}) for some parameter a>0a>0 can be achieved, where the convergence rate aa depends on the approximation process and on the amount of sparsity in the Wiener-Hermite PC expansion coefficients of the GRF under consideration. In the terminology of computational complexity, a prescribed numerical tolerance ε>0\varepsilon>0 can be reached in work and memory of order O(ε1/a)O(\varepsilon^{-1/a}). In particular, the convergence rate aa and the constant hidden in the Landau O()O(\cdot) symbol do not depend on the dimension of the space of active parameters involved in the approximations which we construct. The approximations developed in the present text are constructive and linear and can be realized computationally by deterministic algorithms of so-called “stochastic collocation” or “sparse-grid” type. Error bounds are proved in L2L^{2}-type Bochner spaces with respect to the GM γ\gamma on the input data space of the PDE, in natural Hilbert or Banach spaces of solutions of the PDEs under consideration. Here, it is important to notice that the sparsity of the Wiener-Hermite PC expansion coefficients used in constructive linear approximation algorithms and in estimating convergence rates, takes the form of weighted 2\ell^{2}-summability, but not p\ell^{p}-summability as in best nn-term approximations [72, 9, 8]. Furthermore, p\ell^{p}-summability results are implied from the corresponding weighted 2\ell^{2}-summability ones.

All approximation rates for deterministic sampling strategies in the present text are free from the so-called curse of dimensionality, a terminology coined apparently by R.E. Bellmann (see [17]). The rates are in fact only limited by the sparsity of the Wiener-Hermite PC expansion coefficients of the deterministic, countably-parametric solution families. In particular, dimension-independent convergence rates >1/2>1/2 are possible, provided a sufficient Wiener-Hermite PC expansion coefficient sparsity, that the random inputs feature sufficient pathwise regularity, and the affine representation system (being a tight frame on space of admissible input realizations) are stable in a suitable smoothness scale of inputs.

1.4 Structure and content of this text

We briefly describe the structure and content of the present text.

In Section 2, we collect known facts from functional analysis and GM theory which are required throughout this text. In particular, we review constructions and results on GMs on separable Hilbert and Banach spaces. Special focus will be on constructions via countable products of univariate GMs on countable products of real lines. We also review assorted known results on convergence rates of Lagrangian finite elements for linear, second order, divergence-form elliptic PDEs in polytopal domains D{D} with Lipschitz boundary D\partial{D}.

In Section 3, we address the analyticity and sparsity for elliptic divergence-form PDEs with log-Gaussian coefficients. In Section 3.1, we introduce a model linear, second order elliptic divergence-form PDE with log-Gaussian coefficients, with variational solutions in the “energy space” H01(D)H^{1}_{0}({D}). This equation was investigated with parametric input data in a number of references in recent years [38, 39, 72, 32, 9, 8, 42, 59, 83, 112]. It is considered in this work mainly to develop the holomorphic approach to establish our mathematical approach to parametric holomorphy and sparsity of Wiener-Hermite PC expansions of parametric solutions in a simple setting, and to facilitate comparisons with the mentioned previous works and results. We review known results on its well-posedness in Section 3.1, and Lipschitz continuous dependence on the input data in Section 3.2. We discuss regularity results for parametric coefficients in Section 3.3. Sections 3.4 and 3.5 describe uncertainty modelling by placing GMs on sets of admissible, countably parametric input data, i.e., formalizing mathematically aleatoric uncertainty in input data. Here, the Gaussian series introduced in Section 2.5 will be seen to take a key role in converting operator equations with GRF inputs to infinitely-parametric, deterministic operator equations. The Lipschitz continuous dependence of the solutions on input data from function spaces will imply strong measurability of corresponding random solutions, and render well-defined the uncertainty propagation, i.e., the push-forward of the GM on the input data. In Section 3.6, we connect the quantified holomorphy of the parametric, deterministic solution manifold {u(𝒚):𝒚}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in{\mathbb{R}}^{\infty}\} in the space H01(D)H^{1}_{0}({D}) with a sparsity (weighted 2\ell^{2}-summability and p\ell^{p}-summability) of the coefficients (u𝝂)𝝂(u_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}} of the (H01(D)H^{1}_{0}({D})-valued) Wiener-Hermite PC expansion. With this methodology in place, we show in Section 3.7 how to obtain holomorphic regularity of the parametric solution family {u(𝒚):𝒚}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in{\mathbb{R}}^{\infty}\} in Sobolev spaces Hs(D)H^{s}({D}) of possibly high smoothness order ss\in{\mathbb{N}} and how to derive from here the corresponding sparsity. The argument is self-contained and provides parametric holomorphy for any differentiation order ss\in{\mathbb{N}} in a unified way, in domains D{D} of sufficiently high regularity and for sufficiently high almost sure regularity of coefficient functions. In Section 3.8, we extend these results for linear second order elliptic differential operators in divergence form in a bounded polygonal domain D2{D}\subset\mathbb{R}^{2}. Here, corners are well-known to obstruct high almost sure pathwise regularity in the usual Sobolev and Besov spaces in D{D} for both, PC coefficients and parametric solutions. Therefore, we develop summability of the Wiener-Hermite PC expansion coefficients (u𝝂)𝝂(u_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}} of the random solutions in terms of corner-weighted Sobolev spaces, originating with V.A. Kondrat’ev (see, e.g., [61, 24, 90] and the references there). In Section 3.9, we briefly recall some known related results [32, 37, 38, 39, 72, 10, 11, 9, 8] on p\ell^{p}-summability and weighted 2\ell^{2}-summability of the generalized PC expansion coefficients of solutions to parametric divergence-form elliptic PDEs, as well as applications to best nn-term approximation.

In Section 4, we investigate sparsity of the Wiener-Hermite PC expansions coefficients of holomorphic functions. In Section 4.1, we introduce a concept of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy of parametric deterministic functions on the parameter domain \mathbb{R}^{\infty} taking values in a separable Hilbert space XX. This concept is fairly broad and covers a large range of parametric PDEs depending on log-Gaussian distributed data. In order to extend the results and the approach to bound Wiener-Hermite PC expansion coefficients via quantified holomorphy beyond the simple, second order diffusion equation introduced in Section 3, we address sparsity of the Wiener-Hermite PC expansions coefficients of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. In Section 4.2, we show that composite functions of a certain type are (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic under certain conditions. The significance of such functions is that they cover solution operators of a collection of linear, elliptic divergence-form PDEs in a unified way along with structurally similar PDEs with log-Gaussian random input data. This will allow to apply the ensuing results on convergence rates of deterministic collocation and quadrature algorithms to a wide range of PDEs with GRF inputs and functionals on their random solutions. In Section 4.3, we analyze some examples of holomorphic functions which are solutions to certain PDEs, including linear elliptic divergence-form PDEs with parametric diffusion coefficient, linear parabolic PDEs with parametric coefficient, linear elastostatics equations with log-Gaussian modulus of elasticity, Maxwell equations with log-Gaussian permittivity.

In Section 5, we apply the preceding abstract results on parametric holomorphy to establish quantified holomorphy of countably-parametric, posterior densities of corresponding BIPs where the uncertain input of the forward PDE is a countably-parametric GRF taking values in a separable Banach space of inputs. As an example, we analyze the BIP for the parametric diffusion coefficient of the diffusion equation with parametric log-Gaussian inputs.

In Section 6, we discuss deterministic interpolation and quadrature algorithms for approximation and numerical integration of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. Such algorithms are necessary for the approximation of certain statistical quantities (expectations, statistical moments) of the parametric solutions with respect to a GM on the parameter space. The proposed algorithms are variants and generalizations of so-called “stochastic collocation” or “sparse-grid” type approximation, and proved to outperform sampling methods such as MC methods, under suitable sparsity conditions on coefficients of the Wiener-Hermite PC expansion of integrands. In the quadrature case, they are also known as “Smolyak quadrature” methods. Their common feature is a) the deterministic nature of the algorithms, and b) the possibility of achieving convergence rates >1/2>1/2 independent of the dimension of parameters and therefore the curse of dimensionality is broken. They offer, in particular, the perspective of deterministic numerical approximations for GRFs under nonlinear pushforwards (being realized via the deterministic data-to-solution map of the PDE of interest). The decisive analytic property to be established are dimension-explicit estimates of individual Wiener-Hermite PC expansion coefficients of parametric solutions, and based on these, sharp summability estimates of norms of the coefficients of Wiener-Hermite PC expansion of parametric, deterministic solution families are given. In Sections 6.1 and 6.2, we construct sparse-grid Smolyak-type interpolation and quadrature algorithms. In Sections 6.3 and 6.4, we prove the convergence rates of interpolation and quadrature algorithms for (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions.

Section 7 is devoted to multilevel interpolation and quadrature of parametric holomorphic functions. We construct deterministic interpolation and quadrature algorithms for generic (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. For linear second order elliptic divergence-form PDEs with log-Gaussian coefficients, the results on the weighted 2\ell^{2}-summability of the Wiener-Hermite PC expansion coefficients of parametric, deterministic solution families with respect to corner-weighted Sobolev spaces on spatial domain D{D} finally also allow to analyze methods for constructive, deterministic linear approximations of parametric solution families. Here, a truncation of Wiener-Hermite PC expansions is combined with approximating the Wiener-Hermite PC expansion coefficients in the norm of the “energy space” H01(D)H^{1}_{0}({D}) of these solutions from finite-dimensional approximation spaces which are customary in the numerical approximation of solution instances. Importantly, required approximation accuracies of the Wiener-Hermite PC expansion coefficients u𝛎u_{\boldsymbol{\nu}} will depend on the relative importance of u𝛎u_{\boldsymbol{\nu}} within the Wiener-Hermite PC expansion. This observation gives rise to multilevel approximations where a prescribed overall accuracy in mean square w.r.t. the GM γ\gamma with respect to H01(D)H^{1}_{0}({D}) will be achieved by a 𝝂{\boldsymbol{\nu}}-dependent discretization level in the physical domain. Multilevel approximation and integration and the corresponding error estimates will be developed in this section in an abstract setting: Besides (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy, it is neccessary to require an assumption on the discretization error in the physical domain in the form of stronger holomorphy of the approximation error in this discretization. A combined assumption for guaranteeing constructive multilevel approximations is formulated in Section 7.1. In Section 7.2 we introduce multilevel algorithms for interpolation and quadrature of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions, and discuss work models and choices of discretization levels. A key for the sparse-grid integration and interpolation approaches is to efficiently numerically allocate discretization levels to Wiener-Hermite PC expansion coefficients. We develop such an approach in Section 7.3. It is based on greedy searches and suitable thresholding of (suitable norms of) Wiener-Hermite PC expansion coefficients and on a-priori bounds for these quantities which are obtained by complex variable arguments. In Sections 7.4 and 7.5, we establish convergence rate bounds of multilevel interpolation and quadrature algorithms for (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. In Section 7.6, we verify the abstract hypotheses of the sparse-grid multilevel approximations for the forward and inverse problems for concrete linear elliptic and parabolic PDEs on corner-weighted Sobolev spaces (Kondrat’ev spaces) with log-GRF inputs. In Section 7.7, we briefly recall some results from [43] (see also, [45] for some corrections) on linear multilevel (fully discrete) interpolation and quadrature in abstract Bochner spaces based on weighted 2\ell^{2}-summabilities. These results are subsequently applied to parametric divergence-form elliptic PDEs and to parametric holomorphic functions.

1.5 Notation and conventions

Additional to the real numbers \mathbb{R}, the complex numbers \mathbb{C}, and the positive integers \mathbb{N}, we set +:={x:x0}\mathbb{R}_{+}:=\{x\in\mathbb{R}\,:\,x\geq 0\} and 0:={0}\mathbb{N}_{0}:=\{0\}\cup\mathbb{N}. We denote by {\mathbb{R}}^{\infty} the set of all sequences 𝒚=(yj)j{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}} with yjy_{j}\in{\mathbb{R}}, and similarly define \mathbb{C}^{\infty}, +\mathbb{R}_{+}^{\infty} and 0\mathbb{N}_{0}^{\infty}. Both, {\mathbb{R}}^{\infty} and {\mathbb{C}}^{\infty}, will be understood with the product topology from {\mathbb{R}} and {\mathbb{C}}, respectively. For 𝜶{\boldsymbol{\alpha}}, 𝜷0d{\boldsymbol{\beta}}\in{\mathbb{N}}_{0}^{d}, d{}d\in\mathbb{N}\cup\{\infty\}, the inequality 𝜷𝜶{\boldsymbol{\beta}}\leq{\boldsymbol{\alpha}} is understood component-wise, i.e., 𝜷𝜶{\boldsymbol{\beta}}\leq{\boldsymbol{\alpha}} if and only if βjαj\beta_{j}\leq\alpha_{j} for all jj.

Denote by {\mathcal{F}} the countable set of all sequences of nonnegative integers 𝝂=(νj)j{\boldsymbol{\nu}}=(\nu_{j})_{j\in{\mathbb{N}}} such that supp(𝝂)\operatorname{supp}({\boldsymbol{\nu}}) is finite, where supp(𝝂):={j:νj0}\operatorname{supp}({\boldsymbol{\nu}}):=\{j\in{\mathbb{N}}:\nu_{j}\neq 0\} denotes the “support” of the multi-index 𝝂{\boldsymbol{\nu}}. Similarly, we define supp(𝝆)\operatorname{supp}({\boldsymbol{\rho}}) of a sequence 𝝆+{\boldsymbol{\rho}}\in{\mathbb{R}}^{\infty}_{+}. For 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}}, and for a sequence 𝒃=(bj)j{\boldsymbol{b}}=(b_{j})_{j\in{\mathbb{N}}} of positive real numbers, the quantities

𝝂!:=jνj!,|𝝂|:=jνj,and𝒃𝝂:=jbjνj{\boldsymbol{\nu}}!:=\prod_{j\in{\mathbb{N}}}\nu_{j}!\,,\qquad|{\boldsymbol{\nu}}|:=\sum_{j\in{\mathbb{N}}}\nu_{j},\qquad\text{and}\qquad{\boldsymbol{b}}^{\boldsymbol{\nu}}:=\prod_{j\in{\mathbb{N}}}b_{j}^{\nu_{j}}

are finite and well-defined.

For a multi-index 𝜶0d{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{d} and a function u(𝒙,𝒚)u({\boldsymbol{x}},{\boldsymbol{y}}) of 𝒙d{\boldsymbol{x}}\in{\mathbb{R}}^{d} and parameter sequence 𝒚{\boldsymbol{y}}\in{\mathbb{R}}^{\infty} we use the notation D𝜶u(𝒙,𝒚)D^{\boldsymbol{\alpha}}u({\boldsymbol{x}},{\boldsymbol{y}}) to indicate the partial derivatives taken with respect to 𝒙{\boldsymbol{x}}. The partial derivative of order 𝜶0{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{\infty} with respect to 𝒚{\boldsymbol{y}} of finite total order |𝜶|=jαj|{\boldsymbol{\alpha}}|=\sum_{j\in{\mathbb{N}}}\alpha_{j} is denoted by 𝜶u(𝒙,𝒚)\partial^{\boldsymbol{\alpha}}u({\boldsymbol{x}},{\boldsymbol{y}}). In order to simplify notation, we will systematically suppress the variable 𝒙Dd{\boldsymbol{x}}\in{D}\subset\mathbb{R}^{d} in mathematical expressions, except when necessary. For example, instead Dv(𝒙)d𝒙\int_{D}v({\boldsymbol{x}})\,\mathrm{d}{\boldsymbol{x}} we will write Dvd𝒙\int_{D}v\,\mathrm{d}{\boldsymbol{x}}, etc. For a Banach space XX, we denote by X:=X+iXX_{{\mathbb{C}}}:=X+{\rm i}X the complexification of XX. The space XX_{\mathbb{C}} is also a Banach space endowed with the (minimal, among several possible equivalent ones, see [91]) norm x1+ix2X:=sup0t2πx1costx2sintX\|x_{1}+{\rm i}x_{2}\|_{X_{\mathbb{C}}}:=\sup_{0\leq t\leq 2\pi}\|x_{1}\cos t-x_{2}\sin t\|_{X}. The space XX^{\infty} is defined in a similar way as {\mathbb{R}}^{\infty}.

By (X,Y)\mathcal{L}(X,Y) we denote the vector space of bounded, linear operators between to Banach spaces XX and YY. With is(X,Y)\mathcal{L}_{{\rm is}}(X,Y) we denote the subspace of boundedly invertible, linear operators from XX to YY.

For a function space X(D)X({{D}}) defined on the domain D{{D}}, if there is no ambiguity, when writing the norm of xX(D)x\in X({{D}}) we will omit D{{D}}, i.e., we write xX\|x\|_{X} instead of xX(D)\|x\|_{X({{D}})}.

For 0<p0<p\leq\infty and a finite or countable index set JJ, we denote by p(J)\ell^{p}(J) the quasi-normed space of all 𝒚=(yj)jJ{\boldsymbol{y}}=(y_{j})_{j\in J} with yjy_{j}\in\mathbb{R}, equipped with the quasi-norm 𝒚p(J):=(jJ|yj|p)1/p\|{\boldsymbol{y}}\|_{{\ell^{p}(J)}}:=\big{(}\sum_{j\in J}|y_{j}|^{p}\big{)}^{1/p} for p<p<\infty, and 𝒚(J):=supjJ|yj|\|{\boldsymbol{y}}\|_{{\ell^{\infty}(J)}}:=\sup_{j\in J}|y_{j}|. Sometimes, we make use of the abbreviation p=p(J)\ell^{p}=\ell^{p}(J) in a particular context if there is no misunderstanding of the meaning. We denote by (𝒆j)jJ({\boldsymbol{e}}_{j})_{j\in J} the standard basis of 2(J)\ell^{2}(J), i.e., 𝒆j=(ej,i)iJ{\boldsymbol{e}}_{j}=(e_{j,i})_{i\in J} with ej,i=1e_{j,i}=1 for i=ji=j and ej,i=0e_{j,i}=0 for iji\not=j.

2 Preliminaries

A key technical ingredient in the analysis of numerical approximations of PDEs with GRF inputs from function spaces, and of numerical methods for their efficient numerical treatment are constructions and numerical approximations of GRFs on real Hilbert and Banach spaces. Due to their high relevance in many areas of science (theoretical physics, quantum field theory, spatial and high-dimensional statistics, etc.), a rich theory has been developed in the past decades and a large body of literature is available now. We recapitulate basic definitions and key results, in particular on GMs, that are necessary for the ensuing developments. We do not attempt to provide a comprehensive survey. We require the exposition on GMs on real-valued Hilbert and Banach spaces, as most PDEs of interest are formulated for real-valued inputs and solutions. However, we crucially use in the ensuing sections of this text analytic continuation of parametric representations to the complex parameter domain. This is required in order to bring to bear complex variable methods for derivative-free, sharp bounds on Hermite expansion coefficients of GRFs. Therefore, we develop in our presentation solvability, well-posedness and regularity for the PDEs that are subject to GRF inputs in Hilbert and Banach spaces of complex-valued fields.

The structure of this section is as follows. In Section 2.1, we recapitulate GMs on finite dimensional spaces, in particular on d\mathbb{R}^{d} and d\mathbb{C}^{d}. In Section 2.2, we extend GMs to separable Banach spaces. Section 2.3 reviews the Cameron-Martin space. In Section 2.4 we recall a notion of Gaussian product measures on a Cartesian product of locally convex spaces. Section 2.5 is devoted to a summary of known representations of a GRF by a Gaussian series. A key object in these and more general spaces is the concept of Parseval frame which we introduce. For details, the reader may consult, for example, the books [3, 87, 21].

In Section 2.6 we recapitulate, from [6, 25, 54], (known) technical results on approximation properties of Lagrangian Finite Elements (FEs for short) in polygonal and polyhedral domains Dd{D}\subset\mathbb{R}^{d}, on regular, simplicial partitions of D{D} with local refinement towards corners (and, in space dimension d=3d=3, towards edges). These will be used in Section 6 in conjunction with collocation approximations in the parameter space of the GRF to build deterministic numerical approximations of solutions in polygonal and in polyhedral domains.

2.1 Finite dimensional Gaussian measures

2.1.1 Univariate Gaussian measures

In dimension d=1d=1, for every μ,σ\mu,\sigma\in\mathbb{R}, there holds the well-known identity

1σ2πexp((yμ)22σ2)dy=1.\frac{1}{\sigma\sqrt{2\pi}}\int_{\mathbb{R}}\exp\left(-\frac{(y-\mu)^{2}}{2\sigma^{2}}\right)\,\mathrm{d}y=1\;.

A Borel probability measure γ\gamma on \mathbb{R} is Gaussian if it is either a Dirac measure δμ\delta_{\mu} at μ\mu\in\mathbb{R} or its density with respect to Lebesgue measure λ\lambda on \mathbb{R} is given by

dγdλ=p(;μ,σ2),p(;μ,σ2):=y1σ2πexp((yμ)22σ2).\frac{\,\mathrm{d}\gamma}{\,\mathrm{d}\lambda}=p(\cdot;\mu,\sigma^{2})\;,\;\;p(\cdot;\mu,\sigma^{2}):=y\mapsto\frac{1}{\sigma\sqrt{2\pi}}\exp\left(-\frac{(y-\mu)^{2}}{2\sigma^{2}}\right)\;.

We shall refer to μ\mu as mean, and to σ2\sigma^{2} as variance of the GM γ\gamma. The case that γ=δμ\gamma=\delta_{\mu} is understood to correspond to σ=0\sigma=0. If σ>0\sigma>0, we shall say that the GM γ\gamma is nondegenerate. Unless explicitly stated otherwise, we assume GMs to be nondegenerate.

For μ=0\mu=0 and σ=1\sigma=1, we shall refer to the GM γ1\gamma_{1} as the standard GM on \mathbb{R}. A GM with μ=0\mu=0 is called centered (or also symmetric). For a GM γ\gamma on \mathbb{R}, there holds

μ=ydγ(y),σ2=(yμ)2dγ(y).\mu=\int_{\mathbb{R}}y\,\mathrm{d}\gamma(y),\quad\sigma^{2}=\int_{\mathbb{R}}(y-\mu)^{2}\,\mathrm{d}\gamma(y).

Let (Ω,𝒜,)(\Omega,{\mathcal{A}},\mathbb{P}) be a probability space with sample space Ω\Omega, σ\sigma-fields 𝒜{\mathcal{A}}, and probability measure \mathbb{P}. A Gaussian random variable (“Gaussian RV” for short) η:Ω\eta:\Omega\to\mathbb{R} is a RV whose law is Gaussian, i.e., it admits a Gaussian distribution. If η\eta is a Gaussian RV with mean μ\mu and variance σ2\sigma^{2} we write η𝒩(μ,σ2)\eta\sim\mathcal{N}(\mu,\sigma^{2}).

Linear transformations of Gaussian RVs are Gaussian: every Gaussian RV η\eta can be written as η=σξ+μ\eta=\sigma\xi+\mu, where ξ\xi is a standard Gaussian RV, i.e., a Gaussian RV whose law is a standard GM on \mathbb{R}.

The Fourier transformation of a GM γ\gamma on \mathbb{R} is defined, for every ξ\xi\in\mathbb{R}, as

γ1^(ξ):=exp(iξy)γ(y)=exp(iμξ12σ2ξ2).\hat{\gamma_{1}}(\xi):=\int_{\mathbb{R}}\exp({\rm i}\xi y)\gamma(y)=\exp\left({\rm i}\mu\xi-\frac{1}{2}\sigma^{2}\xi^{2}\right)\;.

We denote by Φ\Phi the distribution function of γ1\gamma_{1}. For the standard normal distribution

Φ(t)=tp(s;0,1)dst.\Phi(t)=\int_{-\infty}^{t}p(s;0,1)\,\mathrm{d}s\qquad\forall t\in\mathbb{R}.

With the convention Φ1(0):=\Phi^{-1}(0):=-\infty, Φ1(1):=+\Phi^{-1}(1):=+\infty, the inverse function Φ1\Phi^{-1} of Φ\Phi is defined on [0,1][0,1].

2.1.2 Multivariate Gaussian measures

Consider now a finite dimension d>1d>1. A Borel probability measure γ\gamma on (d,(d))(\mathbb{R}^{d},\mathcal{B}(\mathbb{R}^{d})) is called Gaussian if for every f(d,)f\in\mathcal{L}(\mathbb{R}^{d},\mathbb{R}) the measure γf1\gamma\circ f^{-1} is a GM on \mathbb{R}, where as usually, (d)\mathcal{B}(\mathbb{R}^{d}) denotes the σ\sigma-field on d\mathbb{R}^{d}. Since dd is finite, we may identify (d,)\mathcal{L}(\mathbb{R}^{d},\mathbb{R}) with d\mathbb{R}^{d}, and we denote the Euclidean inner product on d\mathbb{R}^{d} by (,)(\cdot,\cdot). The Fourier transform of a Borel measure γ\gamma on d\mathbb{R}^{d} is given by

γ^:d:γ^(𝝃)=dexp(i(𝝃,𝒚))dγ(𝒚).\hat{\gamma}:\mathbb{R}^{d}\to\mathbb{C}:\hat{\gamma}({\boldsymbol{\xi}})=\int_{\mathbb{R}^{d}}\exp\left({\rm i}({\boldsymbol{\xi}},{\boldsymbol{y}})\right)\,\mathrm{d}\gamma({\boldsymbol{y}})\;.

For a GM γ\gamma on d\mathbb{R}^{d}, the Fourier transform γ^\hat{\gamma} uniquely determines γ\gamma.

Proposition 2.1 ([21, Proposition 1.2.2]).

A Borel probability measure γ\gamma on d\mathbb{R}^{d} is Gaussian iff

γ^(𝝃)=exp(i(𝝃,𝝁)12(𝑲𝝃,𝝃)),𝝃d.\hat{\gamma}({\boldsymbol{\xi}})=\exp\left({\rm i}({\boldsymbol{\xi}},{\boldsymbol{\mu}})-\frac{1}{2}(\boldsymbol{K}{\boldsymbol{\xi}},{\boldsymbol{\xi}})\right),\quad{\boldsymbol{\xi}}\in\mathbb{R}^{d}\;.

Here, 𝛍d{\boldsymbol{\mu}}\in\mathbb{R}^{d} and 𝐊d×d\boldsymbol{K}\in\mathbb{R}^{d\times d} is a symmetric positive semidefinite matrix.

We shall say that a GM γ\gamma on d\mathbb{R}^{d} has a density with respect to Lebesgue measure λ\lambda on d\mathbb{R}^{d} iff the matrix 𝐊\boldsymbol{K} is nondegenerate. Then, this density is given by

dγdλ(𝒙):𝒙1(2π)ddet𝑲exp(12(𝑲1(𝒙𝝁),𝒙𝝁)).\frac{\,\mathrm{d}\gamma}{\,\mathrm{d}\lambda}({\boldsymbol{x}}):{\boldsymbol{x}}\mapsto\frac{1}{\sqrt{(2\pi)^{d}\det\boldsymbol{K}}}\exp\left(-\frac{1}{2}(\boldsymbol{K}^{-1}({\boldsymbol{x}}-{\boldsymbol{\mu}}),{\boldsymbol{x}}-{\boldsymbol{\mu}})\right)\;.

Furthermore,

𝝁=d𝒚dγ(𝒚),𝒚,𝒚d:(𝑲𝒚,𝒚)=d(𝒚,𝒙𝝁)(𝒚,𝒙𝝁)dγ(𝒙).{\boldsymbol{\mu}}=\int_{\mathbb{R}^{d}}{\boldsymbol{y}}\,\mathrm{d}\gamma({\boldsymbol{y}}),\quad\forall{\boldsymbol{y}},{\boldsymbol{y}}^{\prime}\in\mathbb{R}^{d}:(\boldsymbol{K}{\boldsymbol{y}},{\boldsymbol{y}}^{\prime})=\int_{\mathbb{R}^{d}}({\boldsymbol{y}},{\boldsymbol{x}}-{\boldsymbol{\mu}})({\boldsymbol{y}}^{\prime},{\boldsymbol{x}}-{\boldsymbol{\mu}})\,\mathrm{d}{\gamma}({\boldsymbol{x}})\;.

The symmetric linear operator 𝒞(d,d)\mathcal{C}\in\mathcal{L}(\mathbb{R}^{d},\mathbb{R}^{d}) defined by the later relation and represented by the symmetric positive definite matrix 𝐊\boldsymbol{K} is the covariance operator associated to the GM γ\gamma on d\mathbb{R}^{d}.

When we do not need to distinguish between the covariance operator 𝒞\mathcal{C} and the covariance matrix 𝑲\boldsymbol{K}, we simply speak of “the covariance” of a GM γ\gamma. If a joint probability distribution of RVs y1,,ydy_{1},\ldots,y_{d} is a GM on d\mathbb{R}^{d} with mean vector 𝝁{\boldsymbol{\mu}} and covariance matrix 𝑲\boldsymbol{K} we write (y1,,yd)𝒩(𝝁,𝑲)(y_{1},\ldots,y_{d})\sim\mathcal{N}({\boldsymbol{\mu}},\boldsymbol{K}).

In what follows, we use γd\gamma_{d} to denote the standard GM on d\mathbb{R}^{d}. Denote by L2(d;γd)L^{2}({\mathbb{R}}^{d};\gamma_{d}) the Hilbert space of all γd\gamma_{d}-measurable, real-valued functions ff on d\mathbb{R}^{d} such that the norm

fL2(d;γd):=(d|f(𝒚)|2dγd(𝒚))1/2\|f\|_{L^{2}({\mathbb{R}}^{d};\gamma_{d})}:=\left(\int_{\mathbb{R}^{d}}|f({\boldsymbol{y}})|^{2}\,\mathrm{d}\gamma_{d}({\boldsymbol{y}})\right)^{1/2}

is finite. The corresponding inner product is denoted by (,)L2(d;γd)(\cdot,\cdot)_{L^{2}({\mathbb{R}}^{d};\gamma_{d})}.

2.1.3 Hermite polynomials

A key role in the ensuing sparsity analysis of parametric solution families is taken by Wiener-Hermite PC expansions. We consider GRF inputs and, accordingly, will employ polynomial systems on {\mathbb{R}} which are orthogonal with respect to the GM γ1\gamma_{1} on {\mathbb{R}}, the so-called Hermite polynomials, as pioneered for the analysis of GRFs by N. Wiener in [109]. To this end, we recapitulate basic definitions and properties, in particular the various normalizations which are met in the literature. Particular attention will be paid to estimates for Hermite coefficients of functions which are holomorphic in a strip, going back to Einar Hille in [69].

Definition 2.2.

For k0k\in\mathbb{N}_{0}, the normalized probabilistic Hermite polynomial HkH_{k} of degree kk on \mathbb{R} is defined by

Hk(x):=(1)kk!exp(x22)dkdxkexp(x22).H_{k}(x):=\frac{(-1)^{k}}{\sqrt{k!}}\exp\left(\frac{x^{2}}{2}\right)\frac{\,\mathrm{d}^{k}}{\,\mathrm{d}x^{k}}\exp\left(-\frac{x^{2}}{2}\right). (2.1)

For every multi-degree 𝛎0m{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{m}, the mm-variate Hermite polynomial H𝛎H_{\boldsymbol{\nu}} is defined by

H𝝂(x1,,xm):=j=1mHνj(xj),xj,j=1,,m.H_{\boldsymbol{\nu}}(x_{1},\ldots,x_{m}):=\prod_{j=1}^{m}H_{\nu_{j}}(x_{j}),\;\;x_{j}\in\mathbb{R},\;j=1,\ldots,m\;.
Remark 2.3.

[Normalizations of Hermite polynomials and Hermite functions]

  1. (i)

    Definition (2.1) provides for every k0k\in{\mathbb{N}}_{0} a polynomial of degree kk. The scaling factor in (2.1) has been chosen to ensure normalization with respect to GM γ1\gamma_{1}, see also Lemma 2.4, item (i).

  2. (ii)

    Other normalizations with at times the same notation are used. The “classical” normalization of HkH_{k} we denote by H~k(x)\tilde{H}_{k}(x). It is defined by (see, e.g., [1, Page 787], and compare (2.1) with [105, Equation (5.5.3)])

    H~k(x/2):=2k/2k!Hk(x).\tilde{H}_{k}(x/\sqrt{2}):=2^{k/2}\sqrt{k!}H_{k}(x).
  3. (iii)

    In [7], so-called “normalized Hermite polynomials” are introduced as

    H~~k(x):=[π2kk!]1/2(1)kexp(x2)dkdxkexp(x2).\tilde{\tilde{H}}_{k}(x):=[\sqrt{\pi}2^{k}k!]^{1/2}(-1)^{k}\exp(x^{2})\frac{\,\mathrm{d}^{k}}{\,\mathrm{d}x^{k}}\exp(-x^{2})\;.

    The system (H~~k)k0(\tilde{\tilde{H}}_{k})_{k\in\mathbb{N}_{0}} is an orthonormal basis (ONB for short) for the space L2(,γ~~)L^{2}({\mathbb{R}},\tilde{\tilde{\gamma}}) with the weight γ~~=exp(x2)dx\tilde{\tilde{\gamma}}=\exp(-x^{2})\,\mathrm{d}x, i.e., (compare, e.g., [105, Eqn. (5.5.1)])

    H~~n(x)H~~n(x)exp(x2)dx=δnn,n,n0.\int_{{\mathbb{R}}}\tilde{\tilde{H}}_{n}(x)\tilde{\tilde{H}}_{n^{\prime}}(x)\exp(-x^{2})\,\mathrm{d}x=\delta_{nn^{\prime}},\;\;n,n^{\prime}\in{\mathbb{N}}_{0}\;.
  4. (iv)

    With the Hermite polynomials H~~k\tilde{\tilde{H}}_{k}, in [69] Hermite functions are introduced for k0k\in{\mathbb{N}}_{0} as

    hk(x):=exp(x2/2)H~~k(x),x.h_{k}(x):=\exp(-x^{2}/2)\tilde{\tilde{H}}_{k}(x)\;,\quad x\in{\mathbb{R}}\;.
  5. (v)

    It has been shown in [69, Theorem 1] that in order for functions f:f:{\mathbb{C}}\to{\mathbb{C}} defined in the strip S(ρ):={z:z=x+iy,x,|y|<ρ}S(\rho):=\{z\in{\mathbb{C}}:z=x+{\rm i}y,\;x\in{\mathbb{R}},\;|y|<\rho\} to admit a Fourier-Hermite expansion

    n=0fnhn(z),fn:=f(x)hn(x)dx=f(x)H~~n(x)exp(x2)dx\sum_{n=0}^{\infty}f_{n}h_{n}(z),\qquad f_{n}:=\int_{{\mathbb{R}}}f(x)h_{n}(x)\,\mathrm{d}x=\int_{{\mathbb{R}}}f(x)\tilde{\tilde{H}}_{n}(x)\exp(-x^{2})\,\mathrm{d}x

    which converges to f(z)f(z) for zS(ρ)z\in S(\rho) a necessary and sufficient condition is that a) ff is holomorphic in S(ρ)S(\rho)\subset{\mathbb{C}} and b) for every 0<ρ<ρ0<\rho^{\prime}<\rho there exist a finite bound B(ρ)B(\rho^{\prime}) and β\beta such that

    |f(x+iy)|B(ρ)exp[|x|(β2y2)1/2],x,|y|ρ.|f(x+{\rm i}y)|\leq B(\rho^{\prime})\exp[-|x|(\beta^{2}-y^{2})^{1/2}]\;,\quad x\in{\mathbb{R}},|y|\leq\rho^{\prime}\;.

    There is a constant C(f)>0C(f)>0 such that for the Fourier-Hermite coefficients fnf_{n}, holds

    |fn|Cexp(ρ2n+1)n0.|f_{n}|\leq C\exp(-\rho\sqrt{2n+1})\quad\forall n\in{\mathbb{N}}_{0}.

We state some basic properties of the Hermite polynomials HkH_{k} defined in (2.1).

Lemma 2.4.

The collection (Hk)k0(H_{k})_{k\in{\mathbb{N}}_{0}} of Hermite polynomials (2.1) in \mathbb{R} has the following properties.

  1. (i)

    (Hk)k0(H_{k})_{k\in{\mathbb{N}}_{0}} is an ONB of the space L2(;γ1)L^{2}(\mathbb{R};\gamma_{1}).

  2. (ii)

    For every kk\in{\mathbb{N}} holds: Hk(x)=kHk1(x)=Hk(x)k+1Hk+1(x)H_{k}^{\prime}(x)=\sqrt{k}H_{k-1}(x)=H_{k}(x)-\sqrt{k+1}H_{k+1}(x).

  3. (iii)

    For all x1,,xmx_{1},\ldots,x_{m}\in\mathbb{R} holds

    i=1mki!Hki(xi)=k1++kmt1k1tmkmexp(i=1mtixi12i=1mti2)t1==tm=0.\prod_{i=1}^{m}\sqrt{k_{i}!}H_{k_{i}}(x_{i})=\frac{\partial^{k_{1}+\ldots+k_{m}}}{\partial t_{1}^{k_{1}}\ldots\partial t_{m}^{k_{m}}}\exp\left(\sum_{i=1}^{m}t_{i}x_{i}-\frac{1}{2}\sum_{i=1}^{m}t_{i}^{2}\right)\mid_{t_{1}=\ldots=t_{m}=0}.
  4. (iv)

    For every fC()f\in C^{\infty}(\mathbb{R}) such that f(k)L2(;γ1)f^{(k)}\in L^{2}(\mathbb{R};\gamma_{1}) for all k0k\in{\mathbb{N}}_{0} holds

    f(x)Hk(x)dγ1(x)=(1)kk!f(k)(x)dγ1(x),\int_{\mathbb{R}}f(x)H_{k}(x)\,\mathrm{d}\gamma_{1}(x)=\frac{(-1)^{k}}{\sqrt{k!}\int_{\mathbb{R}}f^{(k)}(x)\,\mathrm{d}\gamma_{1}(x)}\;,

    and, hence, in L2(;γ1)L^{2}(\mathbb{R};\gamma_{1}),

    f=k0(1)kk!(f(k),1)L2(;γ1)Hk.f=\sum_{k\in{\mathbb{N}}_{0}}\frac{(-1)^{k}}{\sqrt{k!}}(f^{(k)},1)_{L^{2}(\mathbb{R};\gamma_{1})}H_{k}\;.

It follows from item (i) of this lemma in particular that

{H𝝂:𝝂0m}is an ONB ofL2(m;γm).\left\{H_{\boldsymbol{\nu}}:{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{m}\right\}\;\mbox{is an ONB of}\;L^{2}(\mathbb{R}^{m};\gamma_{m})\;.

Denote for k0k\in\mathbb{N}_{0} and mm\in\mathbb{N} by k{\mathcal{H}}_{k} the space of dd-variate Hermite polynomials which are homogeneous of degree kk, i.e.,

k:=span{H𝝂:𝝂0m,|𝝂|=k}.{\mathcal{H}}_{k}:={\rm span}\left\{H_{\boldsymbol{\nu}}:{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{m},|{\boldsymbol{\nu}}|=k\right\}\;.

Then k{\mathcal{H}}_{k} (“homogeneous polynomial chaos of degree kk[109]) is a closed, linear subspace of L2(m;γm)L^{2}(\mathbb{R}^{m};\gamma_{m}) and

L2(m;γm)=k0kinL2(m;γm).L^{2}(\mathbb{R}^{m};\gamma_{m})=\bigoplus_{k\in{\mathbb{N}}_{0}}{\mathcal{H}}_{k}\;\;\mbox{in}\;\;L^{2}(\mathbb{R}^{m};\gamma_{m})\;.

2.2 Gaussian measures on separable locally convex spaces

An important mathematical ingredient in a number of applications, in particular in UQ, Bayesian PDE inversion, risk analysis, but also in statistical learning theory applied to input-output maps for PDEs, is the construction of measures on function spaces. A particular interest is in GMs on separable on Hilbert or Banach or, more generally, on locally convex spaces of uncertain input data for PDEs. Accordingly, we review constructions of such measures, in terms of suitable bases of the input spaces. This implies, in particular, separability of the spaces of admissible PDE inputs or, at least, the uncertain input data being a separably-valued random element of otherwise nonseparable spaces (such as, e.g., L(D)L^{\infty}({{D}})) of valid inputs for the PDE of interest.

Let (Ω,𝒜,μ)(\Omega,\mathcal{A},\mu) be a measure space and 1p1\leq p\leq\infty. Recall that the normed space Lp(Ω,μ)L^{p}(\Omega,\mu) is defined as the space of all μ\mu-measurable functions uu from Ω\Omega to \mathbb{R} such that the norm

uLp(Ω,μ):=(Ω|u(x)|pdμ(x))1/p<.\|u\|_{L^{p}(\Omega,\mu)}:=\ \left(\int_{\Omega}|u(x)|^{p}\,\,\mathrm{d}\mu(x)\right)^{1/p}<\infty.

When p=p=\infty the norm of uL(Ω,μ)u\in L^{\infty}(\Omega,\mu) is given by

uL(Ω,μ):=esssupxΩ|u(x)|.\|u\|_{L^{\infty}(\Omega,\mu)}:=\underset{x\in\Omega}{\operatorname{ess\,sup}}|u(x)|.

If Ωm\Omega\subset{\mathbb{R}}^{m} and μ\mu is the Lebesgue measure, we simply denote these spaces by Lp(Ω)L^{p}(\Omega).

Throughout this section, XX will denote a real separable and locally convex space with Borel σ\sigma-field (X)\mathcal{B}(X) and with dual space XX^{*}.

Example 2.5.

Let {\mathbb{R}}^{\infty} be the linear space of all sequences 𝒚=(yj)j{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}} with yjy_{j}\in{\mathbb{R}}. This linear space becomes a locally convex space (still denoted by {\mathbb{R}}^{\infty}) equipped with the topology generated by the countable family of semi-norms

pj(𝒚):=|yj|,j.p_{j}({\boldsymbol{y}}):=|y_{j}|,\quad j\in\mathbb{N}.

The locally convex space {\mathbb{R}}^{\infty} is separable and complete and, therefore, a Fréchet space. However, it is not normable, and hence not a Banach space.

Example 2.6.

Let Dd{D}\subset\mathbb{R}^{d} be an open bounded Lipschitz domain.

  1. (i)

    The Banach spaces C(D¯)C(\overline{{D}}) and L1(D)L^{1}({D}) are separable.

  2. (ii)

    For 0<s<10<s<1 we denote by Cs(D)C^{s}({D}) the space of ss-Hölder continuous functions in D{D} equipped with the norm and seminorm

    aCs:=aL+|a|Cs,|a|Cs:=sup𝒙,𝒙D,𝒙𝒙|a(𝒙)a(𝒙)||𝒙𝒙|s.\|a\|_{C^{s}}:=\|a\|_{L^{\infty}}+|a|_{C^{s}}\;,\quad|a|_{C^{s}}:=\sup_{{\boldsymbol{x}},{\boldsymbol{x}}^{\prime}\in{D},{\boldsymbol{x}}\neq{\boldsymbol{x}}^{\prime}}\frac{|a({\boldsymbol{x}})-a({\boldsymbol{x}}^{\prime})|}{|{\boldsymbol{x}}-{\boldsymbol{x}}^{\prime}|^{s}}\;.

    Then the Banach space Cs(D)C^{s}({{D}}) is not separable. A separable subspace is

    Cs(D):={aCs(D):𝒙DlimD𝒙𝒙|a(𝒙)a(𝒙)||𝒙𝒙|s=0}.C^{s}_{\circ}({D}):=\bigg{\{}a\in C^{s}({D}):\forall{\boldsymbol{x}}\in{D}\;\lim_{{D}\ni{\boldsymbol{x}}^{\prime}\to{\boldsymbol{x}}}\frac{|a({\boldsymbol{x}})-a({\boldsymbol{x}}^{\prime})|}{|{\boldsymbol{x}}-{\boldsymbol{x}}^{\prime}|^{s}}=0\bigg{\}}\;.

We review and present constructions of GMs γ\gamma on XX.

2.2.1 Cylindrical sets

Cylindrical sets are subsets of XX of the form

C={xX:(l1(x),,ln(x))C0:C0(n),liX},for somen.C=\left\{x\in X:(l_{1}(x),\ldots,l_{n}(x))\in C_{0}:C_{0}\in\mathcal{B}(\mathbb{R}^{n}),l_{i}\in X^{*}\right\},\;\mbox{for some}\;n\in\mathbb{N}\;.

Here, the Borel set C0(n)C_{0}\in\mathcal{B}(\mathbb{R}^{n}) is sometimes referred to as basis of the cylinder CC. We denote by (X)\mathcal{E}(X) the σ\sigma-field generated by all cylindrical subsets of XX. It is the smallest σ\sigma-field for which all continuous linear functionals are measurable. Evidently then (X)(X)\mathcal{E}(X)\subset\mathcal{B}(X), with in general strict inclusion (see, e.g., [21, A.3.8]). If, however, XX is separable, then (X)=(X)\mathcal{E}(X)=\mathcal{B}(X) ([21, Theorem A.3.7]).

Sets of the form

{𝒚:(y1,,yn)B,B(n),n}\left\{{\boldsymbol{y}}\in\mathbb{R}^{\infty}:(y_{1},\ldots,y_{n})\in B,B\in\mathcal{B}(\mathbb{R}^{n}),n\in\mathbb{N}\right\}

generate ()\mathcal{B}(\mathbb{R}^{\infty}) [21, Lemma 2.1.1], and a set CC belongs to (X)\mathcal{B}(X) iff it is of the form

C={xX:(l1(x),,ln(x),)B,forliX,B()},C=\left\{x\in X:\,(l_{1}(x),\ldots,l_{n}(x),\ldots)\in B,\ \mbox{for}\ l_{i}\in X^{*},B\in\mathcal{B}(\mathbb{R}^{\infty})\right\}\;,

(see, e.g., [21, Lemma 2.1.2]).

2.2.2 Definition and basic properties of Gaussian measures

Definition 2.7 ([21, Definition 2.2.1]).

A probability measure γ\gamma defined on the σ\sigma-field (X)\mathcal{E}(X) generated by XX^{*} is called Gaussian if, for any fXf\in X^{*} the induced measure γf1\gamma\circ f^{-1} on \mathbb{R} is Gaussian. The measure γ\gamma is centered or symmetric if all measures γf1\gamma\circ f^{-1}, fXf\in X^{*} are centered.

Let (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}) be a probability space. A random field uu taking values in XX (recall that throughout, XX is a separable locally convex space) is a map u:ΩXu:\Omega\to X such that

B(X):u1(B)𝒜.\forall B\in{\mathcal{B}}(X):\;\;u^{-1}(B)\in{\mathcal{A}}\;.

The law of the random field uu is the probability measure 𝔪u{\mathfrak{m}}_{u} on (X,(X))(X,{\mathcal{B}}(X)) which is defined as

𝔪u(B):=(u1(B)),B(X).{\mathfrak{m}}_{u}(B):=\mathbb{P}(u^{-1}(B)),\quad B\in{\mathcal{B}}(X)\;.

The random field uu is said to be Gaussian if its law is a GM on (X,(X))(X,{\mathcal{B}}(X)).

Images of GMs under continuous affine transformations on XX are Gaussian.

Lemma 2.8 ([21, Lemma 2.2.2]).

Let γ\gamma be a GM on XX and let T:XYT:X\to Y be a linear map to another locally convex space YY such that lTXl\circ T\in X^{*} for all lYl\in Y^{*}. Then γT1\gamma\circ T^{-1} is a GM on YY.

This remains true for the affine map xTx+μx\mapsto Tx+\mu for some μY\mu\in Y.

The Fourier transform of a measure 𝔪{\mathfrak{m}} over (X,(X))(X,\mathcal{B}(X)) is given by

𝔪^:X:f𝔪^(f):=Xexp(if(x))d𝔪(x).\hat{{\mathfrak{m}}}:X^{*}\to\mathbb{C}:f\mapsto\hat{{\mathfrak{m}}}(f):=\int_{X}\exp\left({\rm i}f(x)\right)\,\mathrm{d}{\mathfrak{m}}(x)\;.
Theorem 2.9 ([21, Theorem 2.2.4]).

A measure γ\gamma on XX is Gaussian iff its Fourier transform γ^\hat{\gamma} can be expressed with some linear functional L()L(\cdot) on XX^{*} and a symmetric bilinear form B(.,.)B(.,.) on X×XX^{*}\times X^{*} such that fB(f,f)f\mapsto B(f,f) is nonnegative as

fX:γ^(f)=exp(iL(f)12B(f,f)).\forall f\in X^{*}:\quad\hat{\gamma}(f)=\exp\left({\rm i}L(f)-\frac{1}{2}B(f,f)\right). (2.2)

A GM γ\gamma on XX is therefore characterized by LL and BB. It also follows from (2.2) that a GM γ\gamma on XX is centered iff γ(A)=γ(A)\gamma(A)=\gamma(-A) for all A(X)A\in\mathcal{B}(X), i.e., iff L=0L=0 in (2.2).

Definition 2.10.

Let 𝔪{\mathfrak{m}} be a measure on (X)\mathcal{B}(X) such that XL2(X,𝔪)X^{*}\subset L^{2}(X,{\mathfrak{m}}). Then the element a𝔪(X)a_{\mathfrak{m}}\in(X^{*})^{\prime} in the algebraic dual (X)(X^{*})^{\prime} defined by

a𝔪(f):=Xf(x)d𝔪(x),fX,{a}_{\mathfrak{m}}(f):=\int_{X}f(x)\,\mathrm{d}{\mathfrak{m}}(x),\;f\in X^{*},

is called mean of 𝔪{\mathfrak{m}}.

The operator R𝔪:X(X)R_{\mathfrak{m}}:X^{*}\to(X^{*})^{\prime} defined by

R𝔪(f)(g):=X[f(x)a𝔪(f)][g(x)a𝔪(g)]d𝔪(x)R_{\mathfrak{m}}(f)(g):=\int_{X}[f(x)-{a}_{\mathfrak{m}}(f)][g(x)-{a}_{\mathfrak{m}}(g)]{\,\mathrm{d}{\mathfrak{m}}(x)}

is called covariance operator of 𝔪{\mathfrak{m}}. The quadratic form on XX^{*} is called covariance of 𝔪{\mathfrak{m}}.

When XX is a real separable Hilbert space, one can say more.

Definition 2.11 (Nuclear operators).

Let H1H_{1}, H2H_{2} be real separable Hilbert spaces with the norms H1\|\circ\|_{H_{1}} and H2\|\circ\|_{H_{2}}, respectively, and with corresponding inner products (,)Hi(\cdot,\cdot)_{H_{i}}, i=1,2i=1,2.

A linear operator K(H1,H2)K\in\mathcal{L}(H_{1},H_{2}) is called nuclear or trace class if it can be represented as

uH1:Ku=k(u,x1k)H1x2kinH2.\forall u\in H_{1}:\quad Ku=\sum_{k\in{\mathbb{N}}}(u,x_{1k})_{H_{1}}x_{2k}\;\mbox{in}\;H_{2}\;.

Here, (xik)kHi(x_{ik})_{k\in{\mathbb{N}}}\subset H_{i}, i=1,2i=1,2 are such that kx1kH1x2kH2<\sum_{k\in{\mathbb{N}}}\|x_{1k}\|_{H_{1}}\|x_{2k}\|_{H_{2}}<\infty.

We denote by 1(H1,H2)(H1,H2)\mathcal{L}_{1}(H_{1},H_{2})\subset\mathcal{L}(H_{1},H_{2}) the space of all nuclear operators. This is a separable Banach space when it is endowed with nuclear norm

K1:=inf{kx1kH1x2kH2:Ku=k(u,x1k)H1x2k}.\|K\|_{1}:=\inf\left\{\sum_{k\in{\mathbb{N}}}\|x_{1k}\|_{H_{1}}\|x_{2k}\|_{H_{2}}:Ku=\sum_{k\in{\mathbb{N}}}(u,x_{1k})_{H_{1}}x_{2k}\right\}\,.

When X=H1=H2X=H_{1}=H_{2}, we also write 1(X)\mathcal{L}_{1}(X).

Proposition 2.12 ([21, Theorem 2.3.1]).

Let γ\gamma be a GM on a separable Hilbert space XX with innerproduct (,)X(\cdot,\cdot)_{X}, and let XX^{*} denote its dual, identified with XX via the Riesz isometry.

Then there exist μX\mu\in X and a symmetric, nonnegative nuclear operator K1(X)K\in\mathcal{L}_{1}(X) such that the Fourier transform γ^\hat{\gamma} of γ\gamma is

γ^:X:xexp(i(μ,x)X12(Kx,x)X).\hat{\gamma}:X\to\mathbb{C}:x\mapsto\exp\left({\rm i}(\mu,x)_{X}-\frac{1}{2}(Kx,x)_{X}\right)\,. (2.3)
Remark 2.13.

Consider that XX is a real, separable Hilbert space with innerproduct (,)X(\cdot,\cdot)_{X} and assume given a GM γ\gamma on XX.

  1. (i)

    In (2.3), K(X)K\in\mathcal{L}(X) and μX\mu\in X are determined by

    u,vX:(μ,v)X=X(x,v)Xdγ(x),(Ku,v)X=X(u,xμ)X(v,xμ)Xdγ(x).\forall u,v\in X:(\mu,v)_{X}=\int_{X}(x,v)_{X}\,\mathrm{d}\gamma(x),\quad(Ku,v)_{X}=\int_{X}(u,x-\mu)_{X}(v,x-\mu)_{X}\,\mathrm{d}\gamma(x)\;.

    The closure of X=XX=X^{*} in L2(X;γ)L^{2}(X;\gamma) then equals the completion of XX with respect to the norm xK1/2xX=(Kx,x)Xx\mapsto\|K^{1/2}x\|_{X}=\sqrt{(Kx,x)_{X}}. Let (en)n(e_{n})_{n\in{\mathbb{N}}} denote the ONB of XX formed by eigenvectors of KK, with corresponding real, non-negative eigenvalues kn0k_{n}\in{\mathbb{N}}_{0}, i.e., Ken=knenKe_{n}=k_{n}e_{n} for n=1,2,n=1,2,\ldots. Then the completion can be identified with the weighted sequence (Hilbert) space

    {(xn)n:nknxn2<}.\left\{(x_{n})_{n\in{\mathbb{N}}}:\sum_{n\in{\mathbb{N}}}k_{n}x_{n}^{2}<\infty\right\}\;.

    The nuclear operator KK is the covariance of the GM γ\gamma on the Hilbert space XX.

  2. (ii)

    In coordinates 𝒚=(yj)j2(){\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}}\in\ell^{2}(\mathbb{N}) associated to the ONB (en)n(e_{n})_{n\in{\mathbb{N}}} of XX, (2.3) takes the form

    γ^:2():𝒚exp(inanyn12nknyn2).\hat{\gamma}:\ell^{2}(\mathbb{N})\to\mathbb{C}:{\boldsymbol{y}}\mapsto\exp\left({\rm i}\sum_{n\in{\mathbb{N}}}a_{n}y_{n}-\frac{1}{2}\sum_{n\in{\mathbb{N}}}k_{n}y_{n}^{2}\right)\;.
  3. (iii)

    Consider a=0Xa=0\in X and, for finite nn\in\mathbb{N}, a cylindrical set C=Pn1(B)C=P_{n}^{-1}(B) with PnP_{n} denoting the orthogonal projection onto Xn:=span{ej:j=1,,n}XX_{n}:={\rm span}\{e_{j}:j=1,\ldots,n\}\subset X, and with B(Xn)B\in\mathcal{B}(X_{n}). Then

    γ(C)=Bj=1n(2πkj)1/2exp(12kjyj2)dy1dyn.\gamma(C)=\int_{B}\prod_{j=1}^{n}(2\pi k_{j})^{-1/2}\exp\left(-\frac{1}{2k_{j}}y_{j}^{2}\right)\,\mathrm{d}y_{1}\ldots\,\mathrm{d}y_{n}\;.

For fXf\in X^{*} and xXx\in X, one frequently writes the X×XX^{*}\times X duality pairing as

f(x)=f,x.f(x)=\langle f,x\rangle\;.

With the notation from Definition 2.10, the covariance operator Cg=RγgC_{g}=R_{\gamma_{g}} in Definition 2.10 of a centered, Gaussian random vector g:(Ω,𝒜;γg)Xg:(\Omega,\mathcal{A};\gamma_{g})\to X with Gaussian law γg\gamma_{g} on a separable, real Banach space XX admits the representations

Rγg=Cg:XX:Cgφ:=𝔼φ,gg,Cg:X×X:(ψ,φ)ψ,Cgφ.R_{\gamma_{g}}=C_{g}:X^{*}\to X:C_{g}\varphi:=\mathbb{E}\langle\varphi,g\rangle g,\quad C_{g}:X^{*}\times X^{*}\to\mathbb{R}:(\psi,\varphi)\mapsto\langle\psi,C_{g}\varphi\rangle\;.

2.3 Cameron-Martin space

Let XX be a real separable locally convex space and γ\gamma a GM on (X)\mathcal{E}(X) such that XL2(X;γ)X^{*}\subset L^{2}(X;\gamma). Then, for every φX\varphi\in X^{*}, the image measure φ(γ)\varphi(\gamma) is a GM on \mathbb{R}. By [21, Theorem 3.2.3], there exists a unique aγXa_{\gamma}\in X, the mean of γ\gamma, such that

φX:φ(aγ)=Xφ(h)dγ(h).\forall\varphi\in X^{*}:\quad\varphi(a_{\gamma})=\int_{X}\varphi(h)\,\mathrm{d}\gamma(h)\;.

Denote by XγX^{*}_{\gamma} the closure of the set {φφ(aγ),φX)}\{\varphi-\varphi({a_{\gamma})},\,\varphi\in X^{*})\} embedded into the normed space L2(X;γ)L^{2}(X;\gamma) w.r.t.  its norm.

The covariance operator, RγR_{\gamma}, of γ\gamma is formally given by

φ,ψX:Rγφ,ψ=Xφ(haγ)ψ(haγ)dγ(h).\forall\varphi,\psi\in X^{*}:\quad\langle R_{\gamma}\varphi,\psi\rangle=\int_{X}\varphi(h-{a_{\gamma}})\psi(h-{a_{\gamma}})\,\mathrm{d}\gamma(h)\;. (2.4)

As XX is a separable locally convex space, [21, Theorem 3.2.3] implies that there is a unique linear operator Rγ:XXR_{\gamma}:X^{*}\to X such that (2.4) holds. We define

φX:σ(φ):=Rγφ,φ.\forall\varphi\in X^{*}:\quad\sigma(\varphi):=\sqrt{\langle R_{\gamma}\varphi,\varphi\rangle}\;.

If h=Rγφh=R_{\gamma}\varphi for some φX\varphi\in X^{*}, the map hh:=σ(φ)h\mapsto\|h\|:=\sigma(\varphi) defines a norm on range(Rγ)X{\rm range}(R_{\gamma})\subset X. There holds [21, Lemma 2.4.1] h=hH(γ)=φL2(X;γ)\|h\|=\|h\|_{H(\gamma)}=\|\varphi\|_{L^{2}(X;\gamma)}.

The Cameron-Martin space of the GM γ\gamma on XX is the completion of the range of RγR_{\gamma} in XX with respect to the norm \|\circ\|. The Cameron-Martin space of the GM γ\gamma on XX is denoted by H(γ)H(\gamma). It is also called the reproducing kernel Hilbert space (RKHS for short) of γ\gamma on XX.

By [21, Theorem 3.2.7], H(γ)H(\gamma) is a separable Hilbert space, and H(γ)XH(\gamma)\subset X with continuous embedding, according to [21, Proposition 2.4.6]. In case that XYX\subset Y for another Banach space, with continuous and linear embedding, the Cameron-Martin spaces for XX and YY coincide. For example, in the context of Remark 2.13, item (i), H(γ)=K(Xγ)H(\gamma)=K(X^{*}_{\gamma}).

Being a Hilbert space, introduce an innerproduct (,)H(γ)(\cdot,\cdot)_{H(\gamma)} on H(γ)H(\gamma) compatible with the norm H(γ)\|\circ\|_{H(\gamma)} via the parallelogram law. Then there holds

φXfH(γ):(f,Rγφ)H(γ)=φ(f).\forall\varphi\in X^{*}\;\forall f\in H(\gamma):\quad(f,R_{\gamma}\varphi)_{H(\gamma)}=\varphi(f)\;.

Since H(γ)H(\gamma) is also separable, there is an ONB.

Proposition 2.14 ([21, Theorem 3.5.10, Corollary 3.5.11]).

For a centered GM on a real, separable Banach space XX with norm X\|\circ\|_{X}, there exists an ONB (en)n(e_{n})_{n\in\mathbb{N}} of the Cameron-Martin space H(γ)XH(\gamma)\subset X such that

nenX2<,φX:Rγφ=nφ(en)en.\sum_{n\in{\mathbb{N}}}\|e_{n}\|_{X}^{2}<\infty\;,\qquad\forall\varphi\in X^{*}:\;R_{\gamma}\varphi=\sum_{n\in{\mathbb{N}}}\varphi(e_{n})e_{n}\;.

We remark that Proposition 2.14 is not true for arbitrary ONB (en)n(e_{n})_{n\in{\mathbb{N}}} of H(γ)H(\gamma).

2.4 Gaussian product measures

We recall a notion of product measures which gives an efficient method to construct Gaussian measures on a countable Cartesian product of locally convex spaces.

Definition 2.15 (Product measure, [21, p. 372]).

Let μn\mu_{n} be probability measures defined on σ\sigma-fields n\mathcal{B}_{n} in locally convex spaces XnX_{n}. Put

X:=nXn.X:=\prod_{n\in\mathbb{N}}X_{n}.

Let

:=nn\mathcal{B}:=\bigotimes_{n\in\mathbb{N}}\mathcal{B}_{n}

be the σ\sigma-field generated by all the sets of the form

B=B1×B2××Bn×Xn+1×Xn+2×,Bii.B=B_{1}\times B_{2}\times\ldots\times B_{n}\times X_{n+1}\times X_{n+2}\times\ldots,\ B_{i}\in\mathcal{B}_{i}. (2.5)

The product measure

μ:=nμn\mu:=\bigotimes_{n\in\mathbb{N}}\mu_{n}

is the probability measure on \mathcal{B} defined by μ(B):=i=1nμi(Bi)\mu(B):=\prod_{i=1}^{n}\mu_{i}(B_{i}) for the sets BB of the form (2.5).

Example 2.16 ([21, Example 2.3.8]).

Let (μn)n(\mu_{n})_{n\in\mathbb{N}} be a sequence of GMs. Then the product measure μ:=nμn\mu:=\otimes_{n\in\mathbb{N}}\mu_{n} is a GM on X:=nXnX:=\prod_{n\in\mathbb{N}}X_{n}. The Cameron-Martin space H(μ)H(\mu) of μ\mu is the Hilbert direct sum of spaces H(μn)H(\mu_{n}), i.e.,

H(μ)={h=(hj)jX:hjH(μj),hH(μ)2=jhjH(μj)2}.H(\mu)=\left\{h=(h_{j})_{j\in\mathbb{N}}\in X:\,h_{j}\in H(\mu_{j}),\|h\|^{2}_{H(\mu)}=\sum_{j\in\mathbb{N}}\|h_{j}\|^{2}_{H(\mu_{j})}\right\}.

The space XμX^{*}_{\mu} is the set of all functions of the form

φjfj(φj),fjXμj,jσ(fj)2<,\varphi\mapsto\sum_{j\in\mathbb{N}}f_{j}(\varphi_{j}),\quad f_{j}\in X^{*}_{\mu_{j}},\quad\sum_{j\in\mathbb{N}}\sigma(f_{j})^{2}<\infty,

and

aμ(f)=jaμj(fj),f=(fj)jX.a_{\mu}(f)=\sum_{j\in\mathbb{N}}a_{\mu_{j}}(f_{j}),\quad\forall f=(f_{j})_{j\in\mathbb{N}}\in X^{*}.
Example 2.17 ([21, Example 2.3.5]).

Denote by (γ1,n)n(\gamma_{1,n})_{n\in{\mathbb{N}}} a sequence of standard GMs on (,())(\mathbb{R},\mathcal{B}(\mathbb{R})). Then the product measure

γ=nγ1,n\gamma=\bigotimes_{n\in{\mathbb{N}}}\gamma_{1,n}

is a centered GM on \mathbb{R}^{\infty}. Furthermore, H(γ)=2()H(\gamma)=\ell^{2}(\mathbb{N}) and Xγ2()X^{*}_{\gamma}\simeq\ell^{2}(\mathbb{N}). If μ\mu is a GM on {\mathbb{R}}^{\infty}, then by a result of Fernique, the measures γ\gamma and μ\mu are either mutually singular or equivalent [21, Theorem 2.12.9]. The locally convex space {\mathbb{R}}^{\infty} with the product measure γ\gamma of standard GMs is the main parametric domain in the stochastic setting of UQ problems for PDEs with GRF inputs considered in this text.

2.5 Gaussian series

A key role in the numerical analysis of PDEs with GRF inputs from separable Banach spaces EE is played by representing these GRFs in terms of series with respect to suitable representation systems (ψj)jE(\psi_{j})_{j\in{\mathbb{N}}}\in E^{\infty} of EE with random coefficients. There arises the question of admissibility of (ψj)jE(\psi_{j})_{j\in{\mathbb{N}}}\in E^{\infty} so as to allow a) to transfer randomness of function space-valued inputs to a parametric, deterministic representation (as is customary, for example, in the transition from nonparametric to parametric models in statistics) and b) to ensure suitability for numerical approximation.

Items a) and b) are closely related to the selection of stable bases for EE, with item b) mandating additional requirements, such as efficient accessibility for float point computations, quadrature, etc.

We first present an abstract result, Theorem 2.21 and then, in Sections 2.5.2 and 2.5.3, we review several concrete constructions of such series. We discuss in Sections 2.5.2 and 2.5.3 several examples, in particular the classical Karhunen-Loève Expansion [77, 101] of GRFs taking values in separable Hilbert space. All examples will be admissible in parametrizing GRF input data for PDEs and of Gaussian priors in the ensuing sparsity and approximation rate analysis in Section 3 and the following sections.

2.5.1 Some abstract results

We place ourselves in the setting of a real separable locally convex space XX, with a GM γ\gamma on XX, and with associated Cameron-Martin Hilbert space H(γ)XH(\gamma)\subset X as introduced in Section 2.3.

We first consider expansions of Gaussian random vectors with respect to orthonormal bases (ej)j(e_{j})_{j\in\mathbb{N}} of the Cameron-Martin space H(γ)H(\gamma). As linear transformations of GM are Gaussian (see Lemma 2.8), we admit a linear transformation AA.

Theorem 2.18 ([21, Theorems. 3.5.1, 3.5.7, (3.5.4)]).

Let γ\gamma be a centered GM on a real separable locally convex space XX with Cameron-Martin space H(γ)H(\gamma) and with some ONB (ej)j(e_{j})_{j\in{\mathbb{N}}} of H(γ)H(\gamma). Let further denote (yj)j(y_{j})_{j\in{\mathbb{N}}} any sequence of independent standard Gaussian RVs on a probability space (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}) and let A(H(γ))A\in\mathcal{L}(H(\gamma)) be arbitrary.

Then the Gaussian series

jyj(ω)Aej\sum_{j\in{\mathbb{N}}}y_{j}(\omega)Ae_{j}

converges \mathbb{P}-a.s. in XX. The law of its limit is a centered GM λ\lambda with covariance RλR_{\lambda} given by

Rλ(f)(g)=(ARγ(f),ARγ(g))H(γ).R_{\lambda}(f)(g)=\left(A^{*}R_{\gamma}(f),A^{*}R_{\gamma}(g)\right)_{H(\gamma)}\;.

Furthermore, there holds of independent standard Gaussian RVs on a probability space (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}).

Xf(x)γ(dx)=Ωf(jyj(ω)ej)d(ω).\int_{X}f(x)\gamma(\,\mathrm{d}x)=\int_{\Omega}f\bigg{(}\sum_{j\in{\mathbb{N}}}y_{j}(\omega)e_{j}\bigg{)}\,\mathrm{d}\mathbb{P}(\omega)\;.

If XX is a real separable Banach space XX with norm X\|\circ\|_{X}, for all sufficiently small constants c>0c>0 holds

limnΩexp(cj=nyj(ω)AejX2)d(ω)=1.\lim_{n\to\infty}\int_{\Omega}\exp\Bigg{(}c\bigg{\|}\sum_{j=n}^{\infty}y_{j}(\omega)Ae_{j}\bigg{\|}_{X}^{2}\Bigg{)}\,\mathrm{d}\mathbb{P}(\omega)=1\,.

In particular, for every p[1,)p\in[1,\infty) we have j=nyjAejXp0\big{\|}\sum_{j=n}^{\infty}y_{j}Ae_{j}\big{\|}_{X}^{p}\to 0 in L1(Ω,)L^{1}(\Omega,\mathbb{P}) as nn\to\infty.

Often, in numerical applications, ensuring orthonormality of the basis elements could be computationally costly. It is therefore of some interest to consider Gaussian series with respect to more general representation systems (ψj)j(\psi_{j})_{j\in\mathbb{N}}. An important notion is admissibility of such systems.

Definition 2.19.

Let XX be a real, separable locally convex space, and let g:(Ω,𝒜,)Xg:(\Omega,\mathcal{A},\mathbb{P})\to X be a centered Gaussian random vector with law γg=X\gamma_{g}=\mathbb{P}_{X}. Let further (yj)j(y_{j})_{j\in\mathbb{N}} be a sequence of i.i.d. standard real Gaussian RVs yj𝒩(0,1)y_{j}\sim\mathcal{N}(0,1).

A sequence (ψj)jX(\psi_{j})_{j\in{\mathbb{N}}}\in X^{\infty} is called admissible for gg if

jyjψjconverges -a.s. in Xandg=jyjψj.\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}\;\;\mbox{converges $\mathbb{P}$-a.s. in }\;X\;\;\mbox{and}\;\;g=\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}\;.

To state the next theorem, we recall the notion of frames in separable Hilbert space (see, e.g., [65] and the references there for background and theory of frames. In the terminology of frame theory, Parseval frames correspond to tight frames with frame bounds equal to 11).

Definition 2.20.

A sequence (ψj)jH(\psi_{j})_{j\in\mathbb{N}}\subset H in a real separable Hilbert space HH with inner product (,)H(\cdot,\cdot)_{H} is a Parseval frame of HH if

fH:f=j(ψj,f)HψjinH.\forall f\in H:\ \ \ f=\sum_{j\in{\mathbb{N}}}(\psi_{j},f)_{H}\,\psi_{j}\ \ \mbox{in}\ \ H\;.

The following result, from [88], characterizes admissible affine representation systems for GRFs uu taking values in real, separable Banach spaces XX.

Theorem 2.21 ([88, Theorem 1]).

We have the following.

  1. (i)

    In a real, separable Banach space XX with a centered GM γ\gamma on XX, a representation system 𝚿=(ψj)jX{{\boldsymbol{\Psi}}}=(\psi_{j})_{j\in{\mathbb{N}}}\in X^{\infty} is admissible for γ\gamma iff 𝚿{{\boldsymbol{\Psi}}} is a Parseval frame for the Cameron-Martin space H(γ)XH(\gamma)\subset X, i.e.,

    fH(γ):fH(γ)2=j|f,ψj|2.\forall f\in H(\gamma):\quad\|f\|_{H(\gamma)}^{2}=\sum_{j\in{\mathbb{N}}}|\langle f,\psi_{j}\rangle|^{2}.
  2. (ii)

    Let uu denote a GRF taking values in XX with law γ\gamma and with RKHS H(γ)H(\gamma). For a countable collection 𝚿=(ψj)jX{{\boldsymbol{\Psi}}}=(\psi_{j})_{j\in{\mathbb{N}}}\in X^{\infty} the following are equivalent:

    • (i)

      𝚿{{\boldsymbol{\Psi}}} is a Parseval frame of H(γ)H(\gamma) and

    • (ii)

      there is a sequence 𝒚=(yj)j{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}} of i.i.d standard Gaussian RVs yjy_{j} such that there holds γa.s.\gamma-a.s. the representation

      u=jyjψjinH(γ).u=\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}\quad\mbox{in}\quad H(\gamma)\;.
  3. (iii)

    Consider a GRF uu taking values in XX with law γ\gamma and covariance Rγ(X,X)R_{\gamma}\in\mathcal{L}(X^{\prime},X). If Rγ=SSR_{\gamma}=SS^{\prime} with S(K,X)S\in\mathcal{L}(K,X) for some separable Hilbert space KK, for any Parseval frame 𝚽=(φj)j{{\boldsymbol{\Phi}}}=(\varphi_{j})_{j\in{\mathbb{N}}} of KK, the countable collection 𝚿=S𝚽=(Sφj)j{{\boldsymbol{\Psi}}}=S{{\boldsymbol{\Phi}}}=(S\varphi_{j})_{j\in{\mathbb{N}}} is a Parseval frame of the RKHS H(γ)H(\gamma) of uu.

The last assertion in the preceding result is [88, Proposition 1]. It generalizes the observation that for a symmetric positive definite matrix 𝑴{\boldsymbol{M}} in d{\mathbb{R}}^{d}, any factorization 𝑴=𝑳𝑳{\boldsymbol{M}}={\boldsymbol{L}}{\boldsymbol{L}}^{\top} implies that for z𝒩(0,𝑰)z\sim{\mathcal{N}}(0,{\boldsymbol{I}}) it holds 𝑳z𝒩(0,𝑴){\boldsymbol{L}}z\sim{\mathcal{N}}(0,{\boldsymbol{M}}). The result is useful in building customized representation systems 𝚿{{\boldsymbol{\Psi}}} which are frames of a GRF uu with computationally convenient properties in particular applications.

We review several widely used constructions of Parseval frames. These comprise expansions in eigenfunctions of the covariance operator KK (referred to also as principal component analysis, or as “Karhunen-Loève expansions”), but also “eigenvalue-free” multiresolution constructions (generalizing the classical Lévy-Cieselski construction of the Brownian bridge) for various geometric settings, in particular bounded subdomains of euclidean space, compact manifolds without boundary etc. Any of these constructions will be admissible choices as representation system of the GRF input of PDEs to render these PDEs parametric-deterministic where, in turn, our parametric regularity results will apply.

Example 2.22 (Brownian bridge).

On the bounded time interval [0,T][0,T], consider the Brownian bridge (Bt)t0(B_{t})_{t\geq 0}. It is defined in terms of a Wiener process (Wt)t0(W_{t})_{t\geq 0} by conditioning as

(Bt)0tT:={(Wt)0tT|WT=0}.(B_{t})_{0\leq t\leq T}:=\big{\{}(W_{t})_{0\leq t\leq T}|W_{T}=0\big{\}}. (2.6)

It is a simple example of kriging applied to the GRF WtW_{t}.

The covariance function of the GRF BtB_{t} is easily calculated as

kB(s,t)=𝔼[BsBt]=s(Tt)/Tifs<t.k_{B}(s,t)=\mathbb{E}[B_{s}B_{t}]=s(T-t)/T\;\;\mbox{if}\;\;s<t.

Various other representations of BtB_{t} are

Bt=WttTWT=TtTWt/(Tt).B_{t}=W_{t}-\frac{t}{T}W_{T}=\frac{T-t}{\sqrt{T}}W_{t/(T-t)}.

The RKHS H(γ)H(\gamma) corresponding to the GRF BtB_{t} is the Sobolev space H01(0,T)H^{1}_{0}(0,T).

2.5.2 Karhunen-Loève expansion

A widely used representation system in the analysis and computation of GRFs is the so-called Karhunen-Loève expansion (KL expansion for short) of GRFs, going back to [77]. We present main ideas and definitions, in a generic setting of [79], see also [3, Chap. 3.3].

Let \mathcal{M} be a compact space with metric ρ:×\rho:\mathcal{M}\times\mathcal{M}\to{\mathbb{R}} and with Borel sigma-algebra =()\mathcal{B}=\mathcal{B}(\mathcal{M}). Assume given a Borel measure μ\mu on (,)(\mathcal{M},\mathcal{B}). Let further (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}) be a probability space. Examples are =D\mathcal{M}={{D}} a bounded domain in Euclidean space d{\mathbb{R}}^{d}, with ρ\rho denoting the Euclidean distance between pairs (x,x)(x,x^{\prime}) of points in D{{D}}, and \mathcal{M} being a smooth, closed 22-surface in 3{\mathbb{R}}^{3}, where ρ\rho is the geodesic distance between pairs of points in \mathcal{M}.

Consider a measurable map

Z:(,)(Ω,𝒜):(x,ω)Zx(ω)Z:(\mathcal{M},\mathcal{B})\otimes(\Omega,\mathcal{A})\to{\mathbb{R}}:(x,\omega)\mapsto Z_{x}(\omega)\in{\mathbb{R}}

such that for each xx\in\mathcal{M}, ZxZ_{x} is a centered, Gaussian RV. We call the collection (Zx)x(Z_{x})_{x\in\mathcal{M}} a GRF indexed by \mathcal{M}.

Assume furthermore for all nn\in{\mathbb{N}}, for all x1,,xnx_{1},\ldots,x_{n}\in\mathcal{M} and for every ξ1,,ξn\xi_{1},\ldots,\xi_{n}\in{\mathbb{R}}

i=1nξiZxiis a centered Gaussian RV.\sum_{i=1}^{n}\xi_{i}Z_{x_{i}}\;\;\mbox{is a centered Gaussian RV}.

Then the covariance function

K:×:(x,x)K(x,x)K:\mathcal{M}\times\mathcal{M}\to{\mathbb{R}}:(x,x^{\prime})\mapsto K(x,x^{\prime})

associated with the centered GRF (Zx)x(Z_{x})_{x\in\mathcal{M}} is defined pointwise by

K(x,x):=𝔼[ZxZx]x,x.K(x,x^{\prime}):=\mathbb{E}[Z_{x}Z_{x^{\prime}}]\quad x,x^{\prime}\in\mathcal{M}\;.

Evidently, the covariance function K:×K:\mathcal{M}\times\mathcal{M}\to{\mathbb{R}} corresponding to a Gaussian RV indexed by \mathcal{M} is a real-valued, symmetric, and positive definite function, i.e., there holds

n(xj)1jnn,(ξj)1jnn:1i,jnξiξjK(xi,xj)0.\forall n\in{\mathbb{N}}\;\forall(x_{j})_{1\leq j\leq n}\in\mathcal{M}^{n},\forall(\xi_{j})_{1\leq j\leq n}\in{\mathbb{R}}^{n}:\;\sum_{1\leq i,j\leq n}\xi_{i}\xi_{j}K(x_{i},x_{j})\geq 0\;.

The operator K(L2(,μ),L2(,μ))K\in\mathcal{L}(L^{2}(\mathcal{M},\mu),L^{2}(\mathcal{M},\mu)) defined by

fL2(,μ):(Kf)(x):=K(x,x)f(x)dμ(x)x\forall f\in L^{2}(\mathcal{M},\mu):\;\;(Kf)(x):=\int_{\mathcal{M}}K(x,x^{\prime})f(x^{\prime})\,\mathrm{d}\mu(x^{\prime})\quad x\in\mathcal{M}

is a self-adjoint, compact positive operator on L2(,μ)L^{2}(\mathcal{M},\mu). Furthermore, KK is trace-class and K(L2(,μ))C(,)K(L^{2}(\mathcal{M},\mu))\subset C(\mathcal{M},{\mathbb{R}}).

The spectral theorem for compact, self-adjoint operators on the separable Hilbert space L2(,μ)L^{2}(\mathcal{M},\mu) ensures the existence of a sequence λ1λ20\lambda_{1}\geq\lambda_{2}\geq\ldots\geq 0 of real eigenvalues of KK (counted according to multiplicity and accumulating only at zero) with associated eigenfunctions ψkL2(,μ)\psi_{k}\in L^{2}(\mathcal{M},\mu) normalized in L2(,μ)L^{2}(\mathcal{M},\mu), i.e., for all kk\in{\mathbb{N}} holds

Kψk=λkψkinL2(,μ),ψk(x)ψ(x)dμ(x)=δk,k,.K\psi_{k}=\lambda_{k}\psi_{k}\quad\mbox{in}\;\;L^{2}(\mathcal{M},\mu)\;,\quad\int_{\mathcal{M}}\psi_{k}(x)\psi_{\ell}(x)\,\mathrm{d}\mu(x)=\delta_{k\ell}\;,\;\;k,\ell\in{\mathbb{N}}\;.

Then, there holds ψkC(;)\psi_{k}\in C(\mathcal{M};{\mathbb{R}}) and the sequence (ψk)k(\psi_{k})_{k\in{\mathbb{N}}} is an ONB of L2(,μ)L^{2}(\mathcal{M},\mu). From Mercer’s theorem (see, e.g., [101]), there holds the Mercer expansion

x,x:K(x,x)=kλkψk(x)ψk(x)\forall x,x^{\prime}\in\mathcal{M}:\quad K(x,x^{\prime})=\sum_{k\in{\mathbb{N}}}\lambda_{k}\psi_{k}(x)\psi_{k}(x^{\prime})

with absolute and uniform convergence on ×\mathcal{M}\times\mathcal{M}. This result implies that

limm×|K(x,x)j=1mλjψj(x)ψj(x)|2dμ(x)dμ(x)=0.\lim_{m\to\infty}\int_{\mathcal{M}\times\mathcal{M}}\left|K(x,x^{\prime})-\sum_{j=1}^{m}\lambda_{j}\psi_{j}(x)\psi_{j}(x^{\prime})\right|^{2}\,\mathrm{d}\mu(x)\,\mathrm{d}\mu(x^{\prime})=0\;.

We denote by HL2(Ω,)H\subset L^{2}(\Omega,\mathbb{P}) the L2(,μ)L^{2}(\mathcal{M},\mu) closure of finite linear combinations of (Zx)x(Z_{x})_{x\in\mathcal{M}}. This so-called Gaussian space (e.g. [75]) is a Hilbert space when equipped with the L2(,μ)L^{2}(\mathcal{M},\mu) innerproduct. Then, the sequence (Bk)k(B_{k})_{k\in{\mathbb{N}}}\subset{\mathbb{R}} defined by

k:Bk(ω):=1λkZx(ω)ψk(x)dμ(x)H\forall k\in{\mathbb{N}}:\quad B_{k}(\omega):=\frac{1}{\sqrt{\lambda_{k}}}\int_{\mathcal{M}}Z_{x}(\omega)\psi_{k}(x)\,\mathrm{d}\mu(x)\in H

is a sequence of i.i.d, N(0,1){N}(0,1) RVs. The expression

Z~x(ω):=kλkψk(x)Bk(ω)\tilde{Z}_{x}(\omega):=\sum_{k\in{\mathbb{N}}}\sqrt{\lambda_{k}}\psi_{k}(x)B_{k}(\omega) (2.7)

is a modification of Zx(ω)Z_{x}(\omega), i.e., for every xx\in\mathcal{M} holds that ({Zx=Z~x})=1\mathbb{P}(\{Z_{x}=\tilde{Z}_{x}\})=1, which is referred to as Karhunen-Loève expansion of the GRF {Zx:x}\{Z_{x}:x\in\mathcal{M}\}.

Example 2.23.

[KL expansion of the Brownian bridge (2.6)] On the compact interval =[0,T]\mathcal{M}=[0,T]\subset{\mathbb{R}}, the KL expansion of the Brownian bridge is

Bt=kZk2Tkπsin(kπt/T),t[0,T].B_{t}=\sum_{k\in{\mathbb{N}}}Z_{k}\frac{\sqrt{2T}}{k\pi}\sin(k\pi t/T)\;,\quad t\in[0,T]\;.

Then

H(γ)=H01(0,T)=span{sin(kπt/T):k}.H(\gamma)=H^{1}_{0}(0,T)={\rm span}\{\sin(k\pi t/T):k\in{\mathbb{N}}\}.

In view of GRFs appearing as diffusion coefficients in elliptic and parabolic PDEs, criteria on their path regularity are of some interest. Many such conditions are known and we present some of these, from [3, Chapter 3.2, 3.3].

Proposition 2.24.

For any compact set d\mathcal{M}\subset{\mathbb{R}}^{d}, if for α>0\alpha>0, η>α\eta>\alpha and some constant C>0C>0 holds

𝔼[|Z𝒙+𝒉Z𝒙|α]C|𝒉|2d|log|𝒉||1+η,\mathbb{E}[|Z_{{\boldsymbol{x}}+{\boldsymbol{h}}}-Z_{\boldsymbol{x}}|^{\alpha}]\leq C\frac{|{\boldsymbol{h}}|^{2d}}{|\log|{\boldsymbol{h}}||^{1+\eta}}\;, (2.8)

then

𝒙Z𝒙(ω)C0()a.s.{\boldsymbol{x}}\to Z_{\boldsymbol{x}}(\omega)\in C^{0}(\mathcal{M})\ \ \mathbb{P}-a.s.

Choosing α=2\alpha=2 in (2.8), we obtain for \mathcal{M} such that =D¯\mathcal{M}=\overline{{D}}, where Dd{D}\subset{\mathbb{R}}^{d} is a bounded Lipschitz domain, the sufficient criterion that there exist C>0C>0, η>2\eta>2 with

𝒙D:K(𝒙+𝒉,𝒙+𝒉)K(𝒙+𝒉,𝒙)K(𝒙,𝒙+𝒉)+K(𝒙,𝒙)C|𝒉|2d|log|𝒉||1+η.\forall{\boldsymbol{x}}\in{D}:\quad K({\boldsymbol{x}}+{\boldsymbol{h}},{\boldsymbol{x}}+{\boldsymbol{h}})-K({\boldsymbol{x}}+{\boldsymbol{h}},{\boldsymbol{x}})-K({\boldsymbol{x}},{\boldsymbol{x}}+{\boldsymbol{h}})+K({\boldsymbol{x}},{\boldsymbol{x}})\leq C\frac{|{\boldsymbol{h}}|^{2d}}{|\log|{\boldsymbol{h}}||^{1+\eta}}\;.

This is to hold for some η>2\eta>2 with the covariance kernel KK of the GRF ZZ, in order to ensure that [𝒙Z𝒙]C1(D¯)W1(D)[{\boldsymbol{x}}\mapsto Z_{\boldsymbol{x}}]\in C^{1}(\overline{{D}})\subset W^{1}_{\infty}({D}) \mathbb{P}-a.s., see [3, Theorem 3.2.5, page 49 bottom].

Further examples of explicit Karhunen-Loève expansions of GRFs can be found in [84, 35, 79] and a statement for \mathbb{P}-a.s Hölder continuity of GRFs ZZ on smooth manifolds \mathcal{M} is proved in [4].

2.5.3 Multiresolution representations of GRFs

Karhunen-Loève expansions (2.7) provide an important source of concrete examples of Gaussian series representations of GRFs uu in Theorem 2.18. Since KL expansions involve the eigenfunctions of the covariance operators of the GRF uu, all terms in these expansions are, in general, globally supported in the physical domain \mathcal{M} indexing the GRF uu. Often, it is desirable to have Gaussian series representations of uu in Theorem 2.18 where the elements (en)n(e_{n})_{n\in{\mathbb{N}}} of the representation system are locally supported in the indexing domain \mathcal{M}.

Example 2.25 (Lévy-Cieselsky representation of Brownian bridge, [34]).

Consider the Brownian bridge (Bt)0tT(B_{t})_{0\leq t\leq T} from Examples 2.22, 2.23. For T=1T=1, it may also be represented as Gaussian series (e.g. [34])

Bt=jk=02j1Zjk2j/2h(2jtk)=jk=02j1Zjkψjk(t),t=[0,1],B_{t}=\sum_{j\in{\mathbb{N}}}\sum_{k=0}^{2^{j}-1}Z_{jk}2^{-j/2}h(2^{j}t-k)=\sum_{j\in{\mathbb{N}}}\sum_{k=0}^{2^{j}-1}Z_{jk}\psi_{jk}(t),\quad t\in\mathcal{M}=[0,1]\;,

where

ψjk(t):=2j/2h(2jtk),\psi_{jk}(t):=2^{-j/2}h(2^{j}t-k),

with h(s):=max{12|s1/2|,0}h(s):=\max\{1-2|s-1/2|,0\} denoting the standard, continuous piecewise affine “hat” function on (0,1)(0,1). Here, μ\mu is the Lebesgue measure in =[0,1]\mathcal{M}=[0,1], and Zjk𝒩(0,1)Z_{jk}\sim\mathcal{N}(0,1) are i.i.d standard normal RVs.

By suitable reordering of the index pairs (j,k)(j,k), e.g., via the bijection (j,k)j:=2j+k(j,k)\mapsto j:=2^{j}+k, the representation (2.25) is readily seen to be a special case of Theorem 2.21, item ii). The corresponding system

𝚿={ψjk:j0,0k2j1}{{\boldsymbol{\Psi}}}=\{\psi_{jk}:j\in{\mathbb{N}}_{0},0\leq k\leq 2^{j}-1\}

is, in fact, a basis for C0([0,1]):={vC([0,1]):v(0)=v(1)=0}C_{0}([0,1]):=\{v\in C([0,1]):v(0)=v(1)=0\}, the so-called Schauder basis.

There holds

jk=02j12js|ψjk(t)|<,t[0,1],\sum_{j\in{\mathbb{N}}}\sum_{k=0}^{2^{j}-1}2^{js}|\psi_{jk}(t)|<\infty,\quad t\in[0,1]\;,

for any 0s<1/20\leq s<1/2. The functions ψjk\psi_{jk} are localized in the sense that |supp(ψjk)|=2j|{\rm supp}(\psi_{jk})|=2^{-j} for k=0,1,,2j1.k=0,1,\ldots,2^{j}-1.

Further constructions of such multiresolution representations of GRFs with either Riesz basis or frame properties are available on polytopal domains MdM\subset{\mathbb{R}}^{d}, (e.g. [12], for a needlet multiresolution analysis on the 22-sphere =𝕊2\mathcal{M}=\mathbb{S}^{2} embedded in 3{\mathbb{R}}^{3}, where μ\mu in Section 2.5.2 can be chosen as the surface measure see, also, for representation systems by so-called spherical needlets [92], [13]).

We also mention [5] for optimal approximation rates of truncated wavelet series approximations of fractional Brownian random fields, and to [79] for corresponding spectral representations.

Multiresolution constructions are also available on data-graphs MM (see, e.g., [41] and the references there).

2.5.4 Periodic continuation of a stationary GRF

Let (Z𝒙)𝒙D(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in{D}} be a GRF indexed by Dd{D}\subset{\mathbb{R}}^{d}, where D{{D}} is a bounded domain. We aim for representations of the general form

Z𝒙=jϕj(𝒙)yj,Z_{{\boldsymbol{x}}}=\sum_{j\in{\mathbb{N}}}\phi_{j}({\boldsymbol{x}})y_{j}, (2.9)

where the yjy_{j} are i.i.d. 𝒩(0,1){\mathcal{N}}(0,1) RVs and the (ϕj)j(\phi_{j})_{j\in\mathbb{N}} are a given sequence of functions defined on D{{D}}. One natural choice of ϕj\phi_{j} is ϕj=λjψj\phi_{j}=\sqrt{\lambda_{j}}\psi_{j}, where ψj\psi_{j} and are the eigen-functions and λj\lambda_{j} eigenvalues of the covariance operator. However, Karhunen-Loève eigenfunctions on D{{D}} are typically not explicitly known and globally supported in the physical domain D{{D}}. One of the strategies for deriving better representations over D{{D}} is to view it as the restriction to D{{D}} of a periodic Gaussian process Z𝒙extZ_{\boldsymbol{x}}^{{\rm ext}} defined on a suitable larger torus 𝕋d{\mathbb{T}}^{d}.

Since D{{D}} is bounded, without loss of generality, we may the physical domain D{{D}} to be contained in the box [12,12]d[-\frac{1}{2},\frac{1}{2}]^{d}. We wish to construct a periodic process Z𝒙extZ_{\boldsymbol{x}}^{{\rm ext}} on the torus 𝕋d{\mathbb{T}}^{d} where 𝕋=[,]{\mathbb{T}}=[-\ell,\ell] whose restriction of Z𝒙extZ_{\boldsymbol{x}}^{{\rm ext}} on D{{D}} is such that Z𝒙ext|D=Z𝒙Z_{\boldsymbol{x}}^{{\rm ext}}|_{{D}}=Z_{{\boldsymbol{x}}}. As a consequence, any representation

Z𝒙ext=jyjϕ~jZ_{\boldsymbol{x}}^{{\rm ext}}=\sum_{j\in{\mathbb{N}}}y_{j}\tilde{\phi}_{j}

yields a representation (2.9) where ϕj=ϕ~j|D\phi_{j}=\tilde{\phi}_{j}|_{{{D}}}.

Assume that (Z𝒙)𝒙D(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in{D}} is a restriction of a real-valued, stationary and centered GRF (Z𝒙)𝒙d(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in\mathbb{R}^{d}} on d\mathbb{R}^{d} whose covariance is given in the form

𝔼[Z𝒙Z𝒙]=ρ(𝒙𝒙),𝒙,𝒙d,\mathbb{E}[Z_{{\boldsymbol{x}}}Z_{{\boldsymbol{x}}^{\prime}}]=\rho({\boldsymbol{x}}-{\boldsymbol{x}}^{\prime}),\quad{\boldsymbol{x}},{\boldsymbol{x}}^{\prime}\in\mathbb{R}^{d}, (2.10)

where ρ\rho is a real-valued, even function and its Fourier transform is a non-negative function. The extension is feasible provided that we can find an even and 𝕋d{\mathbb{T}}^{d}-periodic function ρext\rho^{{\rm ext}} which agree with ρ\rho over [1,1]d[-1,1]^{d} such that the Fourier coefficients

c𝒏(ρext)=𝕋dρext(𝝃)exp(iπ(𝒏,𝝃))d𝝃,𝒏dc_{\boldsymbol{n}}(\rho^{{\rm ext}})=\int_{{\mathbb{T}}^{d}}\rho^{{\rm ext}}({\boldsymbol{\xi}})\exp\Big{(}-{\rm i}\frac{\pi}{\ell}({\boldsymbol{n}},{\boldsymbol{\xi}})\Big{)}\,\mathrm{d}{\boldsymbol{\xi}},\qquad{\boldsymbol{n}}\in{\mathbb{Z}}^{d}

are non-negative.

A natural way of constructing the function ρext\rho^{{\rm ext}} is by truncation and periodization. First one chooses a sufficiently smooth and even cutoff function φκ\varphi_{\kappa} such that φκ|[1,1]d=1\varphi_{\kappa}|_{[-1,1]^{d}}=1 and φκ(𝒙)=0\varphi_{\kappa}({\boldsymbol{x}})=0 for 𝒙[κ,κ]d{\boldsymbol{x}}\not\in[-\kappa,\kappa]^{d} where κ=21\kappa=2\ell-1. Then ρext\rho^{{\rm ext}} is defined as the periodization of the truncation ρφκ\rho\varphi_{\kappa}, i.e.,

ρext(𝝃)=𝒏d(ρφκ)(𝝃+2𝒏).\rho^{{\rm ext}}({\boldsymbol{\xi}})=\sum_{{\boldsymbol{n}}\in{\mathbb{Z}}^{d}}(\rho\varphi_{\kappa})({\boldsymbol{\xi}}+2\ell{\boldsymbol{n}}).

It is easily seen that ρext\rho^{{\rm ext}} agrees with ρ\rho over [1,1]d[-1,1]^{d} and

c𝒏(ρext)=ρφκ^(π𝒏).c_{\boldsymbol{n}}(\rho^{{\rm ext}})=\widehat{\rho\varphi_{\kappa}}\Big{(}\frac{\pi}{\ell}{\boldsymbol{n}}\Big{)}.

Therefore c𝒏(ρext)c_{\boldsymbol{n}}(\rho^{{\rm ext}}) is non-negative if we can prove that ρφκ^(𝝃)0\widehat{\rho\varphi_{\kappa}}({\boldsymbol{\xi}})\geq 0 for 𝝃d{\boldsymbol{\xi}}\in{\mathbb{R}}^{d}. The following result is shown in [12].

Theorem 2.26.

Let ρ\rho be an even function on d\mathbb{R}^{d} such that

c(1+|𝝃|2)sρ^(𝝃)C(1+|𝝃|2)r,𝝃dc(1+|{\boldsymbol{\xi}}|^{2})^{-s}\leq\hat{\rho}({\boldsymbol{\xi}})\leq C(1+|{\boldsymbol{\xi}}|^{2})^{-r},\qquad{\boldsymbol{\xi}}\in\mathbb{R}^{d} (2.11)

for some srd/2s\geq r\geq d/2 and 0<cC0<c\leq C and

limR+|𝒙|>R|αρ(𝒙)|d𝒙=0,|𝜶|2s.\lim\limits_{R\to+\infty}\int_{|{\boldsymbol{x}}|>R}|\partial^{\alpha}\rho({\boldsymbol{x}})|\,\mathrm{d}{\boldsymbol{x}}=0,\qquad|{\boldsymbol{\alpha}}|\leq 2\lceil s\rceil.

Then for κ\kappa sufficiently large, there exists φκ\varphi_{\kappa} satisfying φκ|[1,1]d=1\varphi_{\kappa}|_{[-1,1]^{d}}=1 and φκ(𝐱)=0\varphi_{\kappa}({\boldsymbol{x}})=0 for 𝐱[κ,κ]d{\boldsymbol{x}}\not\in[-\kappa,\kappa]^{d} such that

0<ρφκ^(𝝃)C(1+|𝝃|2)r,𝝃d.0<\widehat{\rho\varphi_{\kappa}}({\boldsymbol{\xi}})\leq C(1+|{\boldsymbol{\xi}}|^{2})^{-r},\qquad{\boldsymbol{\xi}}\in\mathbb{R}^{d}.

The assertion in Theorem 2.11 implies that

0<c𝒏(ρext)C(1+|𝒏|2)r,𝒏d.0<c_{\boldsymbol{n}}(\rho^{{\rm ext}})\leq C(1+|{\boldsymbol{n}}|^{2})^{-r},\qquad{\boldsymbol{n}}\in{\mathbb{Z}}^{d}.

In the following we present an explicit construction of the function φκ\varphi_{\kappa} for GRFs with Matérn covariance

ρλ,ν(𝒙):=21νΓ(ν)(2ν|𝒙|λ)νKν(2ν|𝒙|λ),\rho_{\lambda,\nu}({\boldsymbol{x}}):=\frac{2^{1-\nu}}{\Gamma(\nu)}\bigg{(}\frac{\sqrt{2\nu}|{\boldsymbol{x}}|}{\lambda}\bigg{)}^{\nu}K_{\nu}\bigg{(}\frac{\sqrt{2\nu}|{\boldsymbol{x}}|}{\lambda}\bigg{)}\,,

where λ>0\lambda>0, ν>0\nu>0 and KνK_{\nu} is the modified Bessel functions of the second kind. Note that the Matérn covariances satisfy the assumption (2.11) with s=r=ν+d/2s=r=\nu+d/2.

Let P:=2ν+d2+1P:=2\lceil\nu+\frac{d}{2}\rceil+1 and NPN_{P} be the cardinal B-spline function with nodes {P,,1,0}\{-P,\ldots,-1,0\}. For κ>0\kappa>0 we define the even function φCP1()\varphi\in C^{P-1}(\mathbb{R}) by

φ(t)={1if|t|κ/22Pκt+κ/2NP(2Pκξ)dξiftκ/2.\varphi(t)=\begin{cases}1&\text{if}\ \ |t|\leq\kappa/2\\[3.0pt] \displaystyle\frac{2P}{\kappa}\int_{-\infty}^{t+\kappa/2}N_{P}\biggl{(}\frac{2P}{\kappa}\xi\biggr{)}\,\mathrm{d}\xi&\text{if}\ \ t\leq-\kappa/2\,.\end{cases}

It is easy to see that φ(t)=0\varphi(t)=0 if |t|κ|t|\geq\kappa. We now define

φκ(𝒙):=φ(|𝒙|).\varphi_{\kappa}({\boldsymbol{x}}):=\varphi(|{\boldsymbol{x}}|).

With this choice of φκ\varphi_{\kappa}, we have ρext=ρλ,ν\rho^{\rm ext}=\rho_{\lambda,\nu} on [1,1]d[-1,1]^{d} provided that κ+d2.\ell\geq\frac{\kappa+\sqrt{d}}{2}. The required size of κ\kappa is given in the following theorem, see [14, Theorem 10].

Theorem 2.27.

For φκ\varphi_{\kappa} as defined above, there exist constants C1,C2>0C_{1},C_{2}>0 such that for any 0<λ,ν<0<\lambda,\nu<\infty, we have ρλ,νφκ^>0\widehat{\rho_{\lambda,\nu}\varphi_{\kappa}}>0 provided that κ>1\kappa>1 and

κλC1+C2max{ν12(1+|lnν|),ν12}.\frac{\kappa}{\lambda}\geq C_{1}+C_{2}\max\Big{\{}\nu^{\frac{1}{2}}(1+|{\ln\nu}|),\nu^{-\frac{1}{2}}\Big{\}}.
Remark 2.28.

The periodic random field Z𝒙extZ_{{\boldsymbol{x}}}^{\mathrm{{\rm ext}}} on 𝕋d\mathbb{T}^{d} provides a tool for deriving series expansions of the original random field. In contrast to the Karhunen-Loève eigenfunctions on DD, which are typically not explicitly known, the corresponding eigenfunctions ψjext\psi_{j}^{{\rm ext}} of the periodic covariance are explicitly known trigonometric functions and one has the following Karhunen-Loève expansion for the periodized random field:

Z𝒙ext=jyjλjextψjext,yj𝒩(0,1) i.i.d.,Z_{\boldsymbol{x}}^{\mathrm{ext}}=\sum_{j\in{\mathbb{N}}}y_{j}\sqrt{\lambda^{\mathrm{{\rm ext}}}_{j}}\,\psi_{j}^{{\rm ext}},\quad y_{j}\sim\mathcal{N}(0,1)\ \text{ i.i.d.,}

with λjext\lambda^{\mathrm{ext}}_{j} denoting the eigenvalues of the periodized covariance and the ψjext\psi_{j}^{{\rm ext}} are normalized in L2(𝕋d)L^{2}(\mathbb{T}^{d}). Restricting this expansion back to DD, one obtains an exact expansion of the original random field on DD

Z𝒙=jyjλjextψjext|D,yj𝒩(0,1) i.i.d.,Z_{\boldsymbol{x}}=\sum_{j\in{\mathbb{N}}}y_{j}\sqrt{\lambda^{\mathrm{{\rm ext}}}_{j}}\,\psi_{j}^{{\rm ext}}|_{{D}},\quad y_{j}\sim\mathcal{N}(0,1)\ \text{ i.i.d.,} (2.12)

This provides an alternative to the standard KL expansion of Z𝒙Z_{\boldsymbol{x}} in terms of eigenvalues λj\lambda_{j} and eigenfunctions ψj\psi_{j} normalized in L2(D)L^{2}({{D}}). The main difference is that the functions ψjext|D\psi^{\mathrm{ext}}_{j}\big{|}_{{D}} in (2.12) are not L2(D)L^{2}({{D}})-orthogonal. However, these functions are given explicitly, and thus no approximate computation of eigenfunctions is required.

The KL expansion of Z𝒙extZ_{\boldsymbol{x}}^{\rm ext} also enables the construction of alternative expansions of Z𝒙Z_{\boldsymbol{x}} of the basic form (2.12), but with the spatial functions having additional properties. In [12], wavelet-type representations

Z𝒙ext=,ky,kψ,k,y,k𝒩(0,1) i.i.d.,Z_{\boldsymbol{x}}^{{\rm ext}}=\sum_{\ell,k}y_{\ell,k}\psi_{\ell,k},\quad y_{\ell,k}\sim\mathcal{N}(0,1)\ \text{ i.i.d.,}

are constructed where the functions ψ,k\psi_{\ell,k} have the same multilevel-type localisation as the Meyer wavelets. This feature yields improved convergence estimates for tensor Hermite polynomial approximations of solutions of random diffusion equations with log-Gaussian coefficients .

2.5.5 Sampling stationary GRFs

The simulation of GRFs with specified covariance is a fundamental task in computational statistics with a wide range of applications. In this section we present an efficient methods for sampling such fields. Consider a GRF (Z𝒙)𝒙D(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in{{D}}} where D{{D}} is contained in [1/2,1/2]d[-1/2,1/2]^{d}. Assume that (Z𝒙)𝒙D(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in{D}} is a restriction of a real-valued, stationary and centered GRF (Z𝒙)𝒙d(Z_{\boldsymbol{x}})_{{\boldsymbol{x}}\in\mathbb{R}^{d}} on d\mathbb{R}^{d} with covariance given in (2.10). Let mm\in{\mathbb{N}} and 𝒙1,,𝒙M{\boldsymbol{x}}_{1},\ldots,{\boldsymbol{x}}_{M} be M=(m+1)dM=(m+1)^{d} uniform grid points on [1/2,1/2]d[-1/2,1/2]^{d} with grid spacing h=1/mh=1/m. We wish to obtain samples of the Gaussian RV

𝒁=(Z𝒙1,,Z𝒙M){\boldsymbol{Z}}=(Z_{{\boldsymbol{x}}_{1}},\ldots,Z_{{\boldsymbol{x}}_{M}})

with covariance matrix

𝚺=[Σi,j]i,j=1M,Σi,j=ρ(𝒙i𝒙j),i,j=1,,M.\displaystyle\boldsymbol{\Sigma}=[\Sigma_{i,j}]_{i,j=1}^{M},\qquad\Sigma_{i,j}=\rho({\boldsymbol{x}}_{i}-{\boldsymbol{x}}_{j}),\quad i,j=1,\ldots,M. (2.13)

Since 𝚺\boldsymbol{\Sigma} is symmetric positive semidefinite, this can in principle be done by performing the Cholesky factorisation 𝚺=𝑭𝑭\boldsymbol{\Sigma}=\boldsymbol{F}\boldsymbol{F}^{\top} with 𝑭=𝚺1/2\boldsymbol{F}=\boldsymbol{\Sigma}^{1/2}, from which the desired samples are provided by the product 𝑭𝒀\boldsymbol{F}{\boldsymbol{Y}} where 𝒀𝒩(0,𝑰){\boldsymbol{Y}}\sim\mathcal{N}(0,{\boldsymbol{I}}). However, since Σ\Sigma is large and dense when mm is large, this factorisation is prohibitively expensive. Since the covariance matrix 𝚺\boldsymbol{\Sigma} is a nested block Toeplitz matrix under appropriate ordering, an efficient approach is to extend 𝚺\boldsymbol{\Sigma} to a appropriate larger nested block circulant matrix whose spectral decomposition can be rapidly computed using FFT.

For any 1\ell\geq 1 we construct a 22\ell-periodic extension of ρ\rho as follows

ρext(𝒙)=𝒏d(ρχ(,]d)(𝒙+2𝒏),𝒙d.\displaystyle\rho^{\rm ext}({\boldsymbol{x}})=\sum_{{\boldsymbol{n}}\in{\mathbb{Z}}^{d}}\bigl{(}\rho\chi_{(-\ell,\ell]^{d}}\bigr{)}({\boldsymbol{x}}+2\ell{\boldsymbol{n}}),\quad{\boldsymbol{x}}\in\mathbb{R}^{d}\,.

Clearly, ρext\rho^{\rm ext} is 22\ell-periodic and ρext=ρ\rho^{\rm ext}=\rho on [1,1]d[-1,1]^{d}. Denote 𝝃1,,𝝃s{\boldsymbol{\xi}}_{1},\ldots,{\boldsymbol{\xi}}_{s}, s=(2/h)ds=(2\ell/h)^{d}, the uniform grid points on [,]d[-\ell,\ell]^{d} with grid space hh. Let 𝒁ext=(Z𝝃1,,Z𝝃s){\boldsymbol{Z}}^{\rm ext}=(Z_{{\boldsymbol{\xi}}_{1}},\ldots,Z_{{\boldsymbol{\xi}}_{s}}) be the extended GRV with covariance matrix 𝚺ext\boldsymbol{\Sigma}^{\rm ext} whose entries is given by formula (2.13), with ρ\rho replaced by ρext\rho^{\rm ext} and 𝒙i{\boldsymbol{x}}_{i} by 𝝃i{\boldsymbol{\xi}}_{i}. Hence 𝚺\boldsymbol{\Sigma} is embedded into the nested circulant matrix 𝚺ext\boldsymbol{\Sigma}^{\rm ext} which can be diagonalized using FFT (with log-linear complexity) to provide the spectral decomposition

𝚺ext=𝑸ext𝚲ext(𝑸ext),\boldsymbol{\Sigma}^{\rm ext}=\boldsymbol{Q}^{\rm ext}\boldsymbol{\Lambda}^{\rm ext}(\boldsymbol{Q}^{\rm ext})^{\top},

with Λext\Lambda^{\rm ext} diagonal and containing the eigenvalues λjext\lambda^{\rm ext}_{j} of 𝚺ext\boldsymbol{\Sigma}^{\rm ext} and 𝑸ext\boldsymbol{Q}^{\rm ext} being a Fourier matrix. Provided that these eigenvalues are non-negative, the samples of the grid values of ZZ can be drawn as follows. First we draw a random vector (yj)j=1,,s(y_{j})_{j=1,\ldots,s} with yj𝒩(0,1)y_{j}\sim\mathcal{N}(0,1) i.i.d., then compute

𝒁ext=j=1syjλjext𝒒j{\boldsymbol{Z}}^{\text{ext}}=\sum_{j=1}^{s}y_{j}\sqrt{\lambda^{\rm ext}_{j}}\,{\boldsymbol{q}}_{j}

using the FFT, with 𝒒j{\boldsymbol{q}}_{j} the columns of 𝑸ext\boldsymbol{Q}^{\rm ext}. Finally, a sample of 𝒁{\boldsymbol{Z}} is obtained by extracting from 𝒁ext{\boldsymbol{Z}}^{\text{ext}} the entries corresponding to the original grid points.

The above mentioned process is feasible, provided that 𝚺\boldsymbol{\Sigma} is positive semidefinite. The following theorem characterizes the condition on \ell for GRF with Matérn covariance such that 𝚺ext\boldsymbol{\Sigma}^{\rm ext} is positive semidefinite, see [60].

Theorem 2.29.

Let 1/2ν<1/2\leq\nu<\infty, λ1\lambda\leq 1, and h/λe1h/\lambda\leq e^{-1}. Then there exist C1,C2>0C_{1},C_{2}>0 which may depend on dd but are independent of ,h,λ,ν\ell,h,\lambda,\nu, such that Σext\Sigma^{\rm ext} is positive definite if

λC1+C2ν12log(max{λ/h,ν12}).\frac{\ell}{\lambda}\ \geq\ C_{1}\ +\ C_{2}\,\nu^{\frac{1}{2}}\,\log\bigl{(}\max\big{\{}{\lambda}/{h},\,\nu^{\frac{1}{2}}\big{\}}\bigr{)}\,.
Remark 2.30.

For GRF with Matérn covariances, it is well-known (see, e.g. [59, Corollary 5], [10, eq.(64)]) that the exact KL eigenvalues λj\lambda_{j} of Z𝒙Z_{\boldsymbol{x}} in L2(D)L^{2}({{D}}) decay with the rate λjCj(1+2ν/d)\lambda_{j}\ \leq\ Cj^{-(1+2\nu/d)}. It has been proved recently in [14] that the eigenvalue λjext\lambda^{\rm ext}_{j} maintain this rate of decay up to a factor of order 𝒪(|logh|ν)\mathcal{O}(|\!\log h|^{\nu}).

2.6 Finite element discretization

The approximation results and algorithms to be developed in the present text involve, besides the Wiener-Hermite PC expansions with respect to Gaussian co-ordinates 𝒚{\boldsymbol{y}}\in\mathbb{R}^{\infty}, also certain numerical approximations in the physical domain D{D}. Due to their wide use in the numerical solution of elliptic and parabolic PDEs, we opt for considering standard, primal Lagrangian finite element discretizations. We confine the presentation and analysis to Lipschitz polytopal domains Dd{D}\subset\mathbb{R}^{d} with principal interest in d=2d=2 (D{D} is a polygon with straight sides) and d=3d=3 (D{D} is a polyhedron with plane faces). We confine the presentation to so-called primal FE discretizations in D{D} but hasten to add that with minor extra mathematical effort, similar results could be developed also for so-called mixed, or dual FE discretizations (see,e.g., [20] and the references there).

In presenting (known) results on finite element method (FEM for short) convergence rates, we consider separately FEM in polytopal domains Dd{D}\subset{\mathbb{R}}^{d}, d=1,2,3d=1,2,3, and FEM on smooth dd-surfaces Γd+1\Gamma\subset{\mathbb{R}}^{d+1}, d=1,2d=1,2. See [22, 49]

2.6.1 Function spaces

For a bounded domain Dd{D}\subset{\mathbb{R}}^{d}, the usual Sobolev function spaces of integer order s0s\in{\mathbb{N}}_{0} and integrability q[1,]q\in[1,\infty] are denoted by Wqs(D)W^{s}_{q}({D}) with the understanding that Lq(D)=Wq0(D)L^{q}({D})=W^{0}_{q}({D}). The norm of vWqs(D)v\in W^{s}_{q}({D}) is defined by

vWqs:=𝜶+d:|𝜶|sD𝜶vLq.\|v\|_{W^{s}_{q}}:=\ \sum_{{\boldsymbol{\alpha}}\in{\mathbb{Z}}^{d}_{+}:|{\boldsymbol{\alpha}}|\leq s}\|D^{\boldsymbol{\alpha}}v\|_{L^{q}}.

Here D𝜶D^{\boldsymbol{\alpha}} denotes the partial weak derivative of order 𝜶{\boldsymbol{\alpha}}. We refer to any standard text such as [2] for basic properties of these spaces. Hilbertian Sobolev spaces are given for s0s\in{\mathbb{N}}_{0} by Hs(D)=W2s(D)H^{s}({D})=W^{s}_{2}({D}), with the usual understanding that L2(D)=H0(D)L^{2}({D})=H^{0}({D}).

For ss\in{\mathbb{N}}, we call a CsC^{s}-domain Dd{D}\subset{\mathbb{R}}^{d} a bounded domain whose boundary D\partial{D} is locally parameterized in a finite number of co-ordinate systems as a graph of a CsC^{s} function. In a similar way, we shall call Dd{D}\subset{\mathbb{R}}^{d} a Lipschitz domain, when D\partial{D} is, locally, the graph of a Lipschitz function. We refer to [2, 55] and the references there or to [61].

We call polygonal domain a domain D2{D}\subset{\mathbb{R}}^{2} that is a polygon with Lipschitz boundary D\partial{D} (which precludes cusps and slits) and with a finite number of straight sides.

Let D2{D}\subset{\mathbb{R}}^{2} denote an open bounded polygonal domain. We introduce in D{D} a nonnegative function rD:D+r_{D}:{D}\to{\mathbb{R}}_{+} which is smooth in D{D}, and which coincides for 𝒙{\boldsymbol{x}} in a vicinity of each corner 𝒄D{\boldsymbol{c}}\in\partial{D} with the Euclidean distance |𝒙𝒄||{\boldsymbol{x}}-{\boldsymbol{c}}|.

To state elliptic regularity shifts in D{D}, we require certain corner-weighted Sobolev spaces. We require these only for integrability q=2q=2 and for q=q=\infty.

For s0s\in{\mathbb{N}}_{0} and ϰ\varkappa\in{\mathbb{R}} we define

𝒦ϰs(D):={u:D:rD|𝜶|ϰD𝜶uL2(D),|𝜶|s}{\mathcal{K}}^{s}_{\varkappa}({D}):=\big{\{}u:{D}\to{\mathbb{C}}:\ r_{D}^{|{\boldsymbol{\alpha}}|-\varkappa}D^{\boldsymbol{\alpha}}u\in L^{2}({D}),|{\boldsymbol{\alpha}}|\leq s\big{\}}

and

𝒲s(D):={u:D:rD|𝜶|D𝜶uL(D),|𝜶|s}.{\mathcal{W}}^{s}_{\infty}({D}):=\big{\{}u:{D}\to{\mathbb{C}}:\ r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}u\in L^{\infty}({D}),\ |{\boldsymbol{\alpha}}|\leq s\big{\}}.

Here, for 𝜶02{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{2} and as before D𝜶D^{\boldsymbol{\alpha}} denotes the partial weak derivative of order 𝜶{\boldsymbol{\alpha}}.

The corner-weighted norms in these spaces are given by

u𝒦ϰs:=|𝜶|srD|𝜶|ϰD𝜶uL2andu𝒲s:=|𝜶|srD|𝜶|D𝜶uL.\|u\|_{{\mathcal{K}}^{s}_{\varkappa}}:=\sum_{|{\boldsymbol{\alpha}}|\leq s}\|r_{D}^{|{\boldsymbol{\alpha}}|-\varkappa}D^{\boldsymbol{\alpha}}u\|_{L^{2}}\qquad\text{and}\qquad\|u\|_{{\mathcal{W}}^{s}_{\infty}}:=\sum_{|{\boldsymbol{\alpha}}|\leq s}\|r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}u\|_{L^{\infty}}\,.

The function spaces 𝒦ϰs(D){\mathcal{K}}^{s}_{\varkappa}({D}) and 𝒲s(D){\mathcal{W}}^{s}_{\infty}({D}) endowed with these norms are Banach spaces, and 𝒦ϰs(D){\mathcal{K}}^{s}_{\varkappa}({D}) are separable Hilbert spaces. These corner-weighted Sobolev spaces are called Kondrat’ev spaces.

An embedding of these spaces is H01(D)𝒦01(D)H^{1}_{0}({D})\hookrightarrow{\mathcal{K}}^{1}_{0}({D}). This follows from the existence of a constant c(D)>0c({D})>0 such that for every 𝒙D{\boldsymbol{x}}\in{D} holds rD(𝒙)c(D)dist(𝒙,D)r_{D}({\boldsymbol{x}})\geq c({D}){\rm dist}({\boldsymbol{x}},\partial{D}).

2.6.2 Finite element interpolation

In this section, we review some results on FE approximations in polygonal domains D{D} on locally refined triangulations 𝒯\mathcal{T} in D{D}. These results are in principle known for the standard Sobolev spaces Hs(D)H^{s}({D}) and available in the standard texts [23, 33]. For spaces with corner weights in polygonal domains D2{D}\subset{\mathbb{R}}^{2}, such as 𝒦ϰs{\mathcal{K}}^{s}_{\varkappa} and 𝒲s{\mathcal{W}}^{s}_{\infty}, however, which arise in the regularity of the Wiener-Hermite PC expansion coefficient functions for elliptic PDEs in corner domains in Section 3.8 ahead, we provide references to corresponding FE approximation rate bounds.

The corresponding FE spaces involve suitable mesh refinement to compensate for the reduced regularity caused by corner and edge singularities which occur in solutions to elliptic and parabolic boundary value problems in these domains.

We define the FE spaces in a polygonal domain D2{D}\subset{\mathbb{R}}^{2} (see [23, 33] for details). Let 𝒯\mathcal{T} denote a regular triangulation of D¯\overline{D}, i.e., a partition of D¯\overline{D} into a finite number N(𝒯)N(\mathcal{T}) of closed, nondegenerate triangles T𝒯T\in\mathcal{T} (i.e., |T|>0|T|>0) such that for any two T,T𝒯T,T^{\prime}\in\mathcal{T}, the intersection TTT\cap T^{\prime} is either empty, a vertex or an entire edge. We denote the meshwidth of 𝒯\mathcal{T} as

h(𝒯):=max{h(T):T𝒯},whereh(T):=diam(T).h(\mathcal{T}):=\max\{h(T):T\in\mathcal{T}\},\;\;\mbox{where}\;\;h(T):=\mbox{diam}(T)\;.

For T𝒯T\in\mathcal{T}, denote ρ(T)\rho(T) the diameter of the largest circle that can be inscribed into TT. We say 𝒯\mathcal{T} is κ\kappa shape-regular, if

T𝒯:h(T)ρ(T)κ.\forall T\in\mathcal{T}:\;\;\frac{h(T)}{\rho(T)}\leq\kappa\;.

A sequence 𝔗:=(𝒯n)n\mathfrak{T}:=(\mathcal{T}_{n})_{n\in{\mathbb{N}}} is κ\kappa shape-regular if each 𝒯𝔗\mathcal{T}\in\mathfrak{T} is κ\kappa shape-regular, with one common constant κ>1\kappa>1 for all 𝒯𝔗\mathcal{T}\in\mathfrak{T}.

In a polygon D{D}, with a regular, simplicial triangulation 𝒯\mathcal{T}, and for a polynomial degree mm\in{\mathbb{N}}, the Lagrangian FE space Sm(D,𝒯)S^{m}({D},\mathcal{T}) of continuous, piecewise polynomial functions of degree mm on 𝒯\mathcal{T} is defined as

Sm(D,𝒯)={vH1(D):T𝒯:v|Tm}.S^{m}({D},\mathcal{T})=\{v\in H^{1}({D}):\forall T\in\mathcal{T}:v|_{T}\in\mathbb{P}_{m}\}\;.

Here, m:=span{𝒙𝜶:|𝜶|m}\mathbb{P}_{m}:={\rm span}\{{\boldsymbol{x}}^{\boldsymbol{\alpha}}:|{\boldsymbol{\alpha}}|\leq m\} denotes the space of polynomials of 𝒙2{\boldsymbol{x}}\in{\mathbb{R}}^{2} of total degree at most mm. We also define S0m(D,𝒯):=Sm(D,𝒯)H01(D)S^{m}_{0}({D},\mathcal{T}):=S^{m}({D},\mathcal{T})\cap H^{1}_{0}({D}).

The main result on FE approximation rates in a polygon D2{D}\subset{\mathbb{R}}^{2} in corner-weighted spaces 𝒦κs(D){\mathcal{K}}^{s}_{\kappa}({D}) reads as follows.

Proposition 2.31.

Consider a bounded polygonal domain D2{D}\subset{\mathbb{R}}^{2}. Then, for every polynomial degree mm\in{\mathbb{N}}, there exists a sequence (𝒯n)n(\mathcal{T}_{n})_{n\in{\mathbb{N}}} of κ\kappa shape-regular, simplicial triangulations of D{D} such that for every u(H01𝒦λm+1)(D)u\in(H^{1}_{0}\cap{\mathcal{K}}^{m+1}_{\lambda})({D}) for some λ>0\lambda>0, the FE interpolation error converges at rate mm. More precisely, there exists a constant C(D,κ,λ,m)>0C({D},\kappa,\lambda,m)>0 such that for all 𝒯(𝒯n)n\mathcal{T}\in(\mathcal{T}_{n})_{n\in{\mathbb{N}}} and for all u(H01𝒦λm+1)(D)u\in(H^{1}_{0}\cap{\mathcal{K}}^{m+1}_{\lambda})({D}) holds

uI𝒯muH1Ch(𝒯)mu𝒦λm+1.\|u-I^{m}_{\mathcal{T}}u\|_{H^{1}}\leq Ch(\mathcal{T})^{m}\|u\|_{{\mathcal{K}}^{m+1}_{\lambda}}\;.

Equivalently, in terms of the number n:=#(𝒯)n:=\#(\mathcal{T}) of triangles, there holds

uI𝒯muH1Cnm/2u𝒦λm+1.\|u-I^{m}_{\mathcal{T}}u\|_{H^{1}}\leq Cn^{-m/2}\|u\|_{{\mathcal{K}}^{m+1}_{\lambda}}\;. (2.14)

Here, I𝒯m:C0(D¯)Sm(D,𝒯)I^{m}_{\mathcal{T}}:C^{0}(\overline{{D}})\to S^{m}({D},\mathcal{T}) denotes the nodal, Lagrangian interpolant. The constant C>0C>0 depends on mm, D{D} and the shape regularity of 𝒯\mathcal{T}, but is independent of uu.

For a proof of this proposition, we refer, for example, to [25, Theorems 4.2, 4.4].

We remark that due to 𝒦λ2(D)C0(D¯){\mathcal{K}}_{\lambda}^{2}({D})\subset C^{0}(\overline{{D}}), the nodal interpolant I𝒯mI^{m}_{\mathcal{T}} in (2.14) is well-defined. We also remark that the triangulations 𝒯n\mathcal{T}_{n} need not necessarily be nested (the constructions in [6, 25] do not provide nestedness; for a bisection tree construction of (𝒯n)n(\mathcal{T}_{n})_{n\in{\mathbb{N}}} which are nested, such as typically produced by adaptive FE algorithms, with the error bounds (2.14), we refer to [54].

For similar results in polyhedral domains in space dimension d=3d=3, we refer to [27, 26, 86] and to the references there.

3 Elliptic divergence-form PDEs with log-Gaussian coefficient

We present a model second order linear divergence-form PDE with log-Gaussian input data. We review known results on its well-posedness, and Lipschitz continuous dependence on the input data. Particular attention is placed on regularity results in polygonal domains D2{D}\subset{\mathbb{R}}^{2}. Here, solutions belong to Kondrat’ev spaces. We discuss regularity results for parametric coefficients, and establish in particular parametric holomorphy results for the coefficient-to-solution maps.

The outline of this section is as follows. In Section 3.1, we present the strong and variational forms of the PDE, its well-posedness and the continuity of the data-to-solution map in appropriate spaces. Importantly, we do not aim at the most general setting, but to ease notation and for simplicity of presentation we address a rather simple, particular case: in a bounded domain D{D} in Euclidean space d{\mathbb{R}}^{d}. All the ensuing derivations will directly generalize to linear second order elliptic systems. A stronger Lipschitz continuous dependence on data result is stated in Section 3.2. Higher regularity and fractional regularity of the solution provided correspondingly by higher regularity of data are discussed in Section 3.3.

Sections 3.4 and 3.5 describe uncertainty modelling by placing GMs on sets of admissible, countably parametric input data, i.e., formalizing mathematically aleatoric uncertainty in input data. Here, the Gaussian series introduced in Section 2.5 will be seen to take a key role in converting operator equations with GRF inputs to infinitely-parametric, deterministic operator equations. The Lipschitz continuous dependence of the solutions on input data from function spaces will imply strong measurability of corresponding random solutions, and render well-defined the uncertainty propagation, i.e., the push-forward of the GM on the input data.

In Sections 3.63.8, we connect quantified holomorphy of the parametric, deterministic solution manifold {u(𝒚):𝒚}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in{\mathbb{R}}^{\infty}\} with sparsity of the coefficients (u𝝂H)𝝂(\|u_{\boldsymbol{\nu}}\|_{H})_{{\boldsymbol{\nu}}\in{\mathcal{F}}} of Wiener-Hermite PC expansion as elements of certain Sobolev spaces: We start with the case H=H01(D)H=H_{0}^{1}({D}) in Section 3.6 and subsequently discuss higher regularity H=Hs(D)H=H^{s}({D}), ss\in\mathbb{N}, in Section 3.7 and finally HH being a Kondrat’ev space on a bounded polygonal domain D2{D}\subset\mathbb{R}^{2} in Section 3.8.

3.1 Statement of the problem and well-posedness

In a bounded Lipschitz domain Dd{D}\subset{\mathbb{R}}^{d} (d=1,2or 3)(d=1,2\ \text{or}\ 3), consider the linear second order elliptic PDE in divergence-form

Pau:={div(a(𝒙)u(𝒙))τ0(u)}={f(𝒙)inD,0onD.P_{a}u:=\left\{\begin{array}[]{l}-\operatorname{div}(a({\boldsymbol{x}})\nabla u({\boldsymbol{x}}))\\ \tau_{0}(u)\end{array}\right\}=\left\{\begin{array}[]{c}f({\boldsymbol{x}})\;\;\mbox{in}\;\;{D},\\ 0\;\;\mbox{on}\;\;\partial{D}\;.\end{array}\right. (3.1)

Here, τ0:H1(D)H1/2(D)\tau_{0}:H^{1}({D})\to H^{1/2}(\partial{D}) denotes the trace map. With the notation V:=H01(D)V:=H_{0}^{1}({D}) and V=H1(D)V^{*}=H^{-1}({D}), for any fVf\in V^{*}, by the Lax-Milgram lemma the weak formulation given by

uV:Dauvd𝒙=f,vV,V,vV,u\in V:\;\;\int_{{D}}a\nabla u\cdot\nabla v\,\,\mathrm{d}{\boldsymbol{x}}=\langle f,v\rangle_{V^{*},V}\,,\qquad v\in V, (3.2)

admits a unique solution uVu\in V whenever the coefficient aa satisfies the ellipticity assumption

0<amin:=essinf𝒙Da(𝒙)amax=aL<.0<a_{\min}:=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,a({\boldsymbol{x}})\leq a_{\max}=\|a\|_{L^{\infty}}<\infty\;. (3.3)

With vV:=vL2\|v\|_{V}:=\|\nabla v\|_{L^{2}} denoting the norm of vVv\in V, there holds the a-priori estimate

uVfVamin.\|u\|_{V}\leq\frac{\|f\|_{V^{*}}}{a_{\min}}\,. (3.4)

In particular, with

L+(D):={aL(D):amin>0},L^{\infty}_{+}({D}):=\big{\{}a\in L^{\infty}({D}):a_{\min}>0\big{\}},

the data-to-solution operator

𝒮:L+(D)×VV:(a,f)u\mathcal{S}:L^{\infty}_{+}({D})\times V^{*}\to V:(a,f)\mapsto u (3.5)

is continuous.

3.2 Lipschitz continuous dependence

The continuity (3.5) of the data-to-solution map 𝒮\mathcal{S} allows to infer already strong measurability of solutions of (3.1) with respect to random coefficients aa. For purposes of stable numerical approximation, we will be interested in quantitative bounds of the effect of perturbations of the coefficient aa in (3.2) and of the source term data ff on the solution u=𝒮(a,f)u=\mathcal{S}(a,f). Mere continuity of 𝒮\mathcal{S} as a map from L+(D)×VL^{\infty}_{+}({D})\times V^{*} to V=H01(D)V=H^{1}_{0}({D}) will not be sufficient to this end. To quantify the impact of uncertainty in the coefficient aa on the solution uVu\in V, local Hölder or, preferably, Lipschitz continuity of the map 𝒮\mathcal{S} is required, at least locally, close to nominal values of the data (a,f)(a,f).

To this end, consider given a1,a2L+(D)a_{1},a_{2}\in L^{\infty}_{+}({D}), f1,f2L2(D)Vf_{1},f_{2}\in L^{2}({D})\subset V^{*} with corresponding unique solutions ui=𝒮(ai,fi)Vu_{i}=\mathcal{S}(a_{i},f_{i})\in V, i=1,2i=1,2.

Proposition 3.1.

In a bounded Lipschitz domain Dd{D}\subset\mathbb{R}^{d}, for given data bounds ra,rf(0,)r_{a},r_{f}\in(0,\infty), there exist constants cac_{a} and cfc_{f} such that for every aiL+(D)a_{i}\in L^{\infty}_{+}({D}) with log(ai)Lra\|\log(a_{i})\|_{L^{\infty}}\leq r_{a}, and for every fiL2(D)f_{i}\in L^{2}({D}) with fiL2rf\|f_{i}\|_{L^{2}}\leq r_{f}, i=1,2i=1,2, it holds

u1u2VcPa1,mina2,minf1f2L2+f1Vf2Va1,mina2,mina1a2L.\|u_{1}-u_{2}\|_{V}\leq\frac{c_{P}}{a_{1,\min}\wedge a_{2,\min}}\|f_{1}-f_{2}\|_{L^{2}}+\frac{\|f_{1}\|_{V^{*}}\vee\|f_{2}\|_{V^{*}}}{a_{1,\min}a_{2,\min}}\|a_{1}-a_{2}\|_{L^{\infty}}\;. (3.6)

Therefore

𝒮(a1,f1)𝒮(a2,f2)Vcaa1a2L+cff1f2L2,\|\mathcal{S}(a_{1},f_{1})-\mathcal{S}(a_{2},f_{2})\|_{V}\leq c_{a}\|a_{1}-a_{2}\|_{L^{\infty}}+c_{f}\|f_{1}-f_{2}\|_{L^{2}}\;, (3.7)

and

𝒮(a1,f1)𝒮(a2,f2)Vc~alog(a1)log(a2)L+cff1f2L2.\|\mathcal{S}(a_{1},f_{1})-\mathcal{S}(a_{2},f_{2})\|_{V}\leq\tilde{c}_{a}\|\log(a_{1})-\log(a_{2})\|_{L^{\infty}}+c_{f}\|f_{1}-f_{2}\|_{L^{2}}\;. (3.8)

Here, we may take cf=cPexp(ra)c_{f}=c_{P}\exp(r_{a}), ca=cPrfexp(2ra)c_{a}=c_{P}r_{f}\exp(2r_{a}) and c~a=cPrfexp(3ra)\tilde{c}_{a}=c_{P}r_{f}\exp(3r_{a}). The constant cP=c(D)>0c_{P}=c({D})>0 denotes the VL2(D)V-L^{2}({D}) Poincaré constant of D{D}.

The bounds (3.7) and (3.8) follow from the continuous dependence estimates in [15] by elementary manipulations. For a proof (in a slightly more general setting), we also refer to Section 4.3.1 ahead.

3.3 Regularity of the solution

It is well known that weak solutions uVu\in V of the linear elliptic boundary value problem (BVP for short) (3.1) admit higher regularity for more regular data (i.e., coefficient a(𝒙)a({\boldsymbol{x}}), source term f(𝒙)f({\boldsymbol{x}}) and domain D{D}). Standard references for corresponding results are [61, 55]. The proofs in these references cover general, linear elliptic PDEs, with possibly matrix-valued coefficients, and aim at sharp results on the Sobolev and Hölder regularity of solutions, in terms of corresponding regularity of coefficients, source term and boundar D\partial{D}. In order to handle the dependence of solutions on random field and parametric coefficients in a quantitative manner, we develop presently self-contained, straightforward arguments for solution regularity of (3.1).

Here is a first regularity statement, which will be used in several places subsequently. To state it, we denote by WW the normed space of all functions vVv\in V such that ΔvL2(D)\Delta v\in L^{2}({D}). The norm in WW is defined by

vW:=ΔvL2.\|v\|_{W}:=\|\Delta v\|_{L^{2}}.

The map vvWv\mapsto\|v\|_{W} is indeed a norm on WW due to the homogeneous Dirichlet boundary condition of vVv\in V: vW=0\|v\|_{W}=0 implies that vv is harmonic in D{D}, and vVv\in V implies that the trace of vv on D\partial{D} vanishes, whence v=0v=0 in D{D} by the maximum principle.

Proposition 3.2.

Consider the boundary value problem (3.1) in a bounded domain D{D} with Lipschitz boundary, and with aW1(D)a\in W^{1}_{\infty}({D}), fL2(D)f\in L^{2}({D}). Then the weak solution uVu\in V of (3.1) belongs to the space WW and there holds the a-priori estimate

uW1amin(fL2+fVaLamin)camin(1+aLamin)fL2,\|u\|_{W}\ \leq\ \frac{1}{a_{\min}}\left(\|f\|_{L^{2}}+\|f\|_{V^{*}}\frac{\|\nabla a\|_{L^{\infty}}}{a_{\min}}\right)\leq\frac{c}{a_{\min}}\left(1+\frac{\|\nabla a\|_{L^{\infty}}}{a_{\min}}\right)\|f\|_{L^{2}}\;, (3.9)

where amin=min{a(𝐱):𝐱D¯}a_{\min}=\min\{a({\boldsymbol{x}}):{\boldsymbol{x}}\in\overline{{D}}\}.

Proof.

That uVu\in V belongs to WW is verified by observing that under these assumptions, there holds

aΔu=f+auin the sense ofL2(D).-a\Delta u\ =\ f+\nabla a\cdot\nabla u\;\;\mbox{in the sense of}\;\;L^{2}({{D}})\;. (3.10)

The first bound (3.9) follows by elementary argument using (3.4), the second bound by an application of the L2(D)L^{2}({D})-VV^{*} Poincaré inequality in D{D}. ∎

Remark 3.3.

The relevance of the space WW stems from the relation to the corner-weighted Kondrat’ev spaces 𝒦κm(D){\mathcal{K}}^{m}_{\kappa}({D}) which were introduced in Section 2.6.1. When the domain D2{D}\subset{\mathbb{R}}^{2} is a polygon with straight sides, in the presently considered homogeneous Dirichlet boundary conditions on all of D\partial{D}, it holds that W𝒦κ2(D)W\subset{\mathcal{K}}^{2}_{\kappa}({D}) with continuous injection provided that |κ|<π/ω|\kappa|<\pi/\omega where 0<ω<2π0<\omega<2\pi is the largest interior opening angle at the vertices of D{D}. Membership of uu in 𝒦κ2(D){\mathcal{K}}^{2}_{\kappa}({D}) in turn implies optimal approximation rates for standard, Lagrangian FE approximations in D{D} with suitable, corner-refined triangulations in D{D}, see Proposition 2.31.

Remark 3.4.

If the physical domain D{D} is convex or of type C1,1C^{1,1}, then uWu\in W implies that u(H2H01)(D)u\in(H^{2}\cap H^{1}_{0})({{D}}) and (3.9) gives rise to an H2H^{2} a-priori estimate (see, e.g., [61, Theorem 2.2.2.3]).

The regularity in Proposition 3.2 is adequate for diffusion coefficients a(𝒙)a({\boldsymbol{x}}) which are Lipschitz continuous in D{D}, which is essentially (up to modification) W1(D)C0,1(D)W^{1}_{\infty}({D})\simeq C^{0,1}({D}). In view of our interest in admitting diffusion coefficients which are (realizations of) GRF (see Section 3.4), it is clear from Example 2.25 that relevant GRF models may exhibit mere Hölder path regularity.

The Hölder spaces Cs(D)C^{s}({D}) on Lipschitz domains D{D} can be obtained as interpolation spaces, via the so-called KK-method of function space interpolation which we briefly recapitulate (see, e.g., [107, Chapter 1.3], [18]). Two Banach spaces A0,A1A_{0},A_{1} with continuous embedding A1A0A_{1}\hookrightarrow A_{0} with respective norms Ai\|\circ\|_{A_{i}}, i=0,1i=0,1, constitute an interpolation couple. For 0<s<10<s<1, the interpolation space [A0,A1]s,q[A_{0},A_{1}]_{s,q} of smoothness order ss with fine index q[1,]q\in[1,\infty] is defined via the KK-functional: for aA0a\in A_{0}, this functional is given by

K(a,t;A0,A1):=infa1A1{aa1A0+ta1A1},t>0.K(a,t;A_{0},A_{1}):=\inf_{a_{1}\in A_{1}}\{\|a-a_{1}\|_{A_{0}}+t\|a_{1}\|_{A_{1}}\}\;,\quad t>0\;. (3.11)

For 0<s<10<s<1 the intermediate, “interpolation” space of order ss and fine index qq is denoted by [A0,A1]s,q[A_{0},A_{1}]_{s,q}. It is the set of functions aA0a\in A_{0} such that the quantity

a[A0,A1]s,q:={(0(tsK(a,t,A0,A1))qdtt)1/q,1q<,supt>0tsK(a,t,A0,A1),q=\|a\|_{[A_{0},A_{1}]_{s,q}}:=\left\{\begin{array}[]{lr}\left(\int_{0}^{\infty}(t^{-s}K(a,t,A_{0},A_{1}))^{q}\frac{\,\mathrm{d}t}{t}\right)^{1/q}\;,&1\leq q<\infty,\\[4.30554pt] \sup_{t>0}t^{-s}K(a,t,A_{0},A_{1})\;,&q=\infty\end{array}\right. (3.12)

is finite. When the AiA_{i} are Banach spaces, the sets [A0,A1]s,q[A_{0},A_{1}]_{s,q} are Banach spaces with norm given by (3.12). In particular (see, e.g., [2, Lemma 7.36]), in the bounded Lipschitz domain D{D}

Cs(D)=[L(D),W1(D)]s,,0<s<1.C^{s}({D})=[L^{\infty}({D}),W^{1}_{\infty}({D})]_{s,\infty},\quad 0<s<1\;. (3.13)

With the spaces V:=H01(D)V:=H^{1}_{0}({D}) and WVW\subset V, we define the (non-separable, non-reflexive) Banach space

Ws:=[V,W]s,,0<s<1.W^{s}:=[V,W]_{s,\infty}\;,\quad 0<s<1\;. (3.14)

Then there holds the following generalization of (3.9).

Proposition 3.5.

For a bounded Lipschitz domain Dd{D}\subset{\mathbb{R}}^{d}, d2d\geq 2, for every fL2(D)f\in L^{2}({D}) and aCs(D)a\in C^{s}({D}) for some 0<s<10<s<1 with

amin=min{a(𝒙):𝒙D¯}>0,a_{\min}=\min\{a({\boldsymbol{x}}):{\boldsymbol{x}}\in\overline{{D}}\}>0,

the solution uVu\in V of (3.1), (3.2) belongs to WsW^{s}, and there exists a constant c(s,D)c(s,{D}) such that

uWscamin(1+aCs1/samin1/s)fL2.\|u\|_{W^{s}}\leq\frac{c}{a_{\min}}\left(1+\|a\|_{C^{s}}^{1/s}a_{\min}^{-1/s}\right)\|f\|_{L^{2}}\,. (3.15)
Proof.

The estimate follows from the a-priori bounds for s=0s=0 and s=1s=1, i.e., (3.4) and (3.9), by interpolation with the Lipschitz continuity (3.6) of the solution operator.

Let aCs(D)a\in C^{s}({D}) with amin>0a_{\min}>0 be given. From (3.13), for every δ>0\delta>0 exists aδW1,(D)a_{\delta}\in W^{1,\infty}({D}) with

aaδC0CδsaCs,aδW1Cδs1aCs.\|a-a_{\delta}\|_{C^{0}}\leq C\delta^{s}\|a\|_{C^{s}}\;,\quad\|a_{\delta}\|_{W^{1}_{\infty}}\leq C\delta^{s-1}\|a\|_{C^{s}}\;.

From

min𝒙Daδ(𝒙)min𝒙Da(𝒙)aaδC0aminCδsaCs\min_{{\boldsymbol{x}}\in{D}}a_{\delta}({\boldsymbol{x}})\geq\min_{{\boldsymbol{x}}\in{D}}a({\boldsymbol{x}})-\|a-a_{\delta}\|_{C^{0}}\geq a_{\min}-C\delta^{s}\|a\|_{C^{s}}

follows for 0<δ21/sa/aminCs1/s,0<\delta\leq 2^{-1/s}\left\|a/a_{\min}\right\|_{C^{s}}^{-1/s}, that

min𝒙Daδ(𝒙)amin/2.\min_{{\boldsymbol{x}}\in{D}}a_{\delta}({\boldsymbol{x}})\geq a_{\min}/2\;.

For such δ\delta and for fL2(D)f\in L^{2}({D}), (3.1) with aδa_{\delta} admits a unique solution uδVu_{\delta}\in V and from (3.9)

uδW2camin(1+aδLamin)fL2.\|u_{\delta}\|_{W}\leq\frac{2c}{a_{\min}}\left(1+\frac{\|\nabla a_{\delta}\|_{L^{\infty}}}{a_{\min}}\right)\|f\|_{L^{2}}\;.

From (3.6) (with f1=f2=ff_{1}=f_{2}=f) we find

uuδV2camin2aaδLfL2Cδsamin2aCsfL2.\|u-u_{\delta}\|_{V}\leq\frac{2c}{a_{\min}^{2}}\|a-a_{\delta}\|_{L^{\infty}}\|f\|_{L^{2}}\leq C\frac{\delta^{s}}{a_{\min}^{2}}\|a\|_{C^{s}}\|f\|_{L^{2}}\;.

This implies in (3.11) that for some constant C>0C>0 (depending only on D{D} and on ss)

K(u,t,V,W)Camin(δsAs+t(1+δs1As))fL2,t>0K(u,t,V,W)\leq\frac{C}{a_{\min}}\left(\delta^{s}A_{s}+t\left(1+\delta^{s-1}A_{s}\right)\right)\|f\|_{L^{2}}\;,\;\;t>0 (3.16)

where we have set As:=aaminCs[1,)A_{s}:=\big{\|}\frac{a}{a_{\min}}\big{\|}_{C^{s}}\in[1,\infty).

To complete the proof, by (3.14) we bound uWs=supt>0tsK(u,t,V,W)\|u\|_{W^{s}}=\sup_{t>0}t^{-s}K(u,t,V,W). To this end, it suffices to bound K(u,t,V,W)K(u,t,V,W) for 0<t<10<t<1. Given such tt, we choose in the bound (3.16) δ=tδ0(0,δ0)\delta=t\delta_{0}\in(0,\delta_{0}) with δ0:=21/sAs1/s\delta_{0}:=2^{-1/s}A_{s}^{-1/s}. This yields

δsAs+t(1+δs1As)=ts(δ0sAs+t1s+δ0s1As)=ts(21+t1s+2(s1)/sAs1(s1)/s)\delta^{s}A_{s}+t\left(1+\delta^{s-1}A_{s}\right)=t^{s}\left(\delta_{0}^{s}A_{s}+t^{1-s}+\delta_{0}^{s-1}A_{s}\right)=t^{s}\left(2^{-1}+t^{1-s}+2^{-(s-1)/s}A_{s}^{1-(s-1)/s}\right)

and we obtain for 0<t<10<t<1 the bound

tsK(u,t,V,W)Camin(2+2(s1)/sAs1/s)fL2.t^{-s}K(u,t,V,W)\leq\frac{C}{a_{\min}}\left(2+2^{-(s-1)/s}A_{s}^{1/s}\right)\|f\|_{L^{2}}\;.

Adjusting the value of the constant CC, we arrive at (3.15). ∎

3.4 Random input data

We are in particular interested in the input data aa and ff of the elliptic divergence-form PDE (3.1) being not precisely known. The Lipschitz continuous data-dependence in Proposition 3.1 of the variational solution uVu\in V of (3.1) will ensure that small variations in the data (a,f)L+(D)×V(a,f)\in L^{\infty}_{+}({D})\times V^{*} imply corresponding small changes in the (unique) solution uVu\in V. A natural paradigm is to model uncertain data probabilistically. To this end, we work with a base probability space (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}). Given a known right hand side fL2(D)f\in L^{2}({D}), and uncertain diffusion coefficient aEL+(D)a\in E\subseteq L^{\infty}_{+}({D}), where EE denotes a suitable subset of L+(D)L^{\infty}_{+}({D}) of admissible diffusion coefficients, we model the function aa or loga\log a as RVs taking values in a subset EE of L(D)L^{\infty}({D}). We will assume the random data aa to be separably-valued, more precisely, the set EE of admissible random data will almost surely belong to a subset of a separable subspace of L(D)L^{\infty}({D}). See [21, Chap. 2.6] for details on separable-valuedness. Separability of EE is natural from the point of view of numerical approximation of (samples of) random input aa and simplifies many technicalities in the mathematical description; we refer in particular to the construction of GMs on EE in Sections 2.22.5. One valid choice for the space of admissible input data EE consists in E=C(D¯)L+(D)E=C(\overline{{D}})\cap L^{\infty}_{+}({D}). In the log-Gaussian models to be analyzed subsequently, EL+(D)E\subset L^{\infty}_{+}({D}) will be ensured by modelling log(a)\log(a) as a GRF, i.e., we assume the probability measure \mathbb{P} to be such that the law of log(a)\log(a) is a GM on L(D)L^{\infty}({D}) which charges EE, so that the random element log(a(,ω))L+(D)\log(a(\cdot,\omega))\in L^{\infty}_{+}({D}) \mathbb{P}-a.s.. This, in turn, implies with the well-posedness result in Section 3.1 that there exists a unique random solution u(ω)=𝒮(a,f)Vu(\omega)=\mathcal{S}(a,f)\in V \mathbb{P}-a.s.. Furthermore, the Lipschitz continuity (3.8) then implies that the corresponding map ωu(ω)\omega\mapsto u(\omega) is a composition of the measurable map ωlog(a(,ω))\omega\mapsto\log(a(\cdot,\omega)) with the Lipschitz continuous deterministic data-to-solution map 𝒮\mathcal{S}, hence strongly measurable, and thus a RV on (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}) taking values in VV.

3.5 Parametric deterministic coefficient

A key step in the deterministic numerical approximation of the elliptic divergence-form PDE (3.1) with log-Gaussian random inputs (i.e., log(a)\log(a) is a GRF on a suitable locally convex space EE of admissible input data) is to place a GM on EE and to describe the realizations of GRF bb in terms of affine-parametric representations discussed in Section 2.5. In Section 3.5.1, we briefly describe this and in doing so extend a-priori estimates to this resulting deterministic parametric version of elliptic PDE (3.1). Subsequently, in Section 3.5.3, we show that the resulting, countably-parametric, linear elliptic problem admits an extension to certain complex parameter domains, while still remaining well-posed.

3.5.1 Deterministic countably parametric elliptic PDEs

Placing a Gaussian probability measure on the random inputs log(a)\log(a) to the elliptic divergence-form PDE (3.1) can be achieved via Gaussian series as discussed in Section 2.5. Affine-parametric representations which are admissible in the sense of Definition 2.19 of the random input log(a)\log(a) of (3.1), subject to a Gaussian law on the corresponding input locally convex space EE, render the elliptic divergence-form PDE (3.1) with random inputs a deterministic parametric elliptic PDE. More precisely, b:=log(a)b:=\log(a) will depend on the sequence 𝒚=(yj)j{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}} of parameters from the parameter space {\mathbb{R}}^{\infty}. Accordingly, we consider parametric diffusion coefficients a=a(𝒚)a=a({\boldsymbol{y}}), where

𝒚=(yj)jU.{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}}\in U.

Here and throughout the rest of this book we make use of the notation

U:=.U:={\mathbb{R}}^{\infty}.

We develop the holomorphy-based analysis of parametric regularity and Wiener-Hermite PC expansion coefficient sparsity for the model parametric linear second order elliptic divergence-form PDE with so-called “log-affine coefficients”

div(exp(b(𝒚))u(𝒚))=finD,u(𝒚)|D=0,-\operatorname{div}\big{(}\exp(b({\boldsymbol{y}}))\nabla u({\boldsymbol{y}})\big{)}=f\quad\mbox{in}\quad{D}\;,\quad u({\boldsymbol{y}})|_{\partial{D}}=0\;, (3.17)

i.e.,

a(𝒚)=exp(b(𝒚)).a({\boldsymbol{y}})=\exp(b({\boldsymbol{y}})).

Here, the coefficient b(𝒚)=log(a(𝒚))b({\boldsymbol{y}})=\log(a({\boldsymbol{y}})) is assumed to be affine-parametric

b(𝒚)=jyjψj(𝒙),𝒙D,𝒚U.b({\boldsymbol{y}})=\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}({\boldsymbol{x}})\;,\quad{\boldsymbol{x}}\in{D}\;,\quad{\boldsymbol{y}}\in U\;. (3.18)

We assume that ψjEL(D)\psi_{j}\in E\subset L^{\infty}({{D}}) for every jj\in{\mathbb{N}}. For any 𝒚U{\boldsymbol{y}}\in U such that b(𝒚)L(D)b({\boldsymbol{y}})\in L^{\infty}({{D}}), by (3.4) we have the estimate

u(𝒚)VfVa(𝒚)1Lexp(b(𝒚)L)fV.\|u({\boldsymbol{y}})\|_{V}\leq\|f\|_{V^{*}}\|a({\boldsymbol{y}})^{-1}\|_{L^{\infty}}\leq\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\|f\|_{V^{*}}\,. (3.19)

For every 𝒚U{\boldsymbol{y}}\in U satisfying b(𝒚)L(D)b({\boldsymbol{y}})\in L^{\infty}({{D}}), the variational form (3.2) of (3.17) gives rise to the parametric energy norm va(𝒚)\|v\|_{a({\boldsymbol{y}})} on VV which is defined by

va(𝒚)2:=Da(𝒚)|v|2d𝒙,vV.\|v\|_{a({\boldsymbol{y}})}^{2}:=\int_{{D}}a({\boldsymbol{y}})|\nabla v|^{2}\,\mathrm{d}{\boldsymbol{x}}\;,\;\;v\in V.

The norms a(𝒚)\|\circ\|_{a({\boldsymbol{y}})} and V\|\circ\|_{V} are equivalent on VV but not uniformly w.r.t. 𝒚{\boldsymbol{y}}. It holds

exp(b(𝒚)L)vV2va(𝒚)2exp(b(𝒚)L)vV2,vV.\exp(-\|b({\boldsymbol{y}})\|_{L^{\infty}})\|v\|_{V}^{2}\leq\|v\|_{a({\boldsymbol{y}})}^{2}\leq\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\|v\|_{V}^{2},\quad v\in V\;. (3.20)

3.5.2 Probabilistic setting

In a probabilistic setting, the parameter sequence 𝒚{\boldsymbol{y}} is chosen as a sequence of i.i.d.  standard Gaussian RVs 𝒩(0,1){\mathcal{N}}(0,1) and (ψj)j(\psi_{j})_{j\in{\mathbb{N}}} a given sequence of functions in the Banach space L(D)L^{\infty}({D}) to which we refer as representation system of the uncertain input. We then treat (3.17) as the stochastic linear second order elliptic divergence-form PDE with so-called “log-Gaussian coefficients”. We refer to Section 2.5 for the construction of GMs based on affine representation systems (ψj)j(\psi_{j})_{j\in{\mathbb{N}}}. Due to L(D)L^{\infty}({D}) being non-separable, we consider GRFs b(𝒚)b({\boldsymbol{y}}) which take values in separable subspaces EL(D)E\subset L^{\infty}({D}), such as E=C0(D¯)E=C^{0}(\overline{{D}}).

The probability space (Ω,𝒜,)(\Omega,\mathcal{A},\mathbb{P}) from Section 3.4 on the parametric solutions {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} is chosen as (U,(U);γ)(U,\mathcal{B}(U);\gamma). Here and throughout the rest of this book, we make use of the notation: (U)\mathcal{B}(U) is the σ\sigma-field on the locally convex space UU generated by cylinders of Borel sets on {\mathbb{R}}, and γ\gamma is the product measure of the standard GM γ1\gamma_{1} on \mathbb{R} (see the definition in Example 2.17). We shall refer to γ\gamma as the standard GM on UU.

It follows from the a-priori estimate (3.19) that for fVf\in V^{*} the parametric elliptic diffusion problem (3.17) admits a unique solution for parameters 𝒚{\boldsymbol{y}} in the set

U0:={𝒚U:b(𝒚)L(D)}.U_{0}:=\{{\boldsymbol{y}}\in U:b({\boldsymbol{y}})\in L^{\infty}({D})\}\;. (3.21)

The measure γ(U0)\gamma(U_{0}) of the set U0UU_{0}\subset U depends on the structure of 𝒚b(𝒚){\boldsymbol{y}}\mapsto b({\boldsymbol{y}}). The following sufficient condition on the representation system (ψj)j(\psi_{j})_{j\in{\mathbb{N}}} will be assumed throughout.

Assumption 3.6.

For every jj\in{\mathbb{N}}, ψjL(D)\psi_{j}\in L^{\infty}({D}), and there exists a positive sequence (λj)j(\lambda_{j})_{j\in{\mathbb{N}}} such that (exp(λj2))j1()\big{(}\exp(-\lambda_{j}^{2})\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}}) and the series jλj|ψj|\sum_{j\in{\mathbb{N}}}\lambda_{j}|\psi_{j}| converges in L(D)L^{\infty}({D}).

For the statement of the next result, we recall a notion of Bochner spaces. For a measure space (Ω,𝒜,μ)(\Omega,\mathcal{A},\mu) let XX a Banach space and 1p<1\leq p<\infty. Then the Bochner space Lp(Ω,X;μ)L^{p}(\Omega,X;\mu) is defined as the space of all strongly μ\mu-measurable mappings uu from Ω\Omega to XX such that the norm

uLp(Ω,X;μ):=(Ωu(𝒚)Xpdμ(𝒚))1/p<.\|u\|_{L^{p}(\Omega,X;\mu)}:=\ \left(\int_{\Omega}\|u({\boldsymbol{y}})\|_{X}^{p}\,\,\mathrm{d}\mu({\boldsymbol{y}})\right)^{1/p}<\infty. (3.22)

In particular, when (Ω,𝒜,μ)=(U,(U);γ)(\Omega,\mathcal{A},\mu)=(U,\mathcal{B}(U);\gamma), XX is separable and p=2p=2, the hilbertian space L2(U,X;γ)L^{2}(U,X;\gamma) is one of the most important for the problems considered in this book.

The following result was shown in [9, Theorem 2.2].

Proposition 3.7.

Under Assumption 3.6, the set U0U_{0} has full GM, i.e., γ(U0)=1\gamma(U_{0})=1. For all kk\in{\mathbb{N}} there holds, with 𝔼()\mathbb{E}(\cdot) denoting expectation with respect to γ\gamma,

𝔼(exp(kb()L))<.\mathbb{E}\left(\exp(k\|b(\cdot)\|_{L^{\infty}})\right)<\infty\;.

The solution family {u(𝐲):𝐲U0}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U_{0}\} of the parametric elliptic boundary value problem (3.17) is in Lk(U,V;γ)L^{k}(U,V;\gamma) for every finite kk\in{\mathbb{N}}.

3.5.3 Deterministic complex-parametric elliptic PDEs

Towards the aim of establishing sparsity of Wiener-Hermite PC expansions of the parametric solutions {u(𝒚):𝒚U0}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in{U_{0}}\} of (3.17), we extend the deterministic parametric elliptic problem (3.17) from real-valued to complex-valued parameters.

Formally, replacing 𝒚=(yj)jU{\boldsymbol{y}}=(y_{j})_{j\in{\mathbb{N}}}\in U in the coefficient a(𝒚)a({\boldsymbol{y}}) by 𝒛=(zj)j=(yj+iξj)j{\boldsymbol{z}}=(z_{j})_{j\in{\mathbb{N}}}=(y_{j}+{\rm i}\xi_{j})_{j\in{\mathbb{N}}}\in{\mathbb{C}}^{\infty}, the real part of a(𝒛)a({\boldsymbol{z}}) is

[a(𝒛)]=exp(jyjψj(𝒙))cos(jξjψj(𝒙)).\mathfrak{R}[a({\boldsymbol{z}})]=\exp\Bigg{(}{\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}({\boldsymbol{x}})}\Bigg{)}\cos\Bigg{(}\sum_{j\in{\mathbb{N}}}\xi_{j}\psi_{j}({\boldsymbol{x}})\Bigg{)}\,. (3.23)

We find that [a(𝒛)]>0\mathfrak{R}[a({\boldsymbol{z}})]>0 if

jξjψjL<π2.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\xi_{j}\psi_{j}\Bigg{\|}_{L^{\infty}}<\frac{\pi}{2}.

This observation and Proposition 3.7 motivate the study of the analytic continuation of the solution map 𝒚u(𝒚){\boldsymbol{y}}\mapsto u({\boldsymbol{y}}) to 𝒛u(𝒛){\boldsymbol{z}}\mapsto u({\boldsymbol{z}}) for complex parameters 𝒛=(zj)j{\boldsymbol{z}}=(z_{j})_{j\in{\mathbb{N}}} by formally replacing the parameter yjy_{j} by zjz_{j} in the definition of the parametric coefficient aa, where each zjz_{j} lies in the strip

𝒮j(𝝆):={zj:|𝔪zj|<ρj}\mathcal{S}_{j}({\boldsymbol{\rho}}):=\{z_{j}\in{\mathbb{C}}\,:|\mathfrak{Im}z_{j}|<\rho_{j}\} (3.24)

and where ρj>0\rho_{j}>0 and 𝝆=(ρj)j(0,){\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}}\in(0,\infty)^{\infty} is any sequence of positive numbers such that

jρj|ψj|L<π2.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}<\frac{\pi}{2}\,.

3.6 Analyticity and sparsity

We address the analyticity (holomorphy) of the parametric solutions {u(𝒚):𝒚U0}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in{U_{0}}\}. We analyze the sparsity by estimating, in particular, the size of the domains of holomorphy to which the parametric solutions can be extended. We also treat the weighted 2\ell^{2}-summability and p\ell^{p}-summability (sparsity) for the series of Wiener-Hermite the PC expansion coefficients (u𝝂)𝝂(u_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} of u(𝒚)u({\boldsymbol{y}}).

3.6.1 Parametric holomorphy

In this section we establish holomorphic parametric dependence uu on aa and on ff as in [39] by verifying complex differentiability of a suitable complex-parametric extension of 𝒚u(𝒚){\boldsymbol{y}}\mapsto u({\boldsymbol{y}}). We observe that the Lax-Milgram theory can be extended to the case where the coefficient function aa is complex-valued. In this case, V:=H01(D,)V:=H^{1}_{0}({D},{\mathbb{C}}) in (3.2) and the ellipticity assumption (3.3) is extended to the complex domain as

0<ρ(a):=essinf𝒙D(a(𝒙))|a(𝒙)|aL<,𝒙D.0<\rho(a):=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,\Re(a({\boldsymbol{x}}))\leq|a({\boldsymbol{x}})|\leq\|a\|_{L^{\infty}}<\infty,\qquad{\boldsymbol{x}}\in{D}. (3.25)

Under this condition, there exists a unique variational solution uVu\in V of (3.1) and for this solution, the estimate (3.4) remains valid, i.e.,

uVfVρ(a).\|u\|_{V}\leq\frac{\|f\|_{V^{*}}}{\rho(a)}\,. (3.26)

Let 𝝆=(ρj)j[0,){\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} be a sequence of non-negative numbers and assume that 𝔲supp(𝝆){\mathfrak{u}}\subseteq\operatorname{supp}({\boldsymbol{\rho}}) is finite. Define

𝒮𝔲(𝝆):=×j𝔲𝒮j(𝝆),\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}):=\bigtimes_{j\in{\mathfrak{u}}}\mathcal{S}_{j}({\boldsymbol{\rho}})\,, (3.27)

where the strip 𝒮j(𝝆)\mathcal{S}_{j}({\boldsymbol{\rho}}) is given in (3.24). For 𝒚U{\boldsymbol{y}}\in U, put

𝒮𝔲(𝒚,𝝆):={(zj)j:zj𝒮j(𝝆)ifj𝔲andzj=yjifj𝔲}.\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}):=\big{\{}(z_{j})_{j\in{\mathbb{N}}}:z_{j}\in\mathcal{S}_{j}({\boldsymbol{\rho}})\ \text{if}\ j\in{\mathfrak{u}}\ \text{and}\ z_{j}=y_{j}\ \text{if}\ j\not\in{\mathfrak{u}}\big{\}}.
Proposition 3.8.

Let the sequence 𝛒=(ρj)j[0,){\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} satisfy

jρj|ψj|Lκ<π2.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2}\,. (3.28)

Let 𝐲0=(y0,1,y0,2,)U{\boldsymbol{y}}_{0}=(y_{0,1},y_{0,2},\ldots)\in U be such that b(𝐲0)b({\boldsymbol{y}}_{0}) belongs to L(D)L^{\infty}({{D}}), and let 𝔲supp(𝛒){\mathfrak{u}}\subseteq\operatorname{supp}({\boldsymbol{\rho}}) be a finite set.

Then the solution uu of the variational form of (3.17) is holomorphic on 𝒮𝔲(𝛒)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}) as a function of the parameters 𝐳𝔲=(zj)j𝒮𝔲(𝐲0,𝛒){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) taking values in VV with zj=y0,jz_{j}=y_{0,j} for j𝔲j\not\in{\mathfrak{u}} held fixed.

Proof.

Let NN\in{\mathbb{N}}. We denote

𝒮𝔲,N(𝝆):={(yj+iξj)j𝔲𝒮𝔲(𝝆):|yjy0,j|<N}.\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}):=\big{\{}(y_{j}+{\rm i}\xi_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}):|y_{j}-y_{0,j}|<N\big{\}}\,. (3.29)

For 𝒛𝔲=(yj+iξj)j𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{{\mathfrak{u}}}=(y_{j}+{\rm i}\xi_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) with (yj+iξj)j𝔲𝒮𝔲,N(𝝆)(y_{j}+{\rm i}\xi_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}) we have

jyjψjLb(𝒚0)L+j𝔲|(yy0,j)ψj|Lb(𝒚0)L+Nj𝔲|ψj|L=:M<\begin{split}\Bigg{\|}\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}\Bigg{\|}_{L^{\infty}}&\leq\|b({\boldsymbol{y}}_{0})\|_{L^{\infty}}+\Bigg{\|}\sum_{j\in{\mathfrak{u}}}|(y-y_{0,j})\psi_{j}|\Bigg{\|}_{L^{\infty}}\\ &\leq\|b({\boldsymbol{y}}_{0})\|_{L^{\infty}}+N\Bigg{\|}\sum_{j\in{\mathfrak{u}}}|\psi_{j}|\Bigg{\|}_{L^{\infty}}=:M<\infty\end{split}

and

j𝔲ξjψjLj𝔲|ρjψj|Lκ.\Bigg{\|}\sum_{j\in{\mathfrak{u}}}\xi_{j}\psi_{j}\Bigg{\|}_{L^{\infty}}\leq\Bigg{\|}\sum_{j\in{\mathfrak{u}}}|\rho_{j}\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa\,.

Consequently, we obtain from (3.23)

ρ(a(𝒛𝔲))exp(jyjψjL)cos(j𝔲ξjψjL)exp(M)cosκ\rho(a({\boldsymbol{z}}_{{\mathfrak{u}}}))\geq\exp\Bigg{(}-\Bigg{\|}\sum_{j\in{\mathbb{N}}}y_{j}\psi_{j}\Bigg{\|}_{L^{\infty}}\Bigg{)}\cos\Bigg{(}\Bigg{\|}\sum_{j\in{\mathfrak{u}}}\xi_{j}\psi_{j}\Bigg{\|}_{L^{\infty}}\Bigg{)}\geq\exp(-M)\cos\kappa (3.30)

for all 𝒛𝔲𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) with (yj+iξj)j𝔲𝒮𝔲,N(𝝆)(y_{j}+{\rm i}\xi_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}). From this and the analyticity of exponential functions we conclude that the map 𝒛𝔲u(𝒛𝔲){\boldsymbol{z}}_{\mathfrak{u}}\to u({\boldsymbol{z}}_{{\mathfrak{u}}}) is holomorphic on the set 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}), see [37, Pages 22, 23]. Since NN is arbitrary we deduce that the map 𝒛𝔲u(𝒛𝔲){\boldsymbol{z}}_{\mathfrak{u}}\to u({\boldsymbol{z}}_{{\mathfrak{u}}}) is holomorphic on 𝒮𝔲(𝝆)\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{\rho}}). ∎

The analytic continuation of the parametric solutions {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} to 𝒮𝔲(𝝆)\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{\rho}}) leads to a result on parametric VV-regularity.

Lemma 3.9.

Let 𝛒=(ρj)j{\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}} be a non-negative sequence satisfying (3.28). Let 𝐲U{\boldsymbol{y}}\in U with b(𝐲)L(D)b({\boldsymbol{y}})\in L^{\infty}({{D}}) and 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}} such that supp(𝛎)supp(𝛒)\operatorname{supp}({\boldsymbol{\nu}})\subseteq\operatorname{supp}({\boldsymbol{\rho}}). Then we have

𝝂u(𝒚)VC0𝝂!𝝆𝝂exp(b(𝒚)L),\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{V}\leq C_{0}\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}^{\boldsymbol{\nu}}}\exp\big{(}\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)},

where C0=eκ(cosκ)1fVC_{0}=e^{\kappa}(\cos\kappa)^{-1}\|f\|_{V^{*}}.

Proof.

Let 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} such that supp(𝝂)supp(𝝆)\operatorname{supp}({\boldsymbol{\nu}})\subseteq\operatorname{supp}({\boldsymbol{\rho}}). Denote 𝔲=supp(𝝂){\mathfrak{u}}=\operatorname{supp}({\boldsymbol{\nu}}). For fixed variable yjy_{j} with j𝔲j\not\in{\mathfrak{u}}, the map 𝒮𝔲(𝒚,𝝆)𝒛𝔲u(𝒛𝔲)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}})\ni{\boldsymbol{z}}_{{\mathfrak{u}}}\to u({\boldsymbol{z}}_{\mathfrak{u}}) is holomorphic on the domain 𝒮𝔲(𝒚,κ𝝆)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}},\kappa^{\prime}{\boldsymbol{\rho}}) where κ<κκ<π/2\kappa<\kappa\kappa^{\prime}<\pi/2, see Proposition 3.8. Applying Cauchy’s integral formula gives

𝝂u(𝒚)=𝝂!(2πi)|𝔲|𝒞𝒚,𝔲(𝝆)u(𝒛𝔲)j𝔲(zjyj)νj+1j𝔲dzj,\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})=\frac{{\boldsymbol{\nu}}!}{(2\pi i)^{|{\mathfrak{u}}|}}\int_{\mathcal{C}_{{\boldsymbol{y}},{\mathfrak{u}}}({\boldsymbol{\rho}})}\frac{u({\boldsymbol{z}}_{\mathfrak{u}})}{\prod_{j\in{\mathfrak{u}}}(z_{j}-y_{j})^{\nu_{j}+1}}\prod_{j\in{\mathfrak{u}}}\,\mathrm{d}z_{j},

where

𝒞𝒚,𝔲(𝝆):=×j𝔲𝒞𝒚,j(𝝆),𝒞𝒚,j(𝝆):={zj:|zjyj|=ρj}.\mathcal{C}_{{\boldsymbol{y}},{\mathfrak{u}}}({\boldsymbol{\rho}}):=\bigtimes_{j\in{\mathfrak{u}}}\mathcal{C}_{{\boldsymbol{y}},j}({\boldsymbol{\rho}})\,,\qquad\mathcal{C}_{{\boldsymbol{y}},j}({\boldsymbol{\rho}}):=\big{\{}z_{j}\in{\mathbb{C}}:|z_{j}-y_{j}|=\rho_{j}\big{\}}\,. (3.31)

This leads to

𝝂u(𝒚)V𝝂!𝝆𝝂supz𝔲𝒞𝔲(𝒚,𝝆)u(𝒛𝔲)V\begin{split}\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{V}&\leq\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}^{\boldsymbol{\nu}}}\sup_{z_{\mathfrak{u}}\in\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}})}\|u({\boldsymbol{z}}_{\mathfrak{u}})\|_{V}\,\end{split} (3.32)

with

𝒞𝔲(𝒚,𝝆)={(zj)j𝒮𝔲(𝒚,𝝆):(zj)j𝔲𝒞𝒚,𝔲(𝝆)}.\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}})=\big{\{}(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}):\ (z_{j})_{j\in{\mathfrak{u}}}\in\mathcal{C}_{{\boldsymbol{y}},{\mathfrak{u}}}({\boldsymbol{\rho}})\big{\}}\,. (3.33)

Notice that for 𝒛𝔲=(zj)j𝒞𝔲(𝒚,𝝆){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}) we can write zj=yj+ηj+iξj𝒞𝒚,j(𝝆)z_{j}=y_{j}+\eta_{j}+{\rm i}\xi_{j}\in{\mathcal{C}}_{{\boldsymbol{y}},j}({\boldsymbol{\rho}}) with |ηj|ρj|\eta_{j}|\leq\rho_{j}, |ξj|ρj|\xi_{j}|\leq\rho_{j} if j𝔲j\in{\mathfrak{u}} and ηj=ξj=0\eta_{j}=\xi_{j}=0 if j𝔲j\not\in{\mathfrak{u}}. By denoting 𝜼=(ηj)j{\boldsymbol{\eta}}=(\eta_{j})_{j\in{\mathbb{N}}} and 𝝃=(ξj)j{\boldsymbol{\xi}}=(\xi_{j})_{j\in{\mathbb{N}}} we see that b(𝜼)Lκ\|b({\boldsymbol{\eta}})\|_{L^{\infty}}\leq\kappa and b(𝝃)Lκ\|b({\boldsymbol{\xi}})\|_{L^{\infty}}\leq\kappa. Hence we deduce from (3.26) that

u(𝒛𝔲)Vexp(b(𝒚+𝜼)L)cos(b(𝝃)L)fVexp(κ+b(𝒚)L)cosκfV.\begin{split}\|u({\boldsymbol{z}}_{\mathfrak{u}})\|_{V}&\leq\frac{\exp\big{(}\|b({\boldsymbol{y}}+{\boldsymbol{\eta}})\|_{L^{\infty}}\big{)}}{\cos\big{(}\|b({\boldsymbol{\xi}})\|_{L^{\infty}}\big{)}}\|f\|_{V^{*}}\leq\frac{\exp\big{(}\kappa+\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}}{\cos\kappa}\|f\|_{V^{*}}\,.\end{split}

Inserting this into (3.32) we obtain the desired estimate. ∎

3.6.2 Sparsity of Wiener-Hermite PC expansion coefficients

In this section, we will exploit the analyticity of uu to prove a weighted 2\ell^{2}-summability result for the VV-norms of the coefficients in the Wiener-Hermite PC expansion of the solution map 𝒚u(𝒚){\boldsymbol{y}}\to u({\boldsymbol{y}}). Our analysis yields the same p\ell^{p}-summability result as in the papers [9, 8] in the case ψj\psi_{j} have arbitrary supports. In this case, our result implies that the p\ell^{p}-summability of (u𝝂V)(\|u_{\boldsymbol{\nu}}\|_{V})_{{\mathcal{F}}} for 0<p10<p\leq 1 (the sparsity of parametric solutions) follows from the p\ell^{p}-summability of the sequence (jαψjL)j(j^{\alpha}\|\psi_{j}\|_{L^{\infty}})_{j\in{\mathbb{N}}} for some α>1/2\alpha>1/2 which is an improvement over the condition (jψjL)jp()(j\|\psi_{j}\|_{L^{\infty}})_{j\in{\mathbb{N}}}\in\ell^{p}({\mathbb{N}}) in [72], see [9, Section 6.3]. In the case of disjoint or finitely overlapping supports our analysis obtains a weaker result compared to [9, 8]. As observed in [39], one advantage of establishing sparsity of Wiener-Hermite PC expansion coefficients via holomorphy rather than by successive differentiation is that it allows to derive, in a unified way, summability bounds for the coefficients of Wiener-Hermite PC expansion whose size is measured in scales of Sobolev and Besov spaces in the domain D{D}. Using real-variable arguments as, e.g., in [9, 8], establishing sparsity of parametric solutions in Besov spaces in D{D} of higher smoothness seems to require more involved technical and notational developments, according to [8, Comment on Page 2157].

The parametric solution {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} of (3.17) belongs to the space L2(U,V;γ)L^{2}(U,V;\gamma) or more generally, L2(U,(H1+sH01)(D);γ)L^{2}(U,(H^{1+s}\cap H^{1}_{0})({D});\gamma) for ss-order of extra differentiability provided by higher data regularity. We recall from Section 2.1.3 the normalized probabilistic Hermite polynomials (Hk)k0(H_{k})_{k\in{\mathbb{N}}_{0}}. Every uL2(U,X;γ)u\in L^{2}(U,X;\gamma) admits the Wiener-Hermite PC expansion

𝝂u𝝂H𝝂(𝒚),\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}u_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}}({\boldsymbol{y}}), (3.34)

where for 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}},

H𝝂(𝒚)=jHνj(yj),H_{\boldsymbol{\nu}}({\boldsymbol{y}})=\prod_{j\in{\mathbb{N}}}H_{\nu_{j}}(y_{j}),\quad

and

u𝝂:=Uu(𝒚)H𝝂(𝒚)dγ(𝒚)u_{\boldsymbol{\nu}}:=\int_{U}u({\boldsymbol{y}})\,H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}})

are called Wiener-Hermite PC expansion coefficients. Notice that (H𝝂)𝝂(H_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}} forms an ONB of L2(U;γ)L^{2}(U;\gamma).

For every uL2(U,X;γ)u\in L^{2}(U,X;\gamma), there holds the Parseval-type identity

uL2(U,X;γ)2=𝝂u𝝂X2,uL2(U,X;γ).\|u\|^{2}_{L^{2}(U,X;\gamma)}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\|u_{\boldsymbol{\nu}}\|_{X}^{2}\;,\quad u\in L^{2}(U,X;\gamma)\;. (3.35)

The error of approximation of the parametric solution {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} of (3.17) will be measured in the Bochner space L2(U,V;γ)L^{2}(U,V;\gamma). A basic role in this approximation is taken by the Wiener-Hermite PC expansion (3.34) of uu in the space L2(U,V;γ)L^{2}(U,V;\gamma).

For a finite set Λ\Lambda\subset\mathcal{F}, we denote by uΛ=𝝂Λu𝝂u_{\Lambda}=\sum_{{\boldsymbol{\nu}}\in\Lambda}u_{\boldsymbol{\nu}} the corresponding partial sum of the Wiener-Hermite PC expansion (3.34). It follows from (3.35) that

uuΛL2(U,V;γ)2=𝝂\Λu𝝂V2.\|u-u_{\Lambda}\|_{L^{2}(U,V;\gamma)}^{2}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda}\|u_{\boldsymbol{\nu}}\|_{V}^{2}\;.

Therefore, summability results of the coefficients (u𝝂V)𝝂(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in\mathcal{F}} imply convergence rate estimates of finitely truncated expansions uΛnu_{\Lambda_{n}} for suitable sequences (Λn)n(\Lambda_{n})_{n\in{\mathbb{N}}} of sets of nn indices 𝝂{\boldsymbol{\nu}} (see [72, 9, 43]). We next recapitulate some weighted summability results for Wiener-Hermite expansions.

For rr\in{\mathbb{N}} and a sequence of nonnegative numbers ϱ=(ϱj)j{\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}}, we define the Wiener-Hermite weights

β𝝂(r,ϱ):=𝝂r(𝝂𝝂)ϱ2𝝂=j(=0r(νj)ϱj2),𝝂.\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}):=\sum_{\|{\boldsymbol{\nu}}^{\prime}\|_{\ell^{\infty}}\leq r}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\nu}}^{\prime}}{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}^{\prime}}=\prod_{j\in{\mathbb{N}}}\Bigg{(}\sum_{\ell=0}^{r}\binom{\nu_{j}}{\ell}\varrho_{j}^{2\ell}\Bigg{)},\ \ {\boldsymbol{\nu}}\in{\mathcal{F}}\,. (3.36)

The following identity was proved in [9, Theorem 3.3]. For convenience to the reader, we present the proof from that paper.

Lemma 3.10.

Let Assumption 3.6 hold. Let rr\in{\mathbb{N}} and ϱ=(ϱj)j{\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}} be a sequence of nonnegative numbers. Then

𝝂β𝝂(r,ϱ)u𝝂V2=𝝂rϱ2𝝂𝝂!U𝝂u(𝒚)V2dγ(𝒚).\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{V}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\,. (3.37)
Proof.

Recall that p(y):=p(y,0,1)=12πexp(y2/2)p(y):=p(y,0,1)=-\frac{1}{\sqrt{2\pi}}\exp(-y^{2}/2) is the density function of the standard GM on {\mathbb{R}}. Let μ\mu\in{\mathbb{N}}. For a sufficiently smooth, univariate function vL2(;γ)v\in L^{2}(\mathbb{R};\gamma), from Hν(y)=(1)νν!p(ν)(y)p(y)H_{\nu}(y)=\frac{(-1)^{\nu}}{\sqrt{\nu!}}\frac{p^{(\nu)}(y)}{p(y)} we have for νμ\nu\geq\mu

vν\displaystyle v_{\nu} :=v(y)Hν(y)p(y)dy=(1)νν!v(y)p(ν)(y)dy\displaystyle:=\int_{{\mathbb{R}}}v(y)H_{\nu}(y)p(y)\,\mathrm{d}y=\frac{(-1)^{\nu}}{\sqrt{\nu!}}\int_{\mathbb{R}}v(y)p^{(\nu)}(y)\,\mathrm{d}y
=(1)νμν!v(μ)(y)p(νμ)(y)dy=(νμ)!ν!v(μ)(y)Hνμ(y)p(y)dy.\displaystyle=\frac{(-1)^{\nu-\mu}}{\sqrt{\nu!}}\int_{\mathbb{R}}v^{(\mu)}(y)p^{(\nu-\mu)}(y)\,\mathrm{d}y=\sqrt{\frac{(\nu-\mu)!}{\nu!}}\int_{\mathbb{R}}v^{(\mu)}(y)H_{\nu-\mu}(y)p(y)\,\mathrm{d}y.

Hence

ν!μ!(νμ)!vν=1μ!v(μ)(y)Hνμ(y)dγ(y).\sqrt{\frac{\nu!}{\mu!(\nu-\mu)!}}v_{\nu}=\sqrt{\frac{1}{\mu!}}\int_{\mathbb{R}}v^{(\mu)}(y)H_{\nu-\mu}(y)\,\mathrm{d}\gamma(y).

By Parseval’s identity, we have

1μ!|v(μ)(y)|2dγ(y)=νμν!μ!(νμ)!|vν|2=ν0(νμ)|vν|2,\frac{1}{\mu!}\int_{\mathbb{R}}|v^{(\mu)}(y)|^{2}\,\mathrm{d}\gamma(y)=\sum_{\nu\geq\mu}\frac{\nu!}{\mu!(\nu-\mu)!}|v_{\nu}|^{2}=\sum_{\nu\in{\mathbb{N}}_{0}}\binom{\nu}{\mu}|v_{\nu}|^{2}\,,

where we use the convention (νμ)=0\binom{\nu}{\mu}=0 if μ>ν\mu>\nu.

For multi-indices and for uL2(U,V;γ)u\in L^{2}(U,V;\gamma), if 𝝁𝝂{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}, applying the above argument in coordinate-wise for the coefficients

u𝝂=(𝝂𝝁)!𝝂!U𝝁u(𝒚)H𝝂𝝁(𝒚)dγ(𝒚)u_{{\boldsymbol{\nu}}}=\sqrt{\frac{({\boldsymbol{\nu}}-{\boldsymbol{\mu}})!}{{\boldsymbol{\nu}}!}}\int_{U}\partial^{{\boldsymbol{\mu}}}u({\boldsymbol{y}})H_{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})

we get

1𝝁!U𝝁u(𝒚)V2dγ(𝒚)=𝝂𝝁𝝂!(𝝁!(𝝂𝝁)!)u𝝂V2=𝝂(𝝂𝝁)u𝝂V2.\frac{1}{{\boldsymbol{\mu}}!}\int_{U}\|\partial^{{\boldsymbol{\mu}}}u({\boldsymbol{y}})\|_{V}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\geq{\boldsymbol{\mu}}}\frac{{\boldsymbol{\nu}}!}{({\boldsymbol{\mu}}!({\boldsymbol{\nu}}-{\boldsymbol{\mu}})!)}\|u_{{\boldsymbol{\nu}}}\|_{V}^{2}=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}\|u_{{\boldsymbol{\nu}}}\|_{V}^{2}.

Multiplying both sides by ϱ2𝝁{\boldsymbol{\varrho}}^{2{\boldsymbol{\mu}}} and summing over 𝝁{\boldsymbol{\mu}} with 𝝁r\|{\boldsymbol{\mu}}\|_{\ell^{\infty}}\leq r, we obtain

𝝁rϱ2𝝁𝝁!U𝝁u(𝒚)V2dγ(𝒚)=𝝁r𝝂(𝝂𝝁)ϱ2𝝁u𝝂V2=𝝂β𝝂(r,ϱ)u𝝂V2.\sum_{\|{\boldsymbol{\mu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\mu}}}}{{\boldsymbol{\mu}}!}\int_{U}\|\partial^{{\boldsymbol{\mu}}}u({\boldsymbol{y}})\|_{V}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})=\sum_{\|{\boldsymbol{\mu}}\|_{\ell^{\infty}}\leq r}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}{\boldsymbol{\varrho}}^{2{\boldsymbol{\mu}}}\|u_{{\boldsymbol{\nu}}}\|_{V}^{2}=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}.

We recall a summability property of the sequence (β𝝂(r,ϱ)1)j(\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1})_{j\in{\mathbb{N}}} and its proof, given in [9, Lemma 5.1].

Lemma 3.11.

Let 0<p<0<p<\infty and q:=2p2pq:=\frac{2p}{2-p}. Let ϱ=(ϱj)j[0,){\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} be a sequence of positive numbers such that

(ϱj1)jq().(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}).

Then for any rr\in{\mathbb{N}} such that 2r+1<p\frac{2}{r+1}<p, the family (β𝛎(r,ϱ))𝛎(\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}))_{{\boldsymbol{\nu}}\in{\mathcal{F}}} defined in (3.36) for this rr satisfies

𝝂β𝝂(r,ϱ)q/2<.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}})^{-q/2}<\infty\,. (3.38)
Proof.

First we have the decomposition

𝝂b𝝂(r,ϱ)q/2=𝝂j(=0r(νj)ϱj2)q/2=jn0(=0r(n)ϱj2)q/2.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}b_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}})^{-q/2}=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\prod_{j\in{\mathbb{N}}}\bigg{(}\sum_{\ell=0}^{r}\binom{\nu_{j}}{\ell}\varrho_{j}^{2\ell}\bigg{)}^{-q/2}=\prod_{j\in{\mathbb{N}}}\sum_{n\in{\mathbb{N}}_{0}}\bigg{(}\sum_{\ell=0}^{r}\binom{n}{\ell}\varrho_{j}^{2\ell}\bigg{)}^{-q/2}.

For each jj\in{\mathbb{N}} we have

n0(=0r(n)ϱj2)q/2n0[(nmin{n,r})ϱj2min{n,r}]q/2=n=0r1ϱjnq+Cr,qϱjrq,\sum_{n\in{\mathbb{N}}_{0}}\bigg{(}\sum_{\ell=0}^{r}\binom{n}{\ell}\varrho_{j}^{2\ell}\bigg{)}^{-q/2}\leq\sum_{n\in{\mathbb{N}}_{0}}\bigg{[}\binom{n}{\min\{n,r\}}\varrho_{j}^{2\min\{n,r\}}\bigg{]}^{-q/2}=\sum_{n=0}^{r-1}\varrho_{j}^{-nq}+C_{r,q}\varrho_{j}^{-rq}, (3.39)

where

Cr,q:=n=r+(nr)q/2=(r!)q/2n0[(n+1)(n+r)]q/2.C_{r,q}:=\sum_{n=r}^{+\infty}\binom{n}{r}^{-q/2}=(r!)^{q/2}\sum_{n\in{\mathbb{N}}_{0}}\big{[}(n+1)\ldots(n+r)\big{]}^{-q/2}.

Since limn+(n+1)(n+r)nr=1\lim\limits_{n\to+\infty}\frac{(n+1)\ldots(n+r)}{n^{r}}=1, we find that Cr,qC_{r,q} is finite if and only if q>2/rq>2/r. This is equivalent to 2r+1<p\frac{2}{r+1}<p. From the assumption (ϱj1)jq()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}) we find some J>1J>1 such that ϱj>1\varrho_{j}>1 for all j>Jj>J. This implies ϱjnqϱjq\varrho_{j}^{-nq}\leq\varrho_{j}^{-q} for n=1,,rn=1,\ldots,r and j>Jj>J. Therefore, one can bound the right side of (3.39) by 1+(Cr,q+r1)ϱjq1+(C_{r,q}+r-1)\varrho_{j}^{-q}. Hence we obtain

𝝂b𝝂(r,ϱ)q/2\displaystyle\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}b_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}})^{-q/2} Cj>J[1+(Cr,q+r1)ϱjq]\displaystyle\leq C\prod_{j>J}\big{[}1+(C_{r,q}+r-1)\varrho_{j}^{-q}\big{]}
Cj>Jexp((Cr,q+r1)ϱjq)\displaystyle\leq C\prod_{j>J}\exp\Big{(}(C_{r,q}+r-1)\varrho_{j}^{-q}\Big{)}
Cexp((Cr,q+r1)(ϱj1)jqq)\displaystyle\leq C\exp\Big{(}(C_{r,q}+r-1)\|(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\|_{\ell^{q}}^{q}\Big{)}

which is finite since (ϱj1)jq()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}). ∎

In what follows, we denote by (𝒆j)j({\boldsymbol{e}}_{j})_{j\in{\mathbb{N}}} the standard basis of 2()\ell^{2}({\mathbb{N}}), i.e., 𝒆j=(ej,i)i{\boldsymbol{e}}_{j}=(e_{j,i})_{i\in{\mathbb{N}}} with ej,i=1e_{j,i}=1 for i=ji=j and ej,i=0e_{j,i}=0 for iji\not=j. The following lemma was obtained in [38, Lemma 7.1, Theorem 7.2] and [37, Lemma 3.17].

Lemma 3.12.

Let 𝛂=(αj)j{\boldsymbol{\alpha}}=(\alpha_{j})_{j\in{\mathbb{N}}} be a sequence of nonnegative numbers. Then we have the following.

  • (i)

    For 0<p<0<p<\infty, the family (𝜶𝝂)𝝂({\boldsymbol{\alpha}}^{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to p()\ell^{p}({\mathcal{F}}) if and only if 𝜶p<\|{\boldsymbol{\alpha}}\|_{\ell^{p}}<\infty and 𝜶<1\|{\boldsymbol{\alpha}}\|_{\ell^{\infty}}<1.

  • (ii)

    For 0<p10<p\leq 1, the family (𝜶𝝂|𝝂|!/𝝂!)𝝂({\boldsymbol{\alpha}}^{\boldsymbol{\nu}}|{\boldsymbol{\nu}}|!/{\boldsymbol{\nu}}!)_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to p()\ell^{p}({\mathcal{F}}) if and only if 𝜶p<\|{\boldsymbol{\alpha}}\|_{\ell^{p}}<\infty and 𝜶1<1\|{\boldsymbol{\alpha}}\|_{\ell^{1}}<1.

Proof.

Step 1. We prove the first statement. Assume that 𝜶<1\|{\boldsymbol{\alpha}}\|_{\ell^{\infty}}<1. Then we have

𝝂𝜶𝝂p\displaystyle\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}{\boldsymbol{\alpha}}^{{\boldsymbol{\nu}}p} =jn0αjpn=j11αjp\displaystyle=\prod_{j\in{\mathbb{N}}}\sum_{n\in{\mathbb{N}}_{0}}\alpha_{j}^{pn}=\prod_{j\in{\mathbb{N}}}\frac{1}{1-\alpha_{j}^{p}}
=j(1+αjp1αjp)jexp(αjp1αjp)\displaystyle=\prod_{j\in{\mathbb{N}}}\bigg{(}1+\frac{\alpha_{j}^{p}}{1-\alpha_{j}^{p}}\bigg{)}\leq\prod_{j\in{\mathbb{N}}}\exp\bigg{(}\frac{\alpha_{j}^{p}}{1-\alpha_{j}^{p}}\bigg{)}
jexp(11𝜶pαjp)=exp(11𝜶p𝜶pp),\displaystyle\leq\prod_{j\in{\mathbb{N}}}\exp\bigg{(}\frac{1}{1-\|{\boldsymbol{\alpha}}\|_{\ell^{\infty}}^{p}}\alpha_{j}^{p}\bigg{)}=\exp\bigg{(}\frac{1}{1-\|{\boldsymbol{\alpha}}\|_{\ell^{\infty}}^{p}}\|{\boldsymbol{\alpha}}\|_{\ell^{p}}^{p}\bigg{)}\,,

where in the last equality we have used (𝜶𝝂)𝝂p()({\boldsymbol{\alpha}}^{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}).

Since the sequence (αj)j=(𝜶𝒆j)j(\alpha_{j})_{j\in\mathbb{N}}=({\boldsymbol{\alpha}}^{{\boldsymbol{e}}_{j}})_{j\in{\mathbb{N}}} is a subsequence of (α𝝂)𝝂(\alpha^{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}, (𝜶𝝂)𝝂p()({\boldsymbol{\alpha}}^{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) implies 𝜶{\boldsymbol{\alpha}} belong to p()\ell^{p}({\mathbb{N}}). Moreover we have for any j1j\geq 1

n0αjnp=n0𝜶n𝒆jp𝝂𝜶𝝂p<.\sum_{n\in{\mathbb{N}}_{0}}\alpha_{j}^{np}=\sum_{n\in{\mathbb{N}}_{0}}{\boldsymbol{\alpha}}^{n{\boldsymbol{e}}_{j}p}\leq\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}{\boldsymbol{\alpha}}^{{\boldsymbol{\nu}}p}<\infty.

From this we have αjp<1\alpha_{j}^{p}<1 which implies αj<1\alpha_{j}<1 for all jj\in{\mathbb{N}}. Since 𝜶p(){\boldsymbol{\alpha}}\in\ell^{p}({\mathbb{N}}) it is easily seen that 𝜶<1\|{\boldsymbol{\alpha}}\|_{\ell^{\infty}}<1.

Step 2. We prove the second statement. We observe that

𝝂|𝝂|𝝂!𝜶𝝂=k0|𝝂|=k|𝝂|𝝂!𝜶𝝂=k0(jαj)k.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\frac{|{\boldsymbol{\nu}}|}{{\boldsymbol{\nu}}!}{\boldsymbol{\alpha}}^{\boldsymbol{\nu}}=\sum_{k\in{\mathbb{N}}_{0}}\sum_{|{\boldsymbol{\nu}}|=k}\frac{|{\boldsymbol{\nu}}|}{{\boldsymbol{\nu}}!}{\boldsymbol{\alpha}}^{\boldsymbol{\nu}}=\sum_{k\in{\mathbb{N}}_{0}}\bigg{(}\sum_{j\in{\mathbb{N}}}\alpha_{j}\bigg{)}^{k}.

From this we deduce that (𝜶𝝂|𝝂|!/𝝂!)𝝂({\boldsymbol{\alpha}}^{\boldsymbol{\nu}}|{\boldsymbol{\nu}}|!/{\boldsymbol{\nu}}!)_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to 1()\ell^{1}({\mathcal{F}}) if and only if 𝜶1(){\boldsymbol{\alpha}}\in\ell^{1}({\mathbb{N}}) and 𝜶1<1\|{\boldsymbol{\alpha}}\|_{\ell^{1}}<1.

Suppose that

(𝜶𝝂|𝝂|!/𝝂!)𝝂p()({\boldsymbol{\alpha}}^{\boldsymbol{\nu}}|{\boldsymbol{\nu}}|!/{\boldsymbol{\nu}}!)_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}})

for some p(0,1)p\in(0,1). As in Step 1, the sequence (αj)j=(𝜶𝒆j)j(\alpha_{j})_{j\in\mathbb{N}}=({\boldsymbol{\alpha}}^{{\boldsymbol{e}}_{j}})_{j\in{\mathbb{N}}} and (αjn)n0=(𝜶n𝒆j)n0(\alpha_{j}^{n})_{n\in{\mathbb{N}}_{0}}=({\boldsymbol{\alpha}}^{n{\boldsymbol{e}}_{j}})_{n\in{\mathbb{N}}_{0}} are subsequences of (𝜶𝝂)𝝂({\boldsymbol{\alpha}}^{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}. Therefore (𝜶𝝂|𝝂|!/𝝂!)𝝂({\boldsymbol{\alpha}}^{\boldsymbol{\nu}}|{\boldsymbol{\nu}}|!/{\boldsymbol{\nu}}!)_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to p()\ell^{p}({\mathcal{F}}) implies that 𝜶p<\|{\boldsymbol{\alpha}}\|_{\ell^{p}}<\infty and 𝜶1<1\|{\boldsymbol{\alpha}}\|_{\ell^{1}}<1.

Conversely, assume that 𝜶p<\|{\boldsymbol{\alpha}}\|_{\ell^{p}}<\infty and 𝜶1<1\|{\boldsymbol{\alpha}}\|_{\ell^{1}}<1. We put δ:=1𝜶1>0\delta:=1-\|{\boldsymbol{\alpha}}\|_{\ell^{1}}>0 and η:=δ3\eta:=\frac{\delta}{3}. Take JJ large enough such that j>Jαjpη\sum_{j>J}\alpha_{j}^{p}\leq\eta. We define the sequence 𝒄{\boldsymbol{c}} and 𝒅{\boldsymbol{d}} by

cj=(1+η)αj,dj=11+ηc_{j}=(1+\eta)\alpha_{j},\qquad d_{j}=\frac{1}{1+\eta}

if jJj\leq J and

cj=αjp,dj=αj1pc_{j}=\alpha_{j}^{p},\qquad d_{j}=\alpha_{j}^{1-p}

if j>Jj>J. By this construction we have αj=cjdj\alpha_{j}=c_{j}d_{j} for all jj\in{\mathbb{N}}. For the sequence 𝒄{\boldsymbol{c}} we have

𝒄1(1+η)𝜶1+j>Jαjp(1+η)(1δ)+η<1η.\|{\boldsymbol{c}}\|_{\ell^{1}}\leq(1+\eta)\|{\boldsymbol{\alpha}}\|_{\ell^{1}}+\sum_{j>J}\alpha_{j}^{p}\leq(1+\eta)(1-\delta)+\eta<1-\eta.

Next we show that 𝒅<1\|{\boldsymbol{d}}\|_{\ell^{\infty}}<1. Indeed, for 1jJ1\leq j\leq J we have dj=11+η<1d_{j}=\frac{1}{1+\eta}<1 and for j>Jj>J we have

dj=(αjp)(1p)/pη(1p)/p<1.d_{j}=(\alpha_{j}^{p})^{(1-p)/p}\leq\eta^{(1-p)/p}<1.

Moreover since dj(p/(1p))=αjpd_{j}^{(p/(1-p))}=\alpha_{j}^{p} for j>Jj>J we have 𝒅p/(1p)(){\boldsymbol{d}}\in\ell^{p/(1-p)}(\mathbb{N}). Now we get from Hölder’s inequality

𝝂(|𝝂|𝝂!𝜶𝝂)p=𝝂(|𝝂|𝝂!𝒄𝝂)p𝒅p𝝂(𝝂|𝝂|𝝂!𝒄𝝂)p(𝝂𝒅𝝂p/(1p))1p.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\bigg{(}\frac{|{\boldsymbol{\nu}}|}{{\boldsymbol{\nu}}!}{\boldsymbol{\alpha}}^{\boldsymbol{\nu}}\bigg{)}^{p}=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\bigg{(}\frac{|{\boldsymbol{\nu}}|}{{\boldsymbol{\nu}}!}{\boldsymbol{c}}^{\boldsymbol{\nu}}\bigg{)}^{p}{\boldsymbol{d}}^{p{\boldsymbol{\nu}}}\leq\bigg{(}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\frac{|{\boldsymbol{\nu}}|}{{\boldsymbol{\nu}}!}{\boldsymbol{c}}^{\boldsymbol{\nu}}\bigg{)}^{p}\bigg{(}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}{\boldsymbol{d}}^{{\boldsymbol{\nu}}p/(1-p)}\bigg{)}^{1-p}.

Observe that the first factor on the right side is finite since 𝒄1(){\boldsymbol{c}}\in\ell^{1}({\mathbb{N}}) and 𝒄1<1\|{\boldsymbol{c}}\|_{\ell^{1}}<1. Applying the first statement, the second factor on the right side is finite, whence (𝜶𝝂|𝝂|!/𝝂!)𝝂p()({\boldsymbol{\alpha}}^{\boldsymbol{\nu}}|{\boldsymbol{\nu}}|!/{\boldsymbol{\nu}}!)_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}). ∎

With these sequence summability results at hand, we are now in position to formulate Wiener-Hermite summation results for parametric solution families of PDEs with log-Gaussian random field data.

Theorem 3.13 (General case).

Let Assumption 3.6 hold and assume that ϱ=(ϱj)j[0,){\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} is a sequence satisfying (ϱj1)jq()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}) for some 0<q<0<q<\infty. Assume that, for each 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}}, there exists a sequence 𝛒𝛎=(ρ𝛎,j)j[0,){\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} such that supp(𝛎)supp(𝛒𝛎)\operatorname{supp}({\boldsymbol{\nu}})\subseteq\operatorname{supp}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}),

sup𝝂jρ𝝂,j|ψj|Lκ<π2,and𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂<\sup_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2},\qquad\text{and}\qquad\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty (3.40)

with rr\in{\mathbb{N}}, r>2/qr>2/q. Then

𝝂β𝝂(r,ϱ)u𝝂V2<with(β𝝂(r,ϱ)1/2)𝝂q().\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}<\infty\ \ \ with\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}}). (3.41)

Furthermore,

(u𝝂V)𝝂p()with1p=1q+12.(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}})\ \ \ with\ \ \ \frac{1}{p}=\frac{1}{q}+\frac{1}{2}.
Proof.

By Proposition 3.7 Assumption 3.6 implies that b(𝒚)b({\boldsymbol{y}}) belongs to L(D)L^{\infty}({{D}}) for γ\gamma-a.e. 𝒚U{\boldsymbol{y}}\in U and 𝔼(exp(kb(𝒚)L))\mathbb{E}(\exp(k\|b({\boldsymbol{y}})\|_{L^{\infty}})) is finite for all k[0,)k\in[0,\infty).

For 𝒚U{\boldsymbol{y}}\in U such that b(𝒚)L(D)b({\boldsymbol{y}})\in L^{\infty}({{D}}) and 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} with 𝔲=supp(𝝂){\mathfrak{u}}=\operatorname{supp}({\boldsymbol{\nu}}), the solution uu of (3.17) is holomorphic in 𝒮𝔲(𝝆𝝂)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}), see Proposition 3.8. This, (3.40) and Lemmata 3.9 and 3.10 yield that

𝝂β𝝂(r,ϱ)u𝝂V2=𝝂rϱ2𝝂𝝂!U𝝂u(𝒚)V2dγ(𝒚)C02𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂𝔼(exp(2b(𝒚)L))<.\begin{split}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}&=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{V}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\\ &\leq C_{0}^{2}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}\mathbb{E}\left(\exp\big{(}2\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}\right)<\infty.\end{split}

Since r>2qr>\frac{2}{q} and (ϱj1)jq()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}), by Lemma 3.11 the family (β𝝂(r,ϱ)1/2)𝝂\big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to q()\ell^{q}({\mathcal{F}}). The relation (3.41) is proven.

From (3.41), by Hölder’s inequality we get that

𝝂u𝝂Vp(𝝂β𝝂(r,ϱ)u𝝂V2)p/2(𝝂β𝝂(r,ϱ)q/2)1p/2<.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\|u_{\boldsymbol{\nu}}\|_{V}^{p}\leq\Bigg{(}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}\Bigg{)}^{p/2}\Bigg{(}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}})^{-q/2}\Bigg{)}^{1-p/2}<\infty\,.

Corollary 3.14 (The case of global supports).

Assume that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψjL)j1()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}})\ \ \mbox{and}\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<0<q<\infty. Then we have (u𝛎V)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

Proof.

Let 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}}. We define the sequence 𝝆𝝂=(ρ𝝂,j)j{\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}} by ρ𝝂,j:=𝝂j|𝝂|ψjL\rho_{{\boldsymbol{\nu}},j}:=\frac{{\boldsymbol{\nu}}_{j}}{|{\boldsymbol{\nu}}|\|\psi_{j}\|_{L^{\infty}}} for jsupp(𝝂)j\in\operatorname{supp}({\boldsymbol{\nu}}) and ρ𝝂,j=0\rho_{{\boldsymbol{\nu}},j}=0 if jsupp(𝝂)j\not\in\operatorname{supp}({\boldsymbol{\nu}}) and choose ϱ=τ𝝀{\boldsymbol{\varrho}}=\tau{\boldsymbol{\lambda}}, τ\tau is an appropriate positive constant. It is obvious that

sup𝝂jρ𝝂,j|ψj|L1.\sup_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}\big{|}\psi_{j}\big{|}\Bigg{\|}_{L^{\infty}}\leq 1.

We first show that Assumption 3.6 is satisfied for the sequence 𝝀=(λj)j{\boldsymbol{\lambda}}^{\prime}=(\lambda^{\prime}_{j})_{j\in{\mathbb{N}}} with λj:=λj1/2\lambda_{j}^{\prime}:=\lambda_{j}^{1/2} by a similar argument as in [9, Remark 2.5]. From the assumption (λj1)jq()(\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}) we derive that up to a nondecreasing rearrangement, λjCj1/(2q)\lambda^{\prime}_{j}\geq Cj^{1/(2q)} for some C>0C>0. Therefore, (exp(λj2))j1()\big{(}\exp(-{\lambda^{\prime}}_{j}^{2})\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}}). The convergence in L(D)L^{\infty}({{D}}) of jλj|ψj|\sum_{j\in{\mathbb{N}}}\lambda^{\prime}_{j}|\psi_{j}| can be proved as follows.

jλj|ψj|Lsupjλj1/2jλjψjL<.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\lambda^{\prime}_{j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\sup_{j\in{\mathbb{N}}}\lambda_{j}^{-1/2}\sum_{j\in{\mathbb{N}}}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}<\infty.

With r>2/qr>2/q we have

𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂𝝂r|𝝂|2|𝝂|𝝂2𝝂jsupp(𝝂)(τr!λjψjL)2νj(𝝂r|𝝂||𝝂|𝝂𝝂jsupp(𝝂)(τr!λjψjL)νj)2(𝝂r|𝝂|!𝝂!jsupp(𝝂)(eτr!λjψjL)νj)2.\begin{split}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}&\leq\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{|{\boldsymbol{\nu}}|^{2|{\boldsymbol{\nu}}|}}{{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\big{(}\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}^{2\nu_{j}}\\ &\leq\Bigg{(}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}}{{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\big{(}\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}^{\nu_{j}}\Bigg{)}^{2}\\ &\leq\Bigg{(}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{|{\boldsymbol{\nu}}|!}{{\boldsymbol{\nu}}!}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\big{(}e\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}^{\nu_{j}}\Bigg{)}^{2}.\end{split} (3.42)

In the last step we used the inequality

|𝝂||𝝂|𝝂𝝂e|𝝂||𝝂|!𝝂!,\frac{|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}}{{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\leq\frac{e^{|{\boldsymbol{\nu}}|}|{\boldsymbol{\nu}}|!}{{\boldsymbol{\nu}}!},

which is immediately derived from the inequalities m!mmemm!m!\leq m^{m}\leq e^{m}m!. Since (τr!λjψjL)j1()\big{(}\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}}), we can choose a positive number τ\tau so that

(eτr!λjψjL)j1<1.\big{\|}\big{(}e\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}_{j\in{\mathbb{N}}}\big{\|}_{\ell^{1}}<1.

This implies by Lemma 3.12(ii) that the last sum in (3.42) is finite. Applying Theorem 3.13 the desired result follows.

Corollary 3.15 (The case of disjoint supports).

Assuming ψjL(D)\psi_{j}\in L^{\infty}({{D}}) for all jj\in{\mathbb{N}} with disjoint supports and, furthermore, that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψjL)j2()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{2}({\mathbb{N}})\ \ {\rm and}\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<.0<q<\infty. Then (u𝛎V)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

Proof.

Fix 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}}, arbitrary. For this 𝝂{\boldsymbol{\nu}} we define the sequence 𝝆𝝂=(ρj)j{\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{j})_{j\in{\mathbb{N}}} by ρj:=1ψjL\rho_{j}:=\frac{1}{\|\psi_{j}\|_{L^{\infty}}} for jj\in{\mathbb{N}} and ϱ=τ𝝀{\boldsymbol{\varrho}}=\tau{\boldsymbol{\lambda}}, where a positive number τ\tau will be chosen later on. It is clear that

jρj|ψj|L1.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq 1.

Since (λjρj1)j2()\big{(}\lambda_{j}\rho_{j}^{-1}\big{)}_{j\in{\mathbb{N}}}\in\ell^{2}({\mathbb{N}}) and (λj1)jq()(\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}), by Hölder’s inequality we get (ρj1)jq0()(\rho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q_{0}}({\mathbb{N}}) with 1q0=12+1q\frac{1}{q_{0}}=\frac{1}{2}+\frac{1}{q}. Hence, similarly to the proof of Corollary 3.14, we can show that Assumption 3.6 holds for the sequence 𝝀=(λj)j{\boldsymbol{\lambda}}^{\prime}=(\lambda^{\prime}_{j})_{j\in{\mathbb{N}}} with λj:=λj1/2\lambda_{j}^{\prime}:=\lambda_{j}^{1/2}. In addition, with r>2/qr>2/q we have by Lemma 3.12(i)

𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂𝝂r(jsupp(𝝂)(τr!λjψjL)2νj)<,\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}\leq\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\Bigg{(}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\big{(}\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}^{2\nu_{j}}\Bigg{)}<\infty,

since by the condition (τr!λjψjL)j2()\big{(}\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{2}({\mathbb{N}}) a positive number τ\tau can be chosen so that supj(τr!λjψjL)<1\sup_{j\in{\mathbb{N}}}(\tau\sqrt{r!}\lambda_{j}\|\psi_{j}\|_{L^{\infty}})<1. Finally, we apply Theorem 3.13 to obtain the desired results.

Remark 3.16.

We comment on the situation when there exists 𝝆=(ρj)j(0,){\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}}\in(0,\infty)^{\infty} such that

jρj|ψj|L=κ<π2\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}=\kappa<\frac{\pi}{2}

and (ρj1)jq0()(\rho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q_{0}}({\mathbb{N}}) for some 0<q0<0<q_{0}<\infty as given in [9, Theorem 1.2]. We choose ϱ=(ϱj)j{\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}} by

ϱj=ρj1q0/21r!(ρj1)jq0q0/2\varrho_{j}=\rho_{j}^{1-q_{0}/2}\frac{1}{\sqrt{r!}\|(\rho_{j}^{-1})_{j\in{\mathbb{N}}}\|_{\ell^{q_{0}}}^{q_{0}/2}}

and 𝝆𝝂=(ρj)j{\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{j})_{j\in{\mathbb{N}}}. Then we obtain (ϱj1)jq0/(1q0/2)()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q_{0}/(1-q_{0}/2)}({\mathbb{N}}) and

𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂=𝝂r𝝂!jsupp𝝂(ρjq0r!(ρj1)jq0q0)νj<.\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}{\boldsymbol{\nu}}!\prod_{j\in\operatorname{supp}{\boldsymbol{\nu}}}\Bigg{(}\frac{\rho_{j}^{-q_{0}}}{r!\|(\rho_{j}^{-1})_{j\in{\mathbb{N}}}\|_{\ell^{q_{0}}}^{q_{0}}}\Bigg{)}^{\nu_{j}}<\infty.

This implies (u𝝂V)𝝂p()(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with p=q0p=q_{0}.

Remark 3.17.

The p\ell^{p}-summability (u𝝂V)𝝂p()(\|u_{\boldsymbol{\nu}}\|_{V})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) proven in Theorem 3.13, has been used in establishing the convergence rate of the best nn-term approximation of the solution uu to the parametric elliptic PDE (3.17) [9]. However, such a property cannot be used for estimating convergence rates of high-dimensional deterministic numerical approximation constructive schemes such as single-level and multi-level versions of anisotropic sparse-grid Hermite-Smolyak interpolation and quadrature in Section 7. In the latter situation, the weighted 2\ell^{2}-summability presented in Theorem 3.13

𝝂β𝝂(r,ϱ)u𝝂V2<with(β𝝂(r,ϱ)1/2)𝝂q()\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{V}^{2}<\infty\ \ \ \text{with}\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}})

and its generalization

𝝂(σ𝝂u𝝂X)2<with(p𝝂(τ,λ)σ𝝂1)𝝂q()\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}(\sigma_{\boldsymbol{\nu}}\|u_{\boldsymbol{\nu}}\|_{X})^{2}<\infty\quad\ \ \ \text{with}\ \ \ \big{(}p_{{\boldsymbol{\nu}}}(\tau,\lambda)\sigma_{\boldsymbol{\nu}}^{-1}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}}) (3.43)

for a Hilbert space XX are efficiently applied, where 0<q<0<q<\infty, λ,τ0\lambda,\tau\geq 0, (σ𝝂)𝝂\big{(}\sigma_{\boldsymbol{\nu}}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}} is a family of positive numbers and

p𝝂(τ,λ):=j(1+λνj)τ,𝝂.p_{\boldsymbol{\nu}}(\tau,\lambda):=\prod_{j\in{\mathbb{N}}}(1+\lambda\nu_{j})^{\tau},\quad{\boldsymbol{\nu}}\in{\mathcal{F}}.

Weighted summability properties such as (3.43) have been employed in [31, 43, 45, 52] for particular questions in quadrature and interpolation with respect to GMs.

In Sections 6 and 7, we will see that weighted 2\ell^{2}-summabilities of the form (3.43) play a basic role in constructing approximation algorithms of sparse-grid interpolation and quadrature and in establishing their convergence rates.

3.7 Parametric Hs(D)H^{s}({D})-analyticity and sparsity

Whereas the previous results were, in principle, already known from the real-variable analyses in [16, 9, 8], in this and the subsequent sections, we prove via analytic continuation the sparsity of the Wiener-Hermite PC expansion coefficients of the parametric solutions of (3.17) with log-Gaussian coefficient a(𝒚)=exp(b(𝒚))a({\boldsymbol{y}})=\exp(b({\boldsymbol{y}})) when the Wiener-Hermite PC expansion coefficients of the parametric solution family {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} are measured in higher Sobolev norms. In Section 3.8 we shall establish corresponding results when the physical domain D{D} is a plane Lipschitz polygon whose sides are analytic arcs.

3.7.1 Hs(D)H^{s}({D})-analyticity

As Hs(D)H^{s}({D}) regularity in D{D} is relevant in particular in conjunction with Galerkin discretization in D{D} by continuous, piecewise polynomial, Lagrangian FEM, we review the elementary regularity results from Section 3.3. To state them, we recall the Sobolev spaces Hs(D)H^{s}({D}), and Ws(D)W^{s}_{\infty}({D}) of functions vv on D{D} for s0s\in{\mathbb{N}}_{0}, equipped with the respective norms

vHs:=𝒌+d:|𝒌|sD𝒌vL2,vWs:=𝒌+d:|𝒌|sD𝒌vL.\|v\|_{H^{s}}:=\ \sum_{{\boldsymbol{k}}\in{\mathbb{Z}}^{d}_{+}:|{\boldsymbol{k}}|\leq s}\|D^{\boldsymbol{k}}v\|_{L^{2}},\qquad\|v\|_{W^{s}_{\infty}}:=\ \sum_{{\boldsymbol{k}}\in{\mathbb{Z}}^{d}_{+}:|{\boldsymbol{k}}|\leq s}\|D^{\boldsymbol{k}}v\|_{L^{\infty}}.

With these definitions H0(D)=L2(D)H^{0}({D})=L^{2}({{D}}) and W0(D)=L(D)W^{0}_{\infty}({{D}})=L^{\infty}({{D}}). We recall from Section 3.3 that we identify L2(D)L^{2}({{D}}) with its own dual, so that the space H1(D)H^{-1}({{D}}) is defined as the dual of H01(D)H^{1}_{0}({{D}}) with respect to the pivot space L2(D)L^{2}({{D}}).

Lemma 3.18.

Let ss\in{\mathbb{N}} and D{D} be a bounded domain in d{\mathbb{R}}^{d} with either CC^{\infty}-boundary or with convex Cs1C^{s-1}-boundary. Assume that there holds the ellipticity condition (3.25), aWs1(D)a\in W^{s-1}_{\infty}({{D}}) and fHs2(D)f\in H^{s-2}({{D}}). Then the solution uu of (3.1) belongs to Hs(D)H^{s}({D}) and there holds

uHs{fH1ρ(a)s=1,Cd,sρ(a)(fHs2+aWs1uHs1)s>1,\|u\|_{H^{s}}\ \leq\ \begin{cases}\frac{\|f\|_{H^{-1}}}{\rho(a)}&\ s=1,\\[4.30554pt] \frac{C_{d,s}}{\rho(a)}\big{(}\|f\|_{H^{s-2}}+\|a\|_{W^{s-1}_{\infty}}\|u\|_{H^{s-1}}\big{)}&\ s>1,\end{cases} (3.44)

with Cd,sC_{d,s} depending on d,sd,s, and ρ(a)\rho(a) given as in (3.25).

Proof.

Defining, for ss\in{\mathbb{N}}, H0s(D):=(HsH01)(D)H^{s}_{0}({{D}}):=(H^{s}\cap H^{1}_{0})({{D}}), since D{D} is a bounded domain in d{\mathbb{R}}^{d} with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary, we have the following norm equivalence

vHs{vH01,s=1,ΔvHs2,s>1,vH0s,\|v\|_{H^{s}}\ \asymp\ \begin{cases}\|v\|_{H^{1}_{0}},&s=1,\\[4.30554pt] \|\Delta v\|_{H^{s-2}},&s>1,\end{cases}\qquad\forall v\in H^{s}_{0}, (3.45)

see [61, Theorem 2.5.1.1]. The lemma for the case s=1s=1 and s=2s=2 is given in (3.4) and (3.9). We prove the case s>2s>2 by induction on ss. Suppose that the assertion holds true for all s<ss^{\prime}<s. We will prove it for ss. Let a 𝒌+d{\boldsymbol{k}}\in{\mathbb{Z}}^{d}_{+} with |𝒌|=s2|{\boldsymbol{k}}|=s-2 be given. Differentiating both sides of (3.10) and applying the Leibniz rule of multivariate differentiation we obtain

0𝒌𝒌(𝒌𝒌)D𝒌aD𝒌𝒌(Δu)=D𝒌f+0𝒌𝒌(𝒌𝒌)(D𝒌a,D𝒌𝒌u),-\sum_{0\leq{\boldsymbol{k}}^{\prime}\leq{\boldsymbol{k}}}\binom{{\boldsymbol{k}}}{{\boldsymbol{k}}^{\prime}}D^{{\boldsymbol{k}}^{\prime}}aD^{{\boldsymbol{k}}-{\boldsymbol{k}}^{\prime}}\big{(}\Delta u\big{)}\ =\ D^{\boldsymbol{k}}f+\sum_{0\leq{\boldsymbol{k}}^{\prime}\leq{\boldsymbol{k}}}\binom{{\boldsymbol{k}}}{{\boldsymbol{k}}^{\prime}}\big{(}\nabla D^{{\boldsymbol{k}}^{\prime}}a,\nabla D^{{\boldsymbol{k}}-{\boldsymbol{k}}^{\prime}}u\big{)},

see also [8, Lemma 4.3]. Hence,

aD𝒌Δu=D𝒌f+0𝒌𝒌(𝒌𝒌)(D𝒌a,D𝒌𝒌u)+0𝒌𝒌,𝒌0(𝒌𝒌)D𝒌aD𝒌𝒌Δu.-a\,D^{\boldsymbol{k}}\Delta u\,=\,D^{\boldsymbol{k}}f+\sum_{0\leq{\boldsymbol{k}}^{\prime}\leq{\boldsymbol{k}}}\binom{{\boldsymbol{k}}}{{\boldsymbol{k}}^{\prime}}\big{(}\nabla D^{{\boldsymbol{k}}^{\prime}}a,\nabla D^{{\boldsymbol{k}}-{\boldsymbol{k}}^{\prime}}u\big{)}+\sum_{0\leq{\boldsymbol{k}}^{\prime}\leq{\boldsymbol{k}},\,{\boldsymbol{k}}^{\prime}\not=0}\binom{{\boldsymbol{k}}}{{\boldsymbol{k}}^{\prime}}D^{{\boldsymbol{k}}^{\prime}}aD^{{\boldsymbol{k}}-{\boldsymbol{k}}^{\prime}}\Delta u.

Taking the L2L^{2}-norm of both sides, by the ellipticity condition (3.3) we derive the inequality

ρ(a)ΔuHs2Cd,s(fHs2+aWs1uHs1+aWs2ΔuHs3)\rho(a)\,\|\Delta u\|_{H^{s-2}}\ \leq\ C^{\prime}_{d,s}\,\big{(}\|f\|_{H^{s-2}}+\|a\|_{W^{s-1}_{\infty}}\,\|u\|_{H^{s-1}}+\|a\|_{W^{s-2}_{\infty}}\,\|\Delta u\|_{H^{s-3}}\big{)}

which yields (3.44) due to (3.45) and the inequality

aWs2ΔuHs3aWs1uHs1,\|a\|_{W^{s-2}_{\infty}}\,\|\Delta u\|_{H^{s-3}}\leq\|a\|_{W^{s-1}_{\infty}}\,\|u\|_{H^{s-1}},

where Cd,sC^{\prime}_{d,s} is a constant depending on d,sd,s only. By induction, this proves that uu belongs to HsH^{s}.

Corollary 3.19.

Let ss\in{\mathbb{N}} and D{D} be a bounded domain in d{\mathbb{R}}^{d} with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Assume that there holds the ellipticity condition (3.3), aWs1(D)a\in W^{s-1}_{\infty}({{D}}) and fHs2(D)f\in H^{s-2}({{D}}). Then the solution uu of (3.1) belongs to Hs(D)H^{s}({{D}}) and there holds the estimate

uHsfHs2ρ(a){1,s=1,Cd,s(1+aWs1ρ(a))s1,s>1,\|u\|_{H^{s}}\ \leq\ \frac{\|f\|_{H^{s-2}}}{\rho(a)}\begin{cases}1,&\ s=1,\\[4.30554pt] C_{d,s}\Big{(}1+\frac{\|a\|_{W^{s-1}_{\infty}}}{{\rho(a)}}\Big{)}^{s-1},&\ s>1,\end{cases}

where Cd,sC_{d,s} is a constant depending on d,sd,s only, and ρ(a)\rho(a) is given as in (3.25).

We need the following lemma.

Lemma 3.20.

Let ss\in{\mathbb{N}} and assume that b(𝐲)b({\boldsymbol{y}}) belongs to Ws(D)W^{s}_{\infty}({{D}}). Then we have

a(𝒚)WsCa(𝒚)L(1+b(𝒚)Ws)s,\|a({\boldsymbol{y}})\|_{W^{s}_{\infty}}\leq C\|a({\boldsymbol{y}})\|_{L^{\infty}}\big{(}1+\|b({\boldsymbol{y}})\|_{W^{s}_{\infty}}\big{)}^{s}\,,

where the constant CC depends on ss and mm but is independent of 𝐲{\boldsymbol{y}}.

Proof.

For 𝜶=(α1,,αd)0d{\boldsymbol{\alpha}}=(\alpha_{1},\ldots,\alpha_{d})\in{\mathbb{N}}_{0}^{d} with 1|𝜶|s1\leq|{\boldsymbol{\alpha}}|\leq s, we observe that for αj>0\alpha_{j}>0 the product rule implies

D𝜶a(𝒚)=D𝜶𝒆j[a(𝒚)D𝒆jb(𝒚)]=0𝜸𝜶𝒆j(𝜶𝒆j𝜸)D𝜶𝜸b(𝒚)D𝜸a(𝒚).D^{\boldsymbol{\alpha}}a({\boldsymbol{y}})=D^{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\big{[}a({\boldsymbol{y}})D^{{\boldsymbol{e}}_{j}}b({\boldsymbol{y}})\big{]}=\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\binom{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}{{\boldsymbol{\gamma}}}D^{{\boldsymbol{\alpha}}-{\boldsymbol{\gamma}}}b({\boldsymbol{y}})D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\,. (3.46)

Here, we recall that (𝒆j)j=1d({\boldsymbol{e}}_{j})_{j=1}^{d} is the standard basis of d\mathbb{R}^{d}. Taking norms on both sides, we can estimate

D𝜶a(𝒚)L=D𝜶𝒆j[a(𝒚)D𝒆jb(𝒚)]L0𝜸𝜶𝒆j(𝜶𝒆j𝜸)D𝜶𝜸b(𝒚)LD𝜸a(𝒚)LC(0𝜸𝜶𝒆jD𝜸a(𝒚)L)(|𝒌|sD𝒌b(𝒚)L).\begin{split}\|D^{\boldsymbol{\alpha}}a({\boldsymbol{y}})\|_{L^{\infty}}&=\big{\|}D^{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\big{[}a({\boldsymbol{y}})D^{{\boldsymbol{e}}_{j}}b({\boldsymbol{y}})\big{]}\big{\|}_{L^{\infty}}\\ &\leq\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\binom{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}{{\boldsymbol{\gamma}}}\|D^{{\boldsymbol{\alpha}}-{\boldsymbol{\gamma}}}b({\boldsymbol{y}})\|_{L^{\infty}}\|D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\\ &\leq C\Bigg{(}\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\|D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\Bigg{(}\sum_{|{\boldsymbol{k}}|\leq s}\|D^{\boldsymbol{k}}b({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\,.\end{split}

Similarly, each term D𝜸a(𝒚)L\|D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}} with |𝜸|>0|{\boldsymbol{\gamma}}|>0 can be estimated

D𝜸a(𝒚)LC(0𝜸𝜸𝒆jD𝜸a(𝒚)L)(|𝒌|sD𝒌b(𝒚)L)\|D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\leq C\Bigg{(}\sum_{0\leq{\boldsymbol{\gamma}}^{\prime}\leq{\boldsymbol{\gamma}}-{\boldsymbol{e}}_{j}}\|D^{{\boldsymbol{\gamma}}^{\prime}}a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\Bigg{(}\sum_{|{\boldsymbol{k}}|\leq s}\|D^{\boldsymbol{k}}b({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\,

if γj>0\gamma_{j}>0. This implies

D𝜶a(𝒚)LCa(𝒚)L(1+|𝒌|sD𝒌b(𝒚)L)|𝜶|,\begin{split}\|D^{\boldsymbol{\alpha}}a({\boldsymbol{y}})\|_{L^{\infty}}\leq C\|a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{(}1+\sum_{|{\boldsymbol{k}}|\leq s}\|D^{\boldsymbol{k}}b({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}^{|{\boldsymbol{\alpha}}|}\,,\end{split}

for 1|𝜶|s1\leq|{\boldsymbol{\alpha}}|\leq s. Summing up these terms with a(𝒚)L\|a({\boldsymbol{y}})\|_{L^{\infty}} we obtain the desired estimate.

Proposition 3.21.

Let ss\in{\mathbb{N}} and D{D} be a bounded domain in d{\mathbb{R}}^{d} with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Assume that (3.28) holds and all the functions ψj\psi_{j} belong to Ws1(D)W^{s-1}_{\infty}({{D}}). Let 𝔲supp(𝛒){\mathfrak{u}}\subseteq\operatorname{supp}({\boldsymbol{\rho}}) be a finite set and let 𝐲0=(y0,1,y0,2,)U{\boldsymbol{y}}_{0}=(y_{0,1},y_{0,2},\ldots)\in U be such that b(𝐲0)b({\boldsymbol{y}}_{0}) belongs to Ws1(D)W^{s-1}_{\infty}({{D}}). Then the solution uu of (3.17) is holomorphic in 𝒮𝔲(𝛒)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}) as a function in variables 𝐳𝔲=(zj)j𝒮𝔲(𝐲0,𝛒){\boldsymbol{z}}_{{\mathfrak{u}}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) taking values in Hs(D)H^{s}({{D}}) where zj=y0,jz_{j}=y_{0,j} for j𝔲j\not\in{\mathfrak{u}} held fixed .

Proof.

Let 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}) be given in (3.29) and 𝒛𝔲=(yj+iξj)j𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{{\mathfrak{u}}}=(y_{j}+{\rm i}\xi_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) with (yj+iξj)j𝔲𝒮𝔲,N(𝝆)(y_{j}+{\rm i}\xi_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}). Then we have from Corollary 3.19

u(𝒛𝔲)HsCρ(a(𝒛𝔲))(1+ρ(a(𝒛𝔲))a(𝒛𝔲)Ws1)s1.\begin{split}\|u({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{H^{s}}&\leq C\rho(a({\boldsymbol{z}}_{{\mathfrak{u}}}))\Big{(}1+\rho(a({\boldsymbol{z}}_{{\mathfrak{u}}}))\|a({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{W^{s-1}_{\infty}}\Big{)}^{s-1}\,.\end{split}

Using Lemma 3.20 we find

a(𝒛𝔲)Ws1Ca(𝒛𝔲)L(1+b(𝒛𝔲)Ws1)s1Ca(𝒛𝔲)L(1+b(𝒚0)Ws1+j𝔲(yjy0,j+iξj)ψjWs1)s1Ca(𝒛𝔲)L(1+b(𝒚0)Ws1+j𝔲(N+ρj)ψjWs1)s1\begin{split}\|a({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{W_{\infty}^{s-1}}&\leq C\|a({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{L^{\infty}}\big{(}1+\|b({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{W_{\infty}^{s-1}}\big{)}^{s-1}\\ &\leq C\|a({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{L^{\infty}}\Bigg{(}1+\|b({\boldsymbol{y}}_{0})\|_{W_{\infty}^{s-1}}+\Bigg{\|}\sum_{j\in{\mathfrak{u}}}(y_{j}-y_{0,j}+{\rm i}\xi_{j})\psi_{j}\Bigg{\|}_{W_{\infty}^{s-1}}\Bigg{)}^{s-1}\\ &\leq C\|a({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{L^{\infty}}\Bigg{(}1+\|b({\boldsymbol{y}}_{0})\|_{W_{\infty}^{s-1}}+\sum_{j\in{\mathfrak{u}}}(N+\rho_{j})\|\psi_{j}\|_{W_{\infty}^{s-1}}\Bigg{)}^{s-1}\end{split}

and

a(𝒛𝔲)Lexp(b(𝒚0)L+j𝔲(yjy0,j+iξj)ψjL)<.\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}\leq\exp\Bigg{(}\|b({\boldsymbol{y}}_{0})\|_{L^{\infty}}+\Bigg{\|}\sum_{j\in{\mathfrak{u}}}(y_{j}-y_{0,j}+{\rm i}\xi_{j})\psi_{j}\Bigg{\|}_{L^{\infty}}\Bigg{)}<\infty. (3.47)

From this and (3.30) we obtain

u(𝒛𝔲)HsC<\|u({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{H^{s}}\leq C<\infty

which implies the map 𝒛𝔲u(𝒛𝔲){\boldsymbol{z}}_{\mathfrak{u}}\to u({\boldsymbol{z}}_{{\mathfrak{u}}}) is holomorphic on the set 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}) as a consequence of [24, Lemma 2.2]. For more details we refer the reader to [111, Examples 1.2.38 and 1.2.39]. Since NN is arbitrary we conclude that the map 𝒛𝔲u(𝒛𝔲){\boldsymbol{z}}_{\mathfrak{u}}\to u({\boldsymbol{z}}_{{\mathfrak{u}}}) is holomorphic on 𝒮𝔲(𝝆)\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{\rho}}). ∎

3.7.2 Sparsity of Wiener-Hermite PC expansion coefficients

For sparsity of HsH^{s}-norms of Wiener-Hermite PC expansion coefficients we need the following assumption.

Assumption 3.22.

Let ss\in{\mathbb{N}}. For every jj\in{\mathbb{N}}, ψjWs1(D)\psi_{j}\in W^{s-1}_{\infty}({{D}}) and there exists a positive sequence (λj)j(\lambda_{j})_{j\in{\mathbb{N}}} such that (exp(λj2))j1()\big{(}\exp(-\lambda_{j}^{2})\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}}) and the series

jλj|D𝜶ψj|\sum_{j\in{\mathbb{N}}}\lambda_{j}|D^{{\boldsymbol{\alpha}}}\psi_{j}|

converges in L(D)L^{\infty}({{D}}) for all 𝛂0d{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{d} with |𝛂|s1|{\boldsymbol{\alpha}}|\leq s-1.

As a consequence of [9, Theorem 2.2] we have the following

Lemma 3.23.

Let Assumption 3.22 hold. Then the set Us1:={𝐲U:b(𝐲)Ws1(D)}U_{s-1}:=\{{\boldsymbol{y}}\in U:b({\boldsymbol{y}})\in W_{\infty}^{s-1}({D})\} has full measure, i.e., γ(Us1)=1\gamma(U_{s-1})=1. Furthermore, 𝔼(exp(kb()Ws1))\mathbb{E}(\exp(k\|b(\cdot)\|_{W_{\infty}^{s-1}})) is finite for all k[0,).k\in[0,\infty).

The HsH^{s}-analytic continuation of the parametric solutions {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} to 𝒮𝔲(𝝆)\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{\rho}}) leads to the following result on parametric HsH^{s}-regularity.

Lemma 3.24.

Let Dd{D}\subset{\mathbb{R}}^{d} be a bounded domain with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Assume that for each 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}}, there exists a sequence 𝛒𝛎=(ρ𝛎,j)j[0,){\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} such that supp(𝛎)supp(𝛒𝛎)\operatorname{supp}({\boldsymbol{\nu}})\subseteq\operatorname{supp}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}), and such that

sup𝝂|𝜶|s1jρ𝝂,j|D𝜶ψj|Lκ<π2.\sup_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|D^{{\boldsymbol{\alpha}}}\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2}.

Then we have

𝝂u(𝒚)HsC𝝂!𝝆𝝂𝝂exp(b(𝒚)L){1+exp(2b(𝒚)L)(1+b(𝒚)Ws1)s1}s1,\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{H^{s}}\leq C\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\exp\big{(}\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}\Big{\{}1+\exp(2\|b({\boldsymbol{y}})\|_{L^{\infty}})\big{(}1+\|b({\boldsymbol{y}})\|_{W_{\infty}^{s-1}}\big{)}^{s-1}\Big{\}}^{s-1}, (3.48)

where CC is a constant depending on κ\kappa, dd, ss only.

Proof.

Let 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} with 𝔲=supp(𝝂){\mathfrak{u}}=\operatorname{supp}({\boldsymbol{\nu}}) and 𝒚U{\boldsymbol{y}}\in U such that b(𝒚)Ws1(D)b({\boldsymbol{y}})\in W_{\infty}^{s-1}({{D}}). Let furthermore 𝒞𝒚,𝔲(𝝆𝝂){\mathcal{C}}_{{\boldsymbol{y}},{\mathfrak{u}}}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}) and 𝒞𝔲(𝒚,𝝆𝝂){\mathcal{C}}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}}) be given as in (3.31) and (3.33). Using Cauchy’s formula as in the proof of Lemma 3.9 we obtain

𝝂u(𝒚)Hs𝝂!𝝆𝝂𝝂sup𝒛𝔲C𝔲(𝒚,𝝆𝝂)u(𝒛𝔲)Hs.\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{H^{s}}\leq\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\sup_{{\boldsymbol{z}}_{{\mathfrak{u}}}\in C_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}})}\|u({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{H^{s}}\,. (3.49)

For 𝒛𝔲=(zj)jC𝔲(𝒚,𝝆𝝂){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in C_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}}) we can write zj=yj+ηj+iξj𝒞𝒚,j(𝝆𝝂)z_{j}=y_{j}+\eta_{j}+{\rm i}\xi_{j}\in{\mathcal{C}}_{{\boldsymbol{y}},j}({\boldsymbol{\rho}}_{{\boldsymbol{\nu}}}) with |ηj|ρ𝝂,j|\eta_{j}|\leq\rho_{{\boldsymbol{\nu}},j} and |ξj|ρ𝝂,j|\xi_{j}|\leq\rho_{{\boldsymbol{\nu}},j} for j𝔲j\in{\mathfrak{u}} and hence we get

D𝜶b(𝒛𝔲)L=D𝜶(b(𝒚)+j𝔲(ηj+iξj)ψj)LD𝜶b(𝒚)L+2j𝔲ρ𝝂,j|D𝜶ψj|LD𝜶b(𝒚)L+κ2.\begin{split}\|D^{\boldsymbol{\alpha}}b({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}&=\Bigg{\|}D^{\boldsymbol{\alpha}}\Big{(}b({\boldsymbol{y}})+\sum_{j\in{\mathfrak{u}}}(\eta_{j}+{\rm i}\xi_{j})\psi_{j}\Big{)}\Bigg{\|}_{L^{\infty}}\\ &\leq\|D^{\boldsymbol{\alpha}}b({\boldsymbol{y}})\|_{L^{\infty}}+\sqrt{2}\,\Bigg{\|}\sum_{j\in{\mathfrak{u}}}\rho_{{\boldsymbol{\nu}},j}|D^{\boldsymbol{\alpha}}\psi_{j}|\Bigg{\|}_{L^{\infty}}\\ &\leq\|D^{\boldsymbol{\alpha}}b({\boldsymbol{y}})\|_{L^{\infty}}+\kappa\sqrt{2}\,.\end{split}

In addition we have

1ρ(a(𝒛𝔲))exp(b(𝒚+j𝔲ηjψjL)cos(j𝔲ξjψjL)exp(κ+b(𝒚)L)cosκ\frac{1}{\rho(a({\boldsymbol{z}}_{{\mathfrak{u}}}))}\leq\frac{\exp(\|b({\boldsymbol{y}}+\sum_{j\in{\mathfrak{u}}}\eta_{j}\psi_{j}\|_{L^{\infty}})}{\cos(\|\sum_{j\in{\mathfrak{u}}}\xi_{j}\psi_{j}\|_{L^{\infty}})}\leq\frac{\exp\big{(}\kappa+\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}}{\cos\kappa}\, (3.50)

and

a(𝒛𝔲)L=exp(b(𝒚)+j𝔲(ηj+iξj)ψj)Leκ2exp(b(𝒚)L).\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}=\Bigg{\|}\exp\Bigg{(}b({\boldsymbol{y}})+\sum_{j\in{\mathfrak{u}}}(\eta_{j}+{\rm i}\xi_{j})\psi_{j}\Bigg{)}\Bigg{\|}_{L^{\infty}}\leq e^{\kappa\sqrt{2}}\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\,. (3.51)

Consequently, we can bound

a(𝒛𝔲)Ws1Ca(𝒛𝔲)L(1+b(𝒛𝔲)Ws1)s1Cexp(b(𝒚)L)(1+b(𝒚)Ws1)s1.\begin{split}\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{W^{s-1}_{\infty}}&\leq C\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}\big{(}1+\|b({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{W_{\infty}^{s-1}}\big{)}^{s-1}\\ &\leq C\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\big{(}1+\|b({\boldsymbol{y}})\|_{W_{\infty}^{s-1}}\big{)}^{s-1}\,.\end{split}

Now Corollary 3.19 implies the inequality

sup𝒛𝔲C𝔲(𝒚,𝝆𝝂)u(𝒛𝔲)HsCexp(b(𝒚)L){1+exp(2b(𝒚)L)(1+b(𝒚)Ws1)s1}s1,\sup_{{\boldsymbol{z}}_{{\mathfrak{u}}}\in C_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}})}\|u({\boldsymbol{z}}_{\mathfrak{u}})\|_{H^{s}}\leq C\exp\big{(}\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}\Big{\{}1+\exp(2\|b({\boldsymbol{y}})\|_{L^{\infty}})\big{(}1+\|b({\boldsymbol{y}})\|_{W_{\infty}^{s-1}}\big{)}^{s-1}\Big{\}}^{s-1}, (3.52)

which together with (3.49) proves the lemma. ∎

We are now in position to formulate sparsity results for the HsH^{s}-norms of Wiener-Hermite PC expansion coefficients of the solution uu.

Theorem 3.25 (General case).

Let s,rs,r\in{\mathbb{N}} and Dd{D}\subset{\mathbb{R}}^{d} denote a bounded domain with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Let further Assumption 3.22 hold, assume that fHs2(D)f\in H^{s-2}({{D}}), and assume given a sequence ϱ=(ϱj)j(0,){\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}}\subset(0,\infty)^{\infty} that satisfies (ϱj1)jq()(\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}) for some 0<q<0<q<\infty. Assume in addition that, for each 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}}, there exists a sequence 𝛒𝛎=(ρ𝛎,j)j[0,){\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} such that supp(𝛎)supp(𝛒𝛎)\operatorname{supp}({\boldsymbol{\nu}})\subseteq\operatorname{supp}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}), and such that, with r>2/qr>2/q,

sup𝝂|𝜶|s1jρ𝝂,j|D𝜶ψj|Lκ<π2,and𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂<.\sup_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|D^{{\boldsymbol{\alpha}}}\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2},\quad\text{and}\quad\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty.

Then there holds, with β𝛎(r,ϱ)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) as in (3.36),

𝝂β𝝂(r,ϱ)u𝝂Hs2<with(β𝝂(r,ϱ)1/2)𝝂q().\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{H^{s}}^{2}<\infty\ \ \ with\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}})\,. (3.53)

Furthermore,

(u𝝂Hs)𝝂p()with1p=1q+12.(\|u_{\boldsymbol{\nu}}\|_{H^{s}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}})\ \ \ with\ \ \ \frac{1}{p}=\frac{1}{q}+\frac{1}{2}.
Proof.

Arguing as in the proof of [9, Theorem 3.3] we obtain that for any rr\in{\mathbb{N}} there holds following generalization of the Parseval-type identity

𝝂β𝝂(r,ϱ)u𝝂Hs2=𝝂rϱ2𝝂𝝂!U𝝂u(𝒚)Hs2dγ(𝒚).\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{H^{s}}^{2}=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{H^{s}}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\,. (3.54)

By (3.52), Lemma 3.23 and Hölder’s inequality we derive that

U(sup𝒛𝔲C𝔲(𝒚,𝝆𝝂)u(𝒛𝔲)Hs)2dγ(𝒚)C\int_{U}\bigg{(}\sup_{{\boldsymbol{z}}_{{\mathfrak{u}}}\in C_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}})}\|u({\boldsymbol{z}}_{{\mathfrak{u}}})\|_{H^{s}}\bigg{)}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\leq C

and in particular, 𝔼(u(𝒚)Hsk)\mathbb{E}(\|u({\boldsymbol{y}})\|_{H^{s}}^{k}) is finite for all k[0,)k\in[0,\infty). Now (3.54), Lemma 3.24 and our assumption give

𝝂β𝝂(r,ϱ)u𝝂Hs2=𝝂rϱ2𝝂𝝂!U𝝂u(𝒚)Hs2dγ(𝒚)C2𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂Udγ(𝒚)=C2𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂<,\begin{split}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{H^{s}}^{2}&=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{H^{s}}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\\ &\leq C^{2}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}\int_{U}\,\mathrm{d}\gamma({\boldsymbol{y}})=C^{2}\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty,\end{split}

where CC is the constant in (3.48). As in the proof of Theorem 3.13, by Lemma 3.11 the family (β𝝂(r,ϱ)1/2)𝝂\big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}} belongs to q()\ell^{q}({\mathcal{F}}). The relation (3.53) is proven.

The assertion (u𝝂Hs)𝝂p()(\|u_{\boldsymbol{\nu}}\|_{H^{s}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) can be proved in the same way as in the proof of Theorem 3.13. ∎

Similarly to Corollaries 3.14 and 3.15 from Theorem 3.25 we obtain

Corollary 3.26 (The case of global supports).

Let ss\in{\mathbb{N}} and Dd{D}\subset{\mathbb{R}}^{d} denote a bounded domain with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Assume that for all jj\in{\mathbb{N}} holds ψjWs1(D)\psi_{j}\in W^{s-1}_{\infty}({{D}}), and that fHs2(D)f\in H^{s-2}({{D}}). Assume further that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψjWs1)j1()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{W_{\infty}^{s-1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}})\ \ and\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<0<q<\infty. Then we have (u𝛎Hs)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{H^{s}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

Corollary 3.27 (The case of disjoint supports).

Let ss\in{\mathbb{N}} and Dd{D}\subset{\mathbb{R}}^{d} denote a bounded domain with either CC^{\infty}-boundary or convex Cs1C^{s-1}-boundary. Assume that fHs2(D)f\in H^{s-2}({{D}}) and for all jj\in{\mathbb{N}} holds ψjWs1(D)\psi_{j}\in W^{s-1}_{\infty}({{D}}) with disjoint supports. Assume further that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψjWs1)j2()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{W_{\infty}^{s-1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{2}({\mathbb{N}})\ \ and\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<0<q<\infty.

Then (u𝛎Hs)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{H^{s}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

3.8 Parametric Kondrat’ev analyticity and sparsity

In the previous section, we investigated the weighted 2\ell^{2}-summability and p\ell^{p}-summability of Wiener-Hermite PC expansion coefficients of parametric solutions measured in the standard Sobolev spaces Hs(D)H^{s}({D}). We assumed that Dd{D}\subset{\mathbb{R}}^{d} with boundary D\partial{D} of sufficient smoothness (depending on ss). In this section we consider in space dimension d=2d=2 the case when the physical domain D{D} is a polygonal domain. In such domains, elliptic regularity shift results and shift theorems in D{D} hold in Kondrat’ev spaces which are corner-weighted Sobolev spaces. We refer to [61, 90] and the references there for an extensive survey.

To state corresponding results for the log-Gaussian parametric elliptic problems, we first review definitions of the weighted Sobolev spaces of Kondrat’ev type and results from [24] on the holomorphy of parametric solutions in weighted Kondrat’ev spaces in polygonal domains D{D}. Then, we establish summability results of the coefficients of Wiener-Hermite PC expansions of the parametric solutions in Kondrat’ev spaces. FE approximation results for Wiener-Hermite PC expansion coefficient functions which are in these spaces were provided in Section 2.6.

3.8.1 Parametric Kϰs(D)K^{s}_{\varkappa}({{D}})-holomorphy

We recall the Kondrat’ev spaces in a bounded polygonal domain D{D} introduced in Section 2.6.1: for s0s\in{\mathbb{N}}_{0} and ϰ\varkappa\in{\mathbb{R}},

𝒦ϰs(D):={u:D:rD|𝜶|ϰD𝜶uL2(D),|𝜶|s}{\mathcal{K}}^{s}_{\varkappa}({D}):=\big{\{}u:{D}\to{\mathbb{C}}:\ r_{D}^{|{\boldsymbol{\alpha}}|-\varkappa}D^{\boldsymbol{\alpha}}u\in L^{2}({D}),|{\boldsymbol{\alpha}}|\leq s\big{\}}

and

𝒲s(D):={u:D:rD|𝜶|D𝜶uL(D),|𝜶|s}.{\mathcal{W}}^{s}_{\infty}({D}):=\big{\{}u:{D}\to{\mathbb{C}}:\ r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}u\in L^{\infty}({D}),\ |{\boldsymbol{\alpha}}|\leq s\big{\}}.

The weighted Sobolev norms in these spaces are given in Section 2.6.1.

Lemma 3.28.

Let s0s\in{\mathbb{N}}_{0}. Assume that 𝐲U{\boldsymbol{y}}\in U is such that b(𝐲)𝒲s(D)b({\boldsymbol{y}})\in{\mathcal{W}}^{s}_{\infty}({{D}}).

Then

a(𝒚)𝒲sCa(𝒚)L(1+b(𝒚)𝒲s)s,\|a({\boldsymbol{y}})\|_{{\mathcal{W}}^{s}_{\infty}}\leq C\|a({\boldsymbol{y}})\|_{L^{\infty}}\big{(}1+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s}_{\infty}}\big{)}^{s}\,,

where the constant CC depends on ss and mm.

Proof.

The proof proceeds along the lines of the proof of Lemma 3.20. Let 𝜶=(α1,,αd)0d{\boldsymbol{\alpha}}=(\alpha_{1},\ldots,\alpha_{d})\in{\mathbb{N}}_{0}^{d} with 1|𝜶|s1\leq|{\boldsymbol{\alpha}}|\leq s and recall that (𝒆j)j=1d({\boldsymbol{e}}_{j})_{j=1}^{d} is the standard basis of d\mathbb{R}^{d}. Assuming that αj>0\alpha_{j}>0 we have (3.46). We apply corner-weighted norms to both sides of (3.46). This implies

rD|𝜶|D𝜶a(𝒚)L=D𝜶𝒆j[a(𝒚)D𝒆jb(𝒚)]L0𝜸𝜶𝒆j(𝜶𝒆j𝜸)rD|𝜶𝜸|D𝜶𝜸b(𝒚)LrD|𝜸|D𝜸a(𝒚)LC(0𝜸𝜶𝒆jrD|𝜸|D𝜸a(𝒚)L)(|𝒌|srD|𝒌|D𝒌b(𝒚)L)=C(0𝜸𝜶𝒆jrD|𝜸|D𝜸a(𝒚)L)b(𝒚)𝒲s.\begin{split}\|r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}a({\boldsymbol{y}})\|_{L^{\infty}}&=\big{\|}D^{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\big{[}a({\boldsymbol{y}})D^{{\boldsymbol{e}}_{j}}b({\boldsymbol{y}})\big{]}\big{\|}_{L^{\infty}}\\ &\leq\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\binom{{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}{{\boldsymbol{\gamma}}}\|r_{D}^{|{\boldsymbol{\alpha}}-{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\alpha}}-{\boldsymbol{\gamma}}}b({\boldsymbol{y}})\|_{L^{\infty}}\|r_{D}^{|{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\\ &\leq C\Bigg{(}\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\|r_{D}^{|{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\Bigg{(}\sum_{|{\boldsymbol{k}}|\leq s}\|r_{D}^{|{\boldsymbol{k}}|}D^{\boldsymbol{k}}b({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\\ &=C\Bigg{(}\sum_{0\leq{\boldsymbol{\gamma}}\leq{\boldsymbol{\alpha}}-{\boldsymbol{e}}_{j}}\|r_{D}^{|{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s}_{\infty}}\,.\end{split}

Similarly, if γj>0\gamma_{j}>0, each term rD|𝜸|D𝜸a(𝒚)L\|r_{D}^{|{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}} with |𝜸|>0|{\boldsymbol{\gamma}}|>0 can be estimated

rD|𝜸|D𝜸a(𝒚)LC(0𝜸𝜸𝒆jrD|𝜸|D𝜸a(𝒚)L)b(𝒚)𝒲s.\|r_{D}^{|{\boldsymbol{\gamma}}|}D^{{\boldsymbol{\gamma}}}a({\boldsymbol{y}})\|_{L^{\infty}}\leq C\Bigg{(}\sum_{0\leq{\boldsymbol{\gamma}}^{\prime}\leq{\boldsymbol{\gamma}}-{\boldsymbol{e}}_{j}}\|r_{D}^{|{\boldsymbol{\gamma}}^{\prime}|}D^{{\boldsymbol{\gamma}}^{\prime}}a({\boldsymbol{y}})\|_{L^{\infty}}\Bigg{)}\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s}_{\infty}}.

This implies

rD|𝜶|D𝜶a(𝒚)LCa(𝒚)L(1+b(𝒚)𝒲s)|𝜶|,\begin{split}\|r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}a({\boldsymbol{y}})\|_{L^{\infty}}\leq C\|a({\boldsymbol{y}})\|_{L^{\infty}}\big{(}1+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s}_{\infty}}\big{)}^{|{\boldsymbol{\alpha}}|}\,,\end{split}

for 1|𝜶|s1\leq|{\boldsymbol{\alpha}}|\leq s. This finishes the proof.

We recall the following result from [24, Theorem 1].

Theorem 3.29.

Let D2{D}\subset\mathbb{R}^{2} be a polygonal domain, η0>0\eta_{0}>0, ss\in{\mathbb{N}} and Ns=2s+1s2N_{s}=2^{s+1}-s-2. Let aL(D,)a\in L^{\infty}({D},{\mathbb{C}}).

Then there exist τ\tau and CsC_{s} with the following property: for any a𝒲s1(D)a\in{\mathcal{W}}^{s-1}_{\infty}({{D}}) and for any ϰ\varkappa\in\mathbb{R} such that

|ϰ|<η:=min{η0,τ1aL1ρ(a)},|\varkappa|<\eta:=\min\{\eta_{0},\tau^{-1}\|a\|_{L^{\infty}}^{-1}\rho(a)\},

the operator PaP_{a} defined in (3.1) induces an isomorphism

Pa:𝒦ϰ+1s(D){u|D=0}𝒦ϰ1s2(D)P_{a}:{\mathcal{K}}_{\varkappa+1}^{s}({{D}})\cap\{u|_{\partial{D}}=0\}\to{\mathcal{K}}_{\varkappa-1}^{s-2}({{D}})

such that Pa1P_{a}^{-1} depends analytically on the coefficients aa and has norm

Pa1Cs(ρ(a)τ|ϰ|aL)Ns1a𝒲s1Ns.\|P_{a}^{-1}\|\leq C_{s}\big{(}\rho(a)-\tau|\varkappa|\|a\|_{L^{\infty}}\big{)}^{-N_{s}-1}\|a\|_{{\mathcal{W}}_{\infty}^{s-1}}^{N_{s}}\,.

The bound of τ\tau and CsC_{s} depends only on ss, D{D} and η0\eta_{0}.

Applying this result to our setting, we obtain the following parametric regularity.

Theorem 3.30.

Suppose η0>0\eta_{0}>0, ψj𝒲s1(D)\psi_{j}\in{\mathcal{W}}^{s-1}_{\infty}({{D}}) for all jj\in{\mathbb{N}} and that (3.28) holds. Let 𝔲supp(𝛒){\mathfrak{u}}\subseteq\operatorname{supp}({\boldsymbol{\rho}}) be a finite set. Let further 𝐲0=(y0,1,y0,2,)U{\boldsymbol{y}}_{0}=(y_{0,1},y_{0,2},\ldots)\in U be such that b(𝐲0)b({\boldsymbol{y}}_{0}) belongs to 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}). We denote

ϑ:=inf𝒛𝔲𝒮𝔲(𝒚0,𝝆)ρ(a(𝒛𝔲))a(𝒛𝔲)L1.\vartheta:=\inf_{{\boldsymbol{z}}_{\mathfrak{u}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}})}\rho\big{(}a({\boldsymbol{z}}_{\mathfrak{u}})\big{)}\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}^{-1}\,.

Let τ>0\tau>0 be as given in Theorem 3.29.

Then there exists a positive constant CsC_{s} such that for ϰ\varkappa\in\mathbb{R} with |ϰ|min{η0,τ1ϑ/2},|\varkappa|\leq\min\{\eta_{0},\tau^{-1}\vartheta/2\}, and for f𝒦s2ϰ1(D)f\in{\mathcal{K}}^{s-2}_{\varkappa-1}({{D}}), the solution uu of (3.17) is holomorphic in the cylinder 𝒮𝔲(𝛒)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}) as a function in variables 𝐳𝔲=(zj)j𝒮𝔲(𝐲0,𝛒){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) taking values in 𝒦ϰ+1s(D)V{\mathcal{K}}_{\varkappa+1}^{s}({{D}})\cap V, where zj=y0,jz_{j}=y_{0,j} for j𝔲j\not\in{\mathfrak{u}} held fixed. Furthermore, we have the estimate

u(𝒛𝔲)𝒦ϰ+1sCs1(ρ(a(𝒛𝔲))Ns+1a(𝒛𝔲)𝒲s1Ns.\|u({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{K}}_{\varkappa+1}^{s}}\leq C_{s}\frac{1}{\big{(}\rho(a({\boldsymbol{z}}_{\mathfrak{u}})\big{)}^{N_{s}+1}}\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}^{N_{s}}\,.
Proof.

Observe first that for the parametric coefficient a(𝒛𝔲)a({\boldsymbol{z}}_{\mathfrak{u}}), the conditions of Proposition 3.8 are satisfied.

Thus, the solution uu is holomorphic in 𝒮𝔲(𝝆)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}) as a VV-valued map in variables 𝒛𝔲=(zj)j𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}). We assume that ϑ>0\vartheta>0. Let 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}) be given in (3.29) and 𝒛𝔲=(yj+iξj)j𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{{\mathfrak{u}}}=(y_{j}+{\rm i}\xi_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) with (yj+iξj)j𝔲𝒮𝔲,N(𝝆)(y_{j}+{\rm i}\xi_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}). From Lemma 3.28 we have

a(𝒛𝔲)𝒲s1Ca(𝒛𝔲)L(1+b(𝒛𝔲)𝒲s1)s1.\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\leq C\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}\Big{(}1+\|b({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Big{)}^{s-1}\,.

Furthermore

b(𝒛𝔲)𝒲s1=|𝜶|s1rD|𝜶|j(yj+iξj)D𝜶ψjLj𝔲(|yjy0,j|+ρj)ψj𝒲s1+b(𝒚0)𝒲s1<.\begin{split}\|b({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}&=\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}r_{D}^{|{\boldsymbol{\alpha}}|}\sum_{j\in{\mathbb{N}}}(y_{j}+{\rm i}\xi_{j})D^{{\boldsymbol{\alpha}}}\psi_{j}\Bigg{\|}_{L^{\infty}}\\ &\leq\sum_{j\in{\mathfrak{u}}}(|y_{j}-y_{0,j}|+\rho_{j})\|\psi_{j}\|_{{\mathcal{W}}^{s-1}_{\infty}}+\|b({\boldsymbol{y}}_{0})\|_{{\mathcal{W}}^{s-1}_{\infty}}<\infty\,.\end{split}

This together with (3.47) implies a(𝒛𝔲)𝒲s1C\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\leq C. From the condition of ϰ\varkappa we infer |ϰ|τϑ/2|\varkappa|\tau\leq{\vartheta}/{2} which leads to

τ|ϰ|a(𝒛𝔲)Lρ(a(𝒛𝔲))/2.\tau|\varkappa|\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}\leq\rho(a({\boldsymbol{z}}_{\mathfrak{u}}))/2.

As a consequence we obtain

(ρ(a(𝒛𝔲))τ|ϰ|a(𝒛)L)11ρ(a(𝒛𝔲).\big{(}\rho(a({\boldsymbol{z}}_{\mathfrak{u}}))-\tau|\varkappa|\|a({\boldsymbol{z}})\|_{L^{\infty}}\big{)}^{-1}\leq\frac{1}{\rho(a({\boldsymbol{z}}_{\mathfrak{u}})}\,.

Since the function exp\exp is analytic in 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}), the assertion follows for the case ϑ>0\vartheta>0 by applying Theorem 3.29. In addition, for 𝒛𝔲=(zj)j𝒮𝔲(𝒚0,𝝆){\boldsymbol{z}}_{{\mathfrak{u}}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{S}_{{\mathfrak{u}}}({\boldsymbol{y}}_{0},{\boldsymbol{\rho}}) with (zj)j𝔲𝒮𝔲,N(𝝆)(z_{j})_{j\in{\mathfrak{u}}}\in\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}), we have

ρ(a(𝒛𝔲))a(𝒛𝔲)L1C>0,\rho(a({\boldsymbol{z}}_{\mathfrak{u}}))\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}^{-1}\geq C>0,

From this we conclude that uu is holomorphic in the cylinder 𝒮𝔲,N(𝝆)\mathcal{S}_{{\mathfrak{u}},N}({\boldsymbol{\rho}}) as a 𝒦1s(D)V{\mathcal{K}}_{1}^{s}({{D}})\cap V-valued map, by again Theorem 3.29. This completes the proof. ∎

Remark 3.31.

The value of ϑ\vartheta depends on the system (ψj)j(\psi_{j})_{j\in{\mathbb{N}}}. Assume that ψj=jα\psi_{j}=j^{-\alpha} for some α>1\alpha>1. Then for any 𝒚U{\boldsymbol{y}}\in U, 𝝆{\boldsymbol{\rho}} satisfying (3.28), and finite set 𝔲supp(𝝆){\mathfrak{u}}\subset\operatorname{supp}({\boldsymbol{\rho}}) we have

ϑ=inf𝒛𝔲𝒮𝔲(𝒚,𝝆)[exp(j(yj+iξj)jα)]exp(jyjjα)cosκ.\vartheta=\inf_{{\boldsymbol{z}}_{\mathfrak{u}}\in\mathcal{S}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}})}\frac{\Re[\exp(\sum_{j\in{\mathbb{N}}}(y_{j}+{\rm i}\xi_{j})j^{-\alpha})]}{\exp(\sum_{j\in{\mathbb{N}}}y_{j}j^{-\alpha})}\geq\cos\kappa\,.

We consider another case when there exists some ψj\psi_{j} such that ψjC>0\psi_{j}\geq C>0 in an open set Ω\Omega in D{D} and exp(yjψj)L1\|\exp(y_{j}\psi_{j})\|_{L^{\infty}}\geq 1 for all yj0y_{j}\leq 0. With 𝒚0=(,0,yj,0,){\boldsymbol{y}}_{0}=(\ldots,0,y_{j},0,...) and v0C0(Ω)v_{0}\in C_{0}^{\infty}(\Omega) we have in this case

ϑρ(exp(yjψj))0whenyj.\vartheta\leq\rho(\exp(y_{j}\psi_{j}))\to 0\quad\text{when}\quad y_{j}\to-\infty\,.

Hence, only for ϰ=0\varkappa=0 is satisfied Theorem 3.30 in this situation.

Due to this observation, for Kondrat’ev regularity we consider only the case ϰ=0\varkappa=0. In Section 7.6.1, we will present a stronger regularity result for a polygonal domain D2D\subset\mathbb{R}^{2}.

Lemma 3.32.

Let 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}}, f𝒦1s2(D)f\in{\mathcal{K}}_{-1}^{s-2}({{D}}), and assume that ψj𝒲s1(D)\psi_{j}\in{\mathcal{W}}^{s-1}_{\infty}({{D}}) for jj\in{\mathbb{N}}. Let 𝐲U{\boldsymbol{y}}\in U with b(𝐲)𝒲s1(D)b({\boldsymbol{y}})\in{\mathcal{W}}^{s-1}_{\infty}({{D}}). Assume further that there exists a non-negative sequence 𝛒𝛎=(ρ𝛎,j)j{\boldsymbol{\rho}}_{\boldsymbol{\nu}}=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}} such that supp(𝛎)supp(𝛒𝛎)\operatorname{supp}({\boldsymbol{\nu}})\subset\operatorname{supp}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}) and

|𝜶|s1jρ𝝂,j|rD|𝜶|D𝜶ψj|Lκ<π2.\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|r_{D}^{|{\boldsymbol{\alpha}}|}D^{{\boldsymbol{\alpha}}}\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2}\,. (3.55)

Then we have the estimate

𝝂u(𝒚)𝒦s1C𝝂!𝝆𝝂𝝂(exp(b(𝒚)L)2Ns+1(1+b(𝒚)𝒲s1)(s1)Ns.\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{{\mathcal{K}}^{s}_{1}}\leq C\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\big{(}\exp\big{(}\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}^{2N_{s}+1}\Big{(}1+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Big{)}^{(s-1)N_{s}}.
Proof.

Let 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} with 𝔲=supp(𝝂){\mathfrak{u}}=\operatorname{supp}({\boldsymbol{\nu}}). By our assumption, it is clear that (with 𝜶=0{\boldsymbol{\alpha}}=0 in (3.55))

jρ𝝂,j|ψj|Lκ<π2.\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2}\,.

Consequently, if we fix the variable yjy_{j} with j𝔲j\not\in{\mathfrak{u}}, the function uu of (3.17) is holomorphic on the domain 𝒮𝔲(𝝆𝝂)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}), see Theorem 3.30. Hence, applying Cauchy’s formula gives that

𝝂u(𝒚)𝒦s1𝝂!𝝆𝝂𝝂sup𝒛𝔲𝒞𝔲(𝒚,𝝆𝝂)u(𝒛𝔲)𝒦s1C𝝂!𝝆𝝂𝝂sup𝒛𝔲𝒞𝔲(𝒚,𝝆𝝂)1(ρ(a(𝒛𝔲)))Ns+1a(𝒛𝔲)𝒲s1Ns,\begin{split}\|\partial^{{\boldsymbol{\nu}}}u({\boldsymbol{y}})\|_{{\mathcal{K}}^{s}_{1}}&\leq\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\sup_{{\boldsymbol{z}}_{\mathfrak{u}}\in\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}})}\|u({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{K}}^{s}_{1}}\\ &\leq C\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\sup_{{\boldsymbol{z}}_{\mathfrak{u}}\in\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}})}\frac{1}{\big{(}\rho(a({\boldsymbol{z}}_{\mathfrak{u}}))\big{)}^{N_{s}+1}}\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}^{N_{s}},\end{split}

where C𝔲(𝒚,𝝆𝝂)C_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}}) is given as in (3.33). Notice that for 𝒛𝔲=(zj)j𝒞𝔲(𝒚,𝝆𝝂){\boldsymbol{z}}_{\mathfrak{u}}=(z_{j})_{j\in{\mathbb{N}}}\in\mathcal{C}_{\mathfrak{u}}({\boldsymbol{y}},{\boldsymbol{\rho}}_{\boldsymbol{\nu}}), we can write zj=yj+ηj+iξj𝒞𝒚,j(𝝆𝝂)z_{j}=y_{j}+\eta_{j}+{\rm i}\xi_{j}\in{\mathcal{C}}_{{\boldsymbol{y}},j}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}) with |ηj|ρ𝝂,j|\eta_{j}|\leq\rho_{{\boldsymbol{\nu}},j} and |ξj|ρ𝝂,j|\xi_{j}|\leq\rho_{{\boldsymbol{\nu}},j} for j𝔲j\in{\mathfrak{u}}. Hence, by (3.50), (3.51) and

a(𝒛𝔲)𝒲s1Ca(𝒛𝔲)L(1+b(𝒛𝔲)𝒲s1)s1=Cexp(b(𝒚)L)[1+|𝜶|s1rD|𝜶|j(yj+ηj+iξj)D𝜶ψjL]s1=Cexp(b(𝒚)L)[1+|𝜶|s12j𝔲ρ𝝂,j|rD|𝜶|D𝜶ψj|L+b(𝒚)𝒲s1]s1Cexp(b(𝒚)L)(1+2κ+b(𝒚)𝒲s1)s1,\begin{split}\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}&\leq C\|a({\boldsymbol{z}}_{\mathfrak{u}})\|_{L^{\infty}}\Big{(}1+\|b({\boldsymbol{z}}_{\mathfrak{u}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Big{)}^{s-1}\\ &=C\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\Bigg{[}1+\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}r_{D}^{|{\boldsymbol{\alpha}}|}\sum_{j\in{\mathbb{N}}}(y_{j}+\eta_{j}+{\rm i}\xi_{j})D^{{\boldsymbol{\alpha}}}\psi_{j}\Bigg{\|}_{L^{\infty}}\Bigg{]}^{s-1}\\ &=C\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\Bigg{[}1+\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}2\sum_{j\in{\mathfrak{u}}}\rho_{{\boldsymbol{\nu}},j}|r_{D}^{|{\boldsymbol{\alpha}}|}D^{{\boldsymbol{\alpha}}}\psi_{j}|\Bigg{\|}_{L^{\infty}}+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Bigg{]}^{s-1}\\ &\leq C\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\Big{(}1+2\kappa+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Big{)}^{s-1}\,,\end{split}

we obtain the desired result. ∎

3.8.2 Summability of KsϰK^{s}_{\varkappa}-norms of Wiener-Hermite PC expansion coefficients

To establish weighted 2\ell^{2}-summability and p\ell^{p}-summability of KsϰK^{s}_{\varkappa}-norms of Wiener-Hermite PC expansion coefficients we need the following assumption.

Assumption 3.33.

Let ss\in{\mathbb{N}}. All functions ψj\psi_{j} belong to 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}) and there exists a positive sequence (λj)j(\lambda_{j})_{j\in{\mathbb{N}}} such that (exp(λj2))j1()\big{(}\exp(-\lambda_{j}^{2})\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}}) and the series

jλj|rD|𝜶|D𝜶ψj|\sum_{j\in{\mathbb{N}}}\lambda_{j}\left|r_{D}^{|{\boldsymbol{\alpha}}|}D^{{\boldsymbol{\alpha}}}\psi_{j}\right|

converges in L(D)L^{\infty}({{D}}) for all 𝛂0d{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{d} with |𝛂|s1|{\boldsymbol{\alpha}}|\leq s-1.

Lemma 3.34.

Suppose that Assumption 3.33 holds. Then b(𝐲)b({\boldsymbol{y}}) belongs to 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}) γa.e.𝐲U\gamma-a.e.\ {\boldsymbol{y}}\in U. Furthermore, 𝔼(exp(kb(𝐲)𝒲s1))\mathbb{E}(\exp(k\|b({\boldsymbol{y}})\|_{{\mathcal{W}}_{\infty}^{s-1}})) is finite for all k[0,)k\in[0,\infty).

Proof.

Under Assumption 3.33, by [9, Theorem 2.2.] we infer that for 𝜶0d{\boldsymbol{\alpha}}\in{\mathbb{N}}_{0}^{d}, |𝜶|s1|{\boldsymbol{\alpha}}|\leq s-1, the sequence

(j=1NyjrD|𝜶|D𝜶ψj)N\left(\sum_{j=1}^{N}y_{j}r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}\psi_{j}\right)_{N\in{\mathbb{N}}}

converges to some ψ𝜶\psi_{\boldsymbol{\alpha}} in LL^{\infty} for γa.e.𝒚U\gamma-a.e.\ {\boldsymbol{y}}\in U and 𝔼(exp(kψ𝜶(𝒚)L))\mathbb{E}(\exp(k\|\psi_{\boldsymbol{\alpha}}({\boldsymbol{y}})\|_{L^{\infty}})) is finite for all k[0,)k\in[0,\infty). Hence, for γa.e.𝒚U\gamma-a.e.\ {\boldsymbol{y}}\in U, the sequence (j=1Nyjψj)N\big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\big{)}_{N\in{\mathbb{N}}} is a Cauchy sequence in 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}). Since 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}) is a Banach space, the statement follows. ∎

Theorem 3.35 (General case).

Let ss\in{\mathbb{N}}, s2s\geq 2 and D{D} be a bounded curvilinear polygonal domain. Let f𝒦1s2(D)f\in{\mathcal{K}}_{-1}^{s-2}({{D}}) and Assumption 3.33 hold. Assume there exists a sequence

ϱ=(ϱj)j(0,)with(ϱj1)jq(){\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}}\in(0,\infty)^{\infty}\ \ with\ (\varrho_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}})

for some 0<q<0<q<\infty. Assume furthermore that, for each 𝛎{\boldsymbol{\nu}}\in{\mathcal{F}}, there exists a sequence 𝛒𝛎:=(ρ𝛎,j)j[0,){\boldsymbol{\rho}}_{\boldsymbol{\nu}}:=(\rho_{{\boldsymbol{\nu}},j})_{j\in{\mathbb{N}}}\in[0,\infty)^{\infty} such that supp(𝛎)supp(𝛒𝛎)\operatorname{supp}({\boldsymbol{\nu}})\subset\operatorname{supp}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}),

sup𝝂|𝜶|s1jρ𝝂,j|rD|𝜶|D𝜶ψj|Lκ<π2,and𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂<\sup_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\sum_{|{\boldsymbol{\alpha}}|\leq s-1}\Bigg{\|}\sum_{j\in{\mathbb{N}}}\rho_{{\boldsymbol{\nu}},j}|r_{D}^{|{\boldsymbol{\alpha}}|}D^{{\boldsymbol{\alpha}}}\psi_{j}|\Bigg{\|}_{L^{\infty}}\leq\kappa<\frac{\pi}{2},\quad\text{and}\quad\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty\,

with rr\in{\mathbb{N}}, r>2/qr>2/q.

Then

𝝂β𝝂(r,ϱ)u𝝂𝒦s12<with(β𝝂(r,ϱ)1/2)𝝂q(),\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{{\mathcal{K}}^{s}_{1}}^{2}<\infty\ \ \ with\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}}),

where β𝛎(r,ϱ)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) is given in (3.36). Furthermore,

(u𝝂𝒦s1)𝝂p()with1p=1q+12.(\|u_{\boldsymbol{\nu}}\|_{{\mathcal{K}}^{s}_{1}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}})\ \ \ with\ \ \ \frac{1}{p}=\frac{1}{q}+\frac{1}{2}.
Proof.

For each 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} with 𝔲=supp(𝝂){\mathfrak{u}}=\operatorname{supp}({\boldsymbol{\nu}}) and 𝒚U{\boldsymbol{y}}\in U such that b(𝒚)𝒲s1(D)b({\boldsymbol{y}})\in{\mathcal{W}}^{s-1}_{\infty}({{D}}), Assumption 3.33 implies that the solution uu of (3.17) is holomorphic in 𝒮𝔲(𝝆𝝂)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}_{\boldsymbol{\nu}}) as a 𝒦1s(D)V{\mathcal{K}}_{1}^{s}({{D}})\cap V-valued map, see Theorem 3.30.

We obtain from Lemmata 3.32 and 3.34

U𝝂u(𝒚)𝒦s12dγ(𝒚)C𝝂!𝝆𝝂2𝝂U(exp(b(𝒚)L)4Ns+2(1+b(𝒚)𝒲s1)2(s1)Nsdγ(𝒚)C𝝂!𝝆𝝂2𝝂<.\begin{split}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{{\mathcal{K}}^{s}_{1}}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})&\leq C\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}\int_{U}\big{(}\exp\big{(}\|b({\boldsymbol{y}})\|_{L^{\infty}}\big{)}^{4N_{s}+2}\Big{(}1+\|b({\boldsymbol{y}})\|_{{\mathcal{W}}^{s-1}_{\infty}}\Big{)}^{2(s-1)N_{s}}\,\mathrm{d}\gamma({\boldsymbol{y}})\\ &\leq C\frac{{\boldsymbol{\nu}}!}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty\,.\end{split}

This leads to

𝝂β𝝂(r,ϱ)u𝝂𝒦s12=𝝂rϱ2𝝂𝝂!U𝝂u(𝒚)𝒦s12dγ(𝒚)C𝝂r𝝂!ϱ2𝝂𝝆𝝂2𝝂<.\begin{split}\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{{\mathcal{K}}^{s}_{1}}^{2}&=\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{U}\|\partial^{\boldsymbol{\nu}}u({\boldsymbol{y}})\|_{{\mathcal{K}}^{s}_{1}}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})\leq C\sum_{\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r}\frac{{\boldsymbol{\nu}}!{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\rho}}_{\boldsymbol{\nu}}^{2{\boldsymbol{\nu}}}}<\infty\,.\end{split}

The rest of the proof follows similarly to the proof of Theorem 3.13. ∎

Similarly to Corollaries 3.14 and 3.15 from Theorem 3.35 we obtain

Corollary 3.36 (The case of global supports).

Let ss\in{\mathbb{N}}, s2s\geq 2 and D{D} be a bounded curvilinear, polygonal domain. Assume that for all jj\in{\mathbb{N}} holds ψj𝒲s1(D)\psi_{j}\in{\mathcal{W}}^{s-1}_{\infty}({{D}}), and that f𝒦1s2(D)f\in{\mathcal{K}}_{-1}^{s-2}({{D}}). Assume further that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψj𝒲s1)j1()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{{\mathcal{W}}_{\infty}^{s-1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{1}({\mathbb{N}})\ \ and\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<0<q<\infty. Then we have (u𝛎𝒦s1)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{{\mathcal{K}}^{s}_{1}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

Corollary 3.37 (The case of disjoint supports).

Let ss\in{\mathbb{N}}, s2s\geq 2 and Dd{D}\subset{\mathbb{R}}^{d} with d2d\geq 2 be a bounded curvilinear polygonal domain. Assume that all the functions ψj\psi_{j} belong to 𝒲s1(D){\mathcal{W}}^{s-1}_{\infty}({{D}}) and have disjoint supports. Assume further that f𝒦1s2(D)f\in{\mathcal{K}}_{-1}^{s-2}({{D}}) and that there exists a sequence of positive numbers 𝛌=(λj)j{\boldsymbol{\lambda}}=(\lambda_{j})_{j\in{\mathbb{N}}} such that

(λjψj𝒲s1)j2()and(λj1)jq(),\big{(}\lambda_{j}\|\psi_{j}\|_{{\mathcal{W}}_{\infty}^{s-1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{2}({\mathbb{N}})\ \ and\ \ (\lambda_{j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q}({\mathbb{N}}),

for some 0<q<0<q<\infty. Then (u𝛎𝒦s1)𝛎p()(\|u_{\boldsymbol{\nu}}\|_{{\mathcal{K}}^{s}_{1}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p}({\mathcal{F}}) with 1p=1q+12\frac{1}{p}=\frac{1}{q}+\frac{1}{2}.

3.9 Bibliographical remarks

In this section, we briefly recall some known related results in previous works on p\ell^{p}-summability and on weighted 2\ell^{2}-summability of the generalized PC expansion coefficients of solutions to parametric divergence-form elliptic PDEs (3.17), as well as some applications to best nn-term approximation.

A basic role in the approximation and numerical integration for parametric divergence-form elliptic PDEs (3.17) are generalized PC expansions for the dependence on the parametric variables. In [32, 37, 38, 39], based on the conditions (ψjW1)jp()\big{(}\|\psi_{j}\|_{W_{\infty}^{1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{p}({\mathbb{N}}) for some 0<p<10<p<1 on the affine expansion

a(𝒚)=a¯+j=1yjψj,𝒚[0,1],a({\boldsymbol{y}})=\bar{a}+\sum_{j=1}^{\infty}y_{j}\psi_{j},\ \ \ {\boldsymbol{y}}\in[0,1]^{\infty}, (3.56)

the authors have proven p\ell^{p}-summability of the coefficients in a Taylor or Legendre PC expansion and hence proposed adaptive best nn-term rate optimal approximation methods of Galerkin and collocation type by choosing a set of nn largest estimated terms in these expansions. To derive a fully discrete approximation, the best nn-term approximants are then discretized by finite element methods. Some results on convergence rates of Galerkin approximation were proven in [72] for the log-Gaussian expansion (3.18), based on the summability (jψjW1)jp()\big{(}j\|\psi_{j}\|_{W_{\infty}^{1}}\big{)}_{j\in{\mathbb{N}}}\in\ell^{p}({\mathbb{N}}) for some 0<p<10<p<1. However, in these papers possible local support properties of the component functions ψj\psi_{j} were not taken into account.

A different approach to studying summability that takes into account the support properties has been recently proposed in [10] for the affine-parametric case, in [9] for the log-exponential, parametric case, and in [8] for extension of both cases to higher-order Sobolev norms of the corresponding generalized PC expansion coefficients. This approach leads to significant improvements on the results on p\ell^{p}-summability and therefore, on best nn-term semi-discrete and fully discrete approximations when the functions ψj\psi_{j} have limited overlap, such as splines, finite elements or compactly supported wavelet bases. These approximation results provide a benchmark for convergence rates.

We present some results from [9] and [8] on p\ell^{p}-summability and weighted 2\ell^{2}-summability of the Wiener-Hermite PC expansion coefficients of the solution to the parametric divergence-form elliptic PDEs (3.17)–(3.18) which were proven by real-variable bootstrapping arguments.

For convenience, we use the conventions:

W1:=V,W2:=W,H1(D):=V,H0(D):=L2(D),W0,(D):=L(D),W^{1}:=V,\ \ W^{2}:=W,\ \ H^{-1}({{D}}):=V^{\prime},\ \ H^{0}({{D}}):=L^{2}({{D}}),\ \ W^{0,\infty}({{D}}):=L^{\infty}({{D}}),

where we recall W:={vV:ΔvL2(D)},W:=\{v\in V\;:\;\Delta v\in L^{2}({{D}})\}\;, is the space equipped with the norm vW:=ΔvL2.\|v\|_{W}:=\|\Delta v\|_{L^{2}}. The following theorem and lemma were proven in [9] for i=1i=1 and in [8] for i=2i=2.

Theorem 3.38.

Let i=1,2i=1,2. Assume that the right side ff in (3.17) belongs to Hi2(D)H^{i-2}({{D}}), that the domain D{{D}} has Ci1,1C^{i-1,1} smoothness, that all functions ψj\psi_{j} belong to Wi1,(D)W^{i-1,\infty}({{D}}). Assume that there exist a number 0<qi<0<q_{i}<\infty and a sequence ϱi=(ϱi;j)j{\boldsymbol{\varrho}}_{i}=(\varrho_{i;j})_{j\in{\mathbb{N}}} of positive numbers such that (ϱi;j1)jqi()(\varrho_{i;j}^{-1})_{j\in{\mathbb{N}}}\in\ell^{q_{i}}({\mathbb{N}}) and

sup|α|i1jϱi;j|Dαψj|L<.\sup_{|\alpha|\leq i-1}\left\|\sum_{j\in{\mathbb{N}}}\varrho_{i;j}|D^{\alpha}\psi_{j}|\right\|_{L^{\infty}}<\infty. (3.57)

Then we have that for any rr\in{\mathbb{N}},

𝝂(σi;𝝂u𝝂Wi)2<and(σi;𝝂1)𝝂qi(),\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}(\sigma_{i;{\boldsymbol{\nu}}}\|u_{\boldsymbol{\nu}}\|_{W^{i}})^{2}<\infty\quad\text{and}\quad(\sigma_{i;{\boldsymbol{\nu}}}^{-1})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{i}}(\mathcal{F}), (3.58)

where

σi;𝝂2:=𝝂r(𝝂𝝂)ϱi2𝝂.\sigma_{i;{\boldsymbol{\nu}}}^{2}:=\sum_{\|{\boldsymbol{\nu}}^{\prime}\|_{\ell^{\infty}}\leq r}{{\boldsymbol{\nu}}\choose{\boldsymbol{\nu}}^{\prime}}{\boldsymbol{\varrho}}_{i}^{2{\boldsymbol{\nu}}^{\prime}}. (3.59)

Furthermore,

(u𝝂Wi)𝝂2qi/(2+qi)().(\|u_{\boldsymbol{\nu}}\|_{W^{i}})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{2q_{i}/(2+q_{i})}({\mathcal{F}}).

Notice that the assumption (3.57) which give the weighted 2\ell^{2}-summability (3.58), already reflects the support properties of the component functions ψj\psi_{j}.

For τ,λ0\tau,\lambda\geq 0, we define the family

p𝝂(τ,λ):=j(1+λνj)τ,𝝂,p_{\boldsymbol{\nu}}(\tau,\lambda):=\prod_{j\in{\mathbb{N}}}(1+\lambda\nu_{j})^{\tau},\quad{\boldsymbol{\nu}}\in{\mathcal{F}}, (3.60)

with the abbreviation p𝝂(τ):=p𝝂(τ,1)p_{\boldsymbol{\nu}}(\tau):=p_{\boldsymbol{\nu}}(\tau,1).

We make use of the following notation

1:=,2:={𝝂:νj1,j}.\mathcal{F}_{1}:=\mathcal{F},\quad\mathcal{F}_{2}:=\{{\boldsymbol{\nu}}\in\mathcal{F}:\nu_{j}\not=1,\ j\in{\mathbb{N}}\}. (3.61)
Lemma 3.39.

Let 0<q<0<q<\infty, s=1,2s=1,2 and τ,λ0\tau,\lambda\geq 0. Let 𝛒=(ρj)j{\boldsymbol{\rho}}=(\rho_{j})_{j\in{\mathbb{N}}} be a sequence of positive numbers such (ρj1)j(\rho_{j}^{-1})_{j\in{\mathbb{N}}} belongs to q()\ell^{q}({\mathbb{N}}). For rr\in{\mathbb{N}}, define the family (σ𝛎)𝛎(\sigma_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} by

σ𝝂2:=𝝂r(𝝂𝝂)𝝆2𝝂.\sigma_{\boldsymbol{\nu}}^{2}:=\sum_{\|{\boldsymbol{\nu}}^{\prime}\|_{\ell^{\infty}}\leq r}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\nu}}^{\prime}}{\boldsymbol{\rho}}^{2{\boldsymbol{\nu}}^{\prime}}.

Then for any r>2s(τ+1)qr>\frac{2s(\tau+1)}{q}, we have

𝝂sp𝝂(τ,λ)σ𝝂q/s<.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{s}}p_{\boldsymbol{\nu}}(\tau,\lambda)\sigma_{\boldsymbol{\nu}}^{-q/s}<\infty.

This lemma has been proven in [43, Lemma 5.3]. Observe that for s=1s=1 and τ=0\tau=0, an equivalent formulation of Lemma 3.39 is Lemma 3.11.

Theorem 3.38 and Lemma 3.39 directly imply the following corollary.

Corollary 3.40.

Under the assumptions of Theorem 3.38, let s=1,2s=1,2 and τ,λ0\tau,\lambda\geq 0. Then we have that for any r>2s(τ+1)qr>\frac{2s(\tau+1)}{q},

𝝂s(σi;𝝂u𝝂Wi)2<and(p𝝂(τ,λ)σi;𝝂1)𝝂sqi/s(s).\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{s}}(\sigma_{i;{\boldsymbol{\nu}}}\|u_{\boldsymbol{\nu}}\|_{W^{i}})^{2}<\infty\quad\text{and}\quad(p_{\boldsymbol{\nu}}(\tau,\lambda)\sigma_{i;{\boldsymbol{\nu}}}^{-1})_{{\boldsymbol{\nu}}\in\mathcal{F}_{s}}\in\ell^{q_{i}/s}(\mathcal{F}_{s}). (3.62)

As commented in Section 3.6.2, in the case of disjoint or finitely overlapping supports the results on sparsity of Theorem 3.38 and Corollary 3.40 are stronger than those in Sections 3.6.2 and 3.7.2. They play a basic role in best nn-term approximation [9, 8] and linear approximation and quadrature [43] (see also [45]) of the solution to the parametric divergence-form elliptic PDEs (3.17)–(3.18).

4 Sparsity for holomorphic functions

In Section 3 we introduced a concept of holomorphic extensions of countably-parametric families {u(𝒚):𝒚U}V\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset V in the separable Hilbert space VV with respect to the parameter 𝒚{\boldsymbol{y}} into the Cartesian product 𝒮𝔲(𝝆)\mathcal{S}_{\mathfrak{u}}({\boldsymbol{\rho}}) of strips in the complex domain (cp. (3.27)). We now introduce a refinement which is required for the ensuing results on rates of numerical approximation and integration of such families, based on sparsity (weighted 2\ell^{2}-summability) and of Wiener-Hermite PC expansions of {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}: quantified parametric holomorphy of (complex extensions of) the parametric families {u(𝒚):𝒚U}X\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset X for a separable Hilbert space X. Section 4.1 presents a definition of quantified holomorphy of families {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} and discusses the sparsity of the Wiener-Hermite PC expansion coefficients of these families. In Section 4.2, we present the notion (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy of composite functions. In Section 4.3, we analyze some examples of holomorphic functions which are solutions to certain PDEs.

There are two basic steps in the approximations which we consider:
(i) We truncate the countably-parametric family {u(𝒚):𝒚U}X\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset X to a finite number NN\in{\mathbb{N}} of parameters. This step, which is sometimes also referred to as “dimension-truncation”, of course implicitly depends on the enumeration of the coordinates yj𝒚y_{j}\in{\boldsymbol{y}}. We assume throughout that this numbering is fixed by the indexing of the Parseval frame in Theorem 2.21 which frame is used as affine representation system to parametrize the uncertain input a=exp(b)a=\exp(b) of the PDE of interest. We emphasize that the finite dimension NN\in{\mathbb{N}} of the truncated parametric Wiener-Hermite PC expansion is a discretization parameter, and we will be interested in quantitative bounds on the error incurred by restricting {u(𝒚):𝒚U}X\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset X to Wiener-Hermite PC expansions of the first NN active variables only. We denote these restrictions by {uN(𝒚):𝒚U}\{u_{N}({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}.
(ii) The coefficients u𝝂Xu_{\boldsymbol{\nu}}\in X of the resulting, finite-parametric Wiener-Hermite PC expansion, can not be computed exactly, but must be numerically approximated. As is done in stochastic collocation and stochastic Galerkin algorithms, we seek numerical approximations of u𝝂u_{\boldsymbol{\nu}} in suitable, finite-dimensional subspaces XlXX_{l}\subset X. Assuming the collection (Xl)lX(X_{l})_{l\in{\mathbb{N}}}\subset X to be dense in XX, any prescribed tolerance ε>0\varepsilon>0 of approximation of uN(𝒚)u_{N}({\boldsymbol{y}}) in L2(U,X;γ)L^{2}(U,X;\gamma) can be met. For notational convenience, we also set X0={0}X_{0}=\{0\}.
In computational practice, however, given a target accuracy ε(0,1]\varepsilon\in(0,1], one searches an allocation of l:×(0,1]:(𝝂,ε)l(𝝂,ε)l:{\mathcal{F}}\times(0,1]\to{\mathbb{N}}:({\boldsymbol{\nu}},\varepsilon)\mapsto l({\boldsymbol{\nu}},\varepsilon) of discretization levels along the “active” Wiener-Hermite PC expansion coefficients which ensures that the prescribed tolerance ε(0,1]\varepsilon\in(0,1] is met with possibly minimal “computational budget”. We propose and analyze the a-priori construction of an allocation ll which ensures convergence rates of the corresponding collocation approximations which are independent of NN (i.e. they are free from the “curse of dimensionality”). These approximations are based on “stochastic collocation”, i.e. on sampling the parametric family {u(𝒚):𝒚U}V\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset V in a collection of deterministic Gaussian coordinates in UU. We prove, subsequently, dimension-independent convergence rates of the sparse collocation w.r.t. 𝒚U{\boldsymbol{y}}\in U and w.r.t. the subspaces XlXX_{l}\subset X realize convergence rates which are free from the curse of dimensionality. These rates depend only on the summability (resp. sparsity) of the coefficients of the norm of the Wiener-Hermite PC expansion of the parametric family {u(𝒚):𝒚U}\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\} with respect to 𝒚{\boldsymbol{y}}.

4.1 (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-Holomorphy and sparsity

We introduce the concept of “(𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions”, which constitutes a subset of L2(U,X;γ)L^{2}(U,X;\gamma). As such these functions are typically not pointwise well defined for each 𝒚U{\boldsymbol{y}}\in U. In order to still define a suitable form of pointwise function evaluations to be used for numerical algorithms such as sampling at 𝒚U{\boldsymbol{y}}\in U for “stochastic collocation” or for quadrature in “stochastic Galerkin” algorithms, we define them as L2(U,X;γ)L^{2}(U,X;\gamma) limits of certain smooth (pointwise defined) functions, cp. Remark. 4.4 and Example 6.9 ahead.

For NN\in\mathbb{N} and ϱ=(ϱj)j=1N(0,)N{\boldsymbol{\varrho}}=(\varrho_{j})_{j=1}^{N}\in(0,\infty)^{N} set (cp. (3.27))

𝒮(ϱ):={𝒛N:|𝔪zj|<ϱjj}and(ϱ):={𝒛N:|zj|<ϱjj}.{\mathcal{S}}({\boldsymbol{\varrho}}):=\{{\boldsymbol{z}}\in\mathbb{C}^{N}\,:\,|\mathfrak{Im}z_{j}|<\varrho_{j}\leavevmode\nobreak\ \forall j\}\qquad\text{and}\qquad{\mathcal{B}}({\boldsymbol{\varrho}}):=\{{\boldsymbol{z}}\in\mathbb{C}^{N}\,:\,|z_{j}|<\varrho_{j}\leavevmode\nobreak\ \forall j\}. (4.1)
Definition 4.1 ((𝒃,ξ,δ,X{\boldsymbol{b}},\xi,\delta,X)-Holomorphy).

Let XX be a complex, separable Hilbert space, 𝐛=(bj)j(0,){\boldsymbol{b}}=(b_{j})_{j\in\mathbb{N}}\in(0,\infty)^{\infty} and ξ>0\xi>0, δ>0\delta>0.

For NN\in\mathbb{N}, ϱ(0,)N{\boldsymbol{\varrho}}\in(0,\infty)^{N} is called (𝒃,ξ)({\boldsymbol{b}},\xi)-admissible if

j=1Nbjϱjξ.\sum_{j=1}^{N}b_{j}\varrho_{j}\leq\xi\,. (4.2)

A function uL2(U,X;γ)u\in L^{2}(U,X;\gamma) is called (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic if

  1. (i)

    for every NN\in\mathbb{N} there exists uN:NXu_{N}:\mathbb{R}^{N}\to X, which, for every (𝒃,ξ)({\boldsymbol{b}},\xi)-admissible ϱ(0,)N{\boldsymbol{\varrho}}\in(0,\infty)^{N}, admits a holomorphic extension (denoted again by uNu_{N}) from 𝒮(ϱ)X{\mathcal{S}}({\boldsymbol{\varrho}})\to X; furthermore, for all N<MN<M

    uN(y1,,yN)=uM(y1,,yN,0,,0)(yj)j=1NN,u_{N}(y_{1},\dots,y_{N})=u_{M}(y_{1},\dots,y_{N},0,\dots,0)\qquad\forall(y_{j})_{j=1}^{N}\in\mathbb{R}^{N}, (4.3)
  2. (ii)

    for every NN\in\mathbb{N} there exists φN:N+\varphi_{N}:\mathbb{R}^{N}\to\mathbb{R}_{+} such that φNL2(N;γN)δ\|\varphi_{N}\|_{L^{2}(\mathbb{R}^{N};\gamma_{N})}\leq\delta and

    supϱ(0,)Nis (𝒃,ξ)-adm.sup𝒛(ϱ)uN(𝒚+𝒛)XφN(𝒚)𝒚N,\sup_{\begin{subarray}{c}{\boldsymbol{\varrho}}\in(0,\infty)^{N}\\ \text{is $({\boldsymbol{b}},\xi)$-adm.}\end{subarray}}\leavevmode\nobreak\ \sup_{{\boldsymbol{z}}\in{\mathcal{B}}({\boldsymbol{\varrho}})}\|u_{N}({\boldsymbol{y}}+{\boldsymbol{z}})\|_{X}\leq\varphi_{N}({\boldsymbol{y}})\qquad\forall{\boldsymbol{y}}\in\mathbb{R}^{N},
  3. (iii)

    with u~N:UX\tilde{u}_{N}:U\to X defined by u~N(𝒚):=uN(y1,,yN)\tilde{u}_{N}({\boldsymbol{y}}):=u_{N}(y_{1},\dots,y_{N}) for 𝒚U{\boldsymbol{y}}\in U it holds

    limNuu~NL2(U,X;γ)=0.\lim_{N\to\infty}\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}=0.

We interpret the definition of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy in the following remarks.

Remark 4.2.

While the numerical value of ξ>0\xi>0 in Definition 4.1 of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy is of minor importance in the definition, the sequence 𝒃{\boldsymbol{b}} and the constant δ\delta will crucially influence the magnitude of our upper bounds of the Wiener-Hermite PC expansion coefficients: The stronger the decay of 𝒃{\boldsymbol{b}}, the larger we can choose the elements of the sequence ϱ{\boldsymbol{\varrho}}, so that ϱ{\boldsymbol{\varrho}} satisfies (4.2). Hence stronger decay of 𝒃{\boldsymbol{b}} indicates larger domains of holomorphic extension. The constant δ\delta is an upper bound of these extensions in the sense of item (ii). Importantly, the decay of 𝒃{\boldsymbol{b}} will determine the sparsity of the Wiener-Hermite PC expansion coefficients, while decreasing δ\delta by a factor will roughly speaking translate to a decrease of all coefficients by the same factor.

Remark 4.3.

Since uNL2(N,X;γN)u_{N}\in L^{2}(\mathbb{R}^{N},X;\gamma_{N}), the function u~N\tilde{u}_{N} in item (iii) belongs to L2(U,X;γ)L^{2}(U,X;\gamma) by Fubini’s theorem.

Remark 4.4.

[Evaluation of countably-parametric functions] In the following sections, for arbitrary NN\in\mathbb{N} and (yj)j=1NN(y_{j})_{j=1}^{N}\in\mathbb{R}^{N} we will write

u(y1,,yN,0,0,):=uN(y1,,yN).u(y_{1},\dots,y_{N},0,0,\dots):=u_{N}(y_{1},\dots,y_{N}). (4.4)

This is well-defined due to (4.3). Note however that (4.4) should be considered as an abuse of notation, since pointwise evaluations of functions uL2(U,X;γ)u\in L^{2}(U,X;\gamma) are in general not well-defined.

Remark 4.5.

The assumption of XX being separable is not necessary in Definition 4.1: Every function uN:NXu_{N}:\mathbb{R}^{N}\to X as in Definition 4.1 is continuous since it allows a holomorphic extension. Hence,

AN,n:={uN((yj)j=1N):yj[n,n]j}XA_{N,n}:=\{u_{N}((y_{j})_{j=1}^{N})\,:\,y_{j}\in[-n,n]\leavevmode\nobreak\ \forall j\}\subseteq X

is compact and thus there is a countable set XN,nXX_{N,n}\subseteq X which is dense in AN,nA_{N,n} for every NN, nn\in\mathbb{N}. Then nAN,n\bigcup_{n\in\mathbb{N}}A_{N,n} is contained in the (separable) closed span X~\tilde{X} of

N,nXN,nX.\bigcup_{N,n\in\mathbb{N}}X_{N,n}\subseteq X.

Since u~NL2(U,X~;γ)\tilde{u}_{N}\in L^{2}(U,\tilde{X};\gamma) for every NN\in\mathbb{N} we also have

u=limNuNL2(U,X~;γ).u=\lim_{N\to\infty}u_{N}\in L^{2}(U,\tilde{X};\gamma).

Hence, uu is separably valued.

Lemma 4.6.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic, let NN\in\mathbb{N} and 0<κ<ξ<0<\kappa<\xi<\infty. Let uNu_{N}, φN\varphi_{N} be as in Definition 4.1. Then with 𝐛N=(bj)j=1N{\boldsymbol{b}}_{N}=(b_{j})_{j=1}^{N} it holds for every 𝛎0N{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}

𝝂uN(𝒚)X𝝂!|𝝂||𝝂|𝒃N𝝂κ|𝝂|𝝂𝝂φN(𝒚)𝒚N.\|\partial^{{\boldsymbol{\nu}}}u_{N}({\boldsymbol{y}})\|_{X}\leq\frac{{\boldsymbol{\nu}}!|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}{\boldsymbol{b}}_{N}^{\boldsymbol{\nu}}}{\kappa^{|{\boldsymbol{\nu}}|}{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\varphi_{N}({\boldsymbol{y}})\qquad\forall{\boldsymbol{y}}\in\mathbb{R}^{N}.
Proof.

For 𝝂0N{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N} fixed we choose ϱ=(ϱj)j=1N{\boldsymbol{\varrho}}=(\varrho_{j})_{j=1}^{N} with ϱj=κνj|𝝂|bj\varrho_{j}=\kappa\frac{\nu_{j}}{|{\boldsymbol{\nu}}|b_{j}} for jsupp(𝝂)j\in\operatorname{supp}({\boldsymbol{\nu}}) and ϱj=ξκNbj\varrho_{j}=\frac{\xi-\kappa}{Nb_{j}} for jsupp(𝝂)j\not\in\operatorname{supp}({\boldsymbol{\nu}}). Then

j=1Nϱjbj=κjsupp(𝝂)νj|𝝂|+jsupp(𝝂)ξκNξ.\sum_{j=1}^{N}\varrho_{j}b_{j}=\kappa\sum_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\frac{\nu_{j}}{|{\boldsymbol{\nu}}|}+\sum_{j\not\in\operatorname{supp}({\boldsymbol{\nu}})}\frac{\xi-\kappa}{N}\leq\xi.

Hence ϱ{\boldsymbol{\varrho}} is (𝒃,ξ)({\boldsymbol{b}},\xi)-admissible, i.e. there exists a holomorphic extension uN:𝒮(ϱ)Xu_{N}:{\mathcal{S}}({\boldsymbol{\varrho}})\to X as in Definition 4.1 (i)-(ii). Applying Cauchy’s integral formula as in the proof of Lemma 3.9 we obtain the desired estimate. ∎

Let us recall the following. Let again XX be a separable Hilbert space and uL2(U,X;γ)u\in L^{2}(U,X;\gamma). Then

L2(U,X;γ)=L2(U;γ)XL^{2}(U,X;\gamma)=L^{2}(U;\gamma)\otimes X

with Hilbertian tensor product, and uu can be represented in a Wiener-Hermite PC expansion

u=𝝂u𝝂H𝝂,u=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}u_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}}, (4.5)

where

u𝝂=Uu(𝒚)H𝝂(𝒚)dγ(𝒚)u_{\boldsymbol{\nu}}=\int_{U}u({\boldsymbol{y}})H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})

are the Wiener-Hermite PC expansion coefficients. Also, there holds the Parseval-type identity

uL2(U,X;γ)2=𝝂u𝝂X2.\|u\|_{L^{2}(U,X;\gamma)}^{2}=\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\|u_{\boldsymbol{\nu}}\|_{X}^{2}\,.

When uu is (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic, then we have for the functions uN:NXu_{N}:\mathbb{R}^{N}\to X in Definition 4.1

uN=𝝂0NuN,𝝂H𝝂,u_{N}=\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}u_{N,{\boldsymbol{\nu}}}H_{\boldsymbol{\nu}},

where

uN,𝝂=NuN(𝒚)H𝝂(𝒚)dγN(𝒚).u_{N,{\boldsymbol{\nu}}}=\int_{\mathbb{R}^{N}}u_{N}({\boldsymbol{y}})H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma_{N}({\boldsymbol{y}}).

In an analogous manner to (3.36), for rr\in{\mathbb{N}} and a finite sequence of nonnegative numbers ϱN=(ϱj)j=1N{\boldsymbol{\varrho}}_{N}=(\varrho_{j})_{j=1}^{N}, we define

β𝝂(r,ϱN):=𝝂0N:𝝂r(𝝂𝝂)ϱ2𝝂=j=1N(=0r(νj)ϱj2),𝝂0N.\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N}):=\sum_{{\boldsymbol{\nu}}^{\prime}\in\mathbb{N}_{0}^{N}:\ \|{\boldsymbol{\nu}}^{\prime}\|_{\ell^{\infty}}\leq r}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\nu}}^{\prime}}{\boldsymbol{\varrho}}^{2{\boldsymbol{\nu}}^{\prime}}=\prod_{j=1}^{N}\Bigg{(}\sum_{\ell=0}^{r}\binom{\nu_{j}}{\ell}\varrho_{j}^{2\ell}\Bigg{)},\ \ \ {\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}. (4.6)
Lemma 4.7.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic, let NN\in\mathbb{N} and let ϱN=(ϱj)j=1N[0,)N{\boldsymbol{\varrho}}_{N}=(\varrho_{j})_{j=1}^{N}\in[0,\infty)^{N}.

Then, for any fixed rr\in\mathbb{N}, there holds the identity

𝝂0Nβ𝝂(r,ϱN)uN,𝝂X2={𝝂0N:𝝂r}ϱN2𝝂𝝂!N𝝂uN(𝒚)X2dγN(𝒚).\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{2}=\sum_{\{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}\,:\,\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r\}}\frac{{\boldsymbol{\varrho}}_{N}^{2{\boldsymbol{\nu}}}}{{\boldsymbol{\nu}}!}\int_{\mathbb{R}^{N}}\|\partial^{\boldsymbol{\nu}}u_{N}({\boldsymbol{y}})\|_{X}^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}}). (4.7)
Proof.

From Lemma 4.6, for any 𝝂0N{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}, we have with 𝒃N=(bj)j=1N{\boldsymbol{b}}_{N}=(b_{j})_{j=1}^{N}

N𝝂uN(𝒚)X2dγN(𝒚)\displaystyle\int_{\mathbb{R}^{N}}\|\partial^{\boldsymbol{\nu}}u_{N}({\boldsymbol{y}})\|_{X}^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}}) N|𝝂!|𝝂||𝝂|𝒃N𝝂κ|𝝂|𝝂𝝂φN(𝒚)|2dγN(𝒚)\displaystyle\leq\int_{\mathbb{R}^{N}}\Big{|}\frac{{\boldsymbol{\nu}}!|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}{\boldsymbol{b}}_{N}^{\boldsymbol{\nu}}}{\kappa^{|{\boldsymbol{\nu}}|}{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\varphi_{N}({\boldsymbol{y}})\Big{|}^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}})
=(𝝂!|𝝂||𝝂|𝒃N𝝂κ|𝝂|𝝂𝝂)2N|φN(𝒚)|2dγN(𝒚)<\displaystyle=\Big{(}\frac{{\boldsymbol{\nu}}!|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}{\boldsymbol{b}}_{N}^{\boldsymbol{\nu}}}{\kappa^{|{\boldsymbol{\nu}}|}{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\Big{)}^{2}\int_{\mathbb{R}^{N}}\big{|}\varphi_{N}({\boldsymbol{y}})\big{|}^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}})<\infty (4.8)

by our assumption. This condition allows us to integrate by parts as in the proof of [9, Theorem 3.3]. Following the argument there we obtain (4.7). ∎

Theorem 4.8.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) and some p(0,1)p\in(0,1). Let rr\in\mathbb{N}.

Then, with

ϱj:=bjp1ξ4r!𝒃p,j,\varrho_{j}:=b_{j}^{p-1}\frac{\xi}{4\sqrt{r!}\|{\boldsymbol{b}}\|_{\ell^{p}}},\ \ j\in\mathbb{N}, (4.9)

and ϱN=(ϱj)j=1N{\boldsymbol{\varrho}}_{N}=(\varrho_{j})_{j=1}^{N} it holds for all NN\in\mathbb{N},

𝝂0Nβ𝝂(r,ϱN)uN,𝝂X2δ2C(𝒃)<withβ𝝂(r,ϱN)1/2p/(1p)(0N)C(𝒃,ξ)<\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\leq\delta^{2}C({\boldsymbol{b}})<\infty\ \ \ with\ \ \ \|\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})^{-1/2}\|_{\ell^{p/(1-p)}(\mathbb{N}_{0}^{N})}\leq C^{\prime}({\boldsymbol{b}},\xi)<\infty (4.10)

for some constants C(𝐛)C({\boldsymbol{b}}) and C(𝐛,ξ)C^{\prime}({\boldsymbol{b}},\xi) depending on 𝐛{\boldsymbol{b}} and ξ\xi, but independent of δ\delta and NN\in\mathbb{N}.

Furthermore, for every NN\in\mathbb{N} and every q>0q>0 there holds

(uN,𝝂X)𝝂0Nq(0N).(\|u_{N,{\boldsymbol{\nu}}}\|_{X})_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\in\ell^{q}(\mathbb{N}_{0}^{N}).

If q2p2pq\geq\frac{2p}{2-p} then there exists a constant C>0C>0 such that for all NN\in\mathbb{N} holds

(uN,𝝂X)𝝂q(0N)C<.\big{\|}(\|u_{N,{\boldsymbol{\nu}}}\|_{X})_{{\boldsymbol{\nu}}}\big{\|}_{\ell^{q}(\mathbb{N}_{0}^{N})}\leq C<\infty\;.
Proof.

We have

jϱjbj=ξ4r!𝒃pjbjp<,\sum_{j\in\mathbb{N}}\varrho_{j}b_{j}=\frac{\xi}{4\sqrt{r!}\|{\boldsymbol{b}}\|_{\ell^{p}}}\sum_{j\in\mathbb{N}}b_{j}^{p}<\infty,

and (ϱj1)jp/(1p)()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{p/(1-p)}(\mathbb{N}). Set κ:=ξ/2(0,ξ)\kappa:=\xi/2\in(0,\xi). Inserting (4.1) into (4.7) we obtain with ϱN=(ϱj)j=1N{\boldsymbol{\varrho}}_{N}=(\varrho_{j})_{j=1}^{N}

𝝂0Nβ𝝂(r,ϱN)uN,𝝂X2\displaystyle\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{2} δ2{𝝂0N:𝝂r}((𝝂!)1/2|𝝂||𝝂|ϱN𝝂𝒃N𝝂κ|𝝂|𝝂𝝂)2\displaystyle\leq\delta^{2}\sum_{\{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}\,:\,\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r\}}\left(\frac{({\boldsymbol{\nu}}!)^{1/2}|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}{\boldsymbol{\varrho}}_{N}^{{\boldsymbol{\nu}}}{\boldsymbol{b}}_{N}^{{\boldsymbol{\nu}}}}{\kappa^{|{\boldsymbol{\nu}}|}{\boldsymbol{\nu}}^{{\boldsymbol{\nu}}}}\right)^{2}
δ2{𝝂0N:𝝂r}(|𝝂||𝝂|j=1N(bjp2𝒃p)νj𝝂𝝂)2,\displaystyle\leq\delta^{2}\sum_{\{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}\,:\,\|{\boldsymbol{\nu}}\|_{\ell^{\infty}}\leq r\}}\left(\frac{|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}\prod_{j=1}^{N}\Big{(}\frac{b_{j}^{p}}{2\|{\boldsymbol{b}}\|_{\ell^{p}}}\Big{)}^{\nu_{j}}}{{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\right)^{2},

where we used (ϱjbj)2=bj2pκ/(2(r!))(\varrho_{j}b_{j})^{2}=b_{j}^{2p}\kappa/(2(r!)) and the bound

NφN(𝒚)2dγN(𝒚)δ2\int_{\mathbb{R}^{N}}\varphi_{N}({\boldsymbol{y}})^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}})\leq\delta^{2}

from Definition 4.1 (ii). With b~j:=bjp/(2𝒃p)\tilde{b}_{j}:=b_{j}^{p}/(2\|{\boldsymbol{b}}\|_{\ell^{p}}) the last term is bounded independent of NN by δ2C(𝒃)\delta^{2}C({\boldsymbol{b}}) with

C(𝒃):=(𝝂|𝝂||𝝂|𝝂𝝂𝒃~𝝂)1/2,C({\boldsymbol{b}}):=\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\frac{|{\boldsymbol{\nu}}|^{|{\boldsymbol{\nu}}|}}{{\boldsymbol{\nu}}^{\boldsymbol{\nu}}}\tilde{\boldsymbol{b}}^{\boldsymbol{\nu}}\right)^{1/2},

since the 1\ell^{1}-norm is an upper bound of the 2\ell^{2}-norm. As is well-known, the latter quantity is finite due to 𝒃~1<1\|\tilde{\boldsymbol{b}}\|_{\ell^{1}}<1, see, e.g., the argument in [37, Page 61].

Now introduce ϱ~N,j:=ϱj\tilde{\varrho}_{N,j}:=\varrho_{j} if jNj\leq N and ϱ~N,j:=exp(j)\tilde{\varrho}_{N,j}:=\exp(j) otherwise. For any q>0q>0 we then have (ϱ~N,j1)jq()(\tilde{\varrho}_{N,j}^{-1})_{j\in\mathbb{N}}\in\ell^{q}(\mathbb{N}) and by Lemma 3.11 this implies

(β𝝂(r,ϱ~N)1)𝝂q/2()(\beta_{\boldsymbol{\nu}}(r,\tilde{\boldsymbol{\varrho}}_{N})^{-1})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q/2}(\mathcal{F})

as long as r>2/qr>2/q. Using β𝝂(r,ϱ~N)=β(νj)j=1N(r,ϱN)\beta_{{\boldsymbol{\nu}}}(r,\tilde{\boldsymbol{\varrho}}_{N})=\beta_{(\nu_{j})_{j=1}^{N}}(r,{\boldsymbol{\varrho}}_{N}) for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F} with supp𝝂{1,,N}\operatorname{supp}{\boldsymbol{\nu}}\subseteq\{1,\dots,N\} we conclude

(β𝝂(r,ϱN)1)𝝂0Nq/2(0N)(\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{N})^{-1})_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\in\ell^{q/2}(\mathbb{N}_{0}^{N})

for any q>0q>0. Now fix q>0q>0 (and 2/q<r2/q<r\in\mathbb{N}). Then, by Hölder’s inequality with s:=2(q/2)/(1+q/2)s:=2(q/2)/(1+q/2), there holds

𝝂0NuN,𝝂Xs\displaystyle\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{s} =𝝂0NuN,𝝂Xsβ𝝂(r,ϱN)s2β𝝂(r,ϱN)s2\displaystyle=\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{s}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{N})^{\frac{s}{2}}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{N})^{-\frac{s}{2}}
(𝝂0NuN,𝝂X2β𝝂(r,ϱN))s2(𝝂0Nβ𝝂(r,ϱN)s2s)2s2,\displaystyle\leq\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\|u_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{N})\Bigg{)}^{\frac{s}{2}}\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\beta_{{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{N})^{\frac{s}{2-s}}\Bigg{)}^{\frac{2-s}{2}},

which is finite since s/(2s)=q/2s/(2-s)=q/2. Thus we have shown

q>0,N:(uN,𝝂X)𝝂0Nq/(1+q/2)(0N).\forall q>0,N\in\mathbb{N}:\quad(\|u_{N,{\boldsymbol{\nu}}}\|_{X})_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\in\ell^{q/(1+q/2)}(\mathbb{N}_{0}^{N})\;.

Finally, due to (ϱj1)jp/(1p)()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{p/(1-p)}(\mathbb{N}), Lemma 3.11 for all NN\in{\mathbb{N}} it holds

(β𝝂(r,ϱN)1)𝝂0Np/(2(1p))(0N)(\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})^{-1})_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}\in\ell^{p/(2(1-p))}(\mathbb{N}_{0}^{N})

and there exists a constant C(𝒃,ξ)C^{\prime}({\boldsymbol{b}},\xi) such that for all NN\in{\mathbb{N}} it holds

(β𝝂(r,ϱN)1)𝝂p/(2(1p))(0N)C(𝒃,ξ)<.\big{\|}(\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{N})^{-1})_{{\boldsymbol{\nu}}}\big{\|}_{\ell^{p/(2(1-p))}(\mathbb{N}_{0}^{N})}\leq C^{\prime}({\boldsymbol{b}},\xi)<\infty\;.

This completes the proofs of (4.10) and of the last statement. ∎

The following result states the sparsity of Wiener-Hermite PC expansion coefficients of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic maps.

Theorem 4.9.

Under the assumptions of Theorem 4.8 it holds

𝝂β𝝂(r,ϱ)u𝝂X2δ2C(𝒃)<with(β𝝂(r,ϱ)1/2)𝝂p/(1p)(),\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{X}^{2}\leq\delta^{2}C({\boldsymbol{b}})<\infty\ \ \ with\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p/(1-p)}({\mathcal{F}}), (4.11)

where C(𝐛)C({\boldsymbol{b}}) is the same constant as in Theorem 4.8 and β𝛎(r,ϱ)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) is given in (3.36). Furthermore,

(u𝝂X)𝝂2p/(2p)().(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{2p/(2-p)}({\mathcal{F}}).
Proof.

Let u~NL2(U,X;γ)\tilde{u}_{N}\in L^{2}(U,X;\gamma) be as in Definition 4.1 and for 𝝂{\boldsymbol{\nu}}\in\mathcal{F} denote by

u~N,𝝂:=Uu~N(𝒚)H𝝂(𝒚)dγ(𝒚)X\tilde{u}_{N,{\boldsymbol{\nu}}}:=\int_{U}\tilde{u}_{N}({\boldsymbol{y}})H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})\in X

the Wiener-Hermite PC expansion coefficient. By Fubini’s theorem

u~N,𝝂=UuN((yj)j=1N)j=1NHνj(yj)dγN((yj)j=1N)=uN,(νj)j=1N\tilde{u}_{N,{\boldsymbol{\nu}}}=\int_{U}u_{N}((y_{j})_{j=1}^{N})\prod_{j=1}^{N}H_{\nu_{j}}(y_{j})\,\mathrm{d}\gamma_{N}((y_{j})_{j=1}^{N})=u_{N,(\nu_{j})_{j=1}^{N}}

for every 𝝂{\boldsymbol{\nu}}\in\mathcal{F} with supp𝝂{1,,N}\operatorname{supp}{\boldsymbol{\nu}}\subseteq\{1,\dots,N\}. Furthermore, since u~N\tilde{u}_{N} is independent of the variables (yj)j=N+1(y_{j})_{j=N+1}^{\infty} we have u~N,𝝂=0\tilde{u}_{N,{\boldsymbol{\nu}}}=0 whenever supp𝝂{1,,N}\operatorname{supp}{\boldsymbol{\nu}}\subsetneq\{1,\dots,N\}. Therefore Theorem 4.8 implies

𝝂β𝝂(r,ϱ)u~N,𝝂X2C(𝒃)δ2N.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|\tilde{u}_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\leq\frac{C({\boldsymbol{b}})}{\delta^{2}}\qquad\forall N\in\mathbb{N}.

Now fix an arbitrary, finite set Λ\Lambda\subset\mathcal{F}. Because of u~NuL2(U,X;γ)\tilde{u}_{N}\to u\in L^{2}(U,X;\gamma) it holds

limNu~N,𝝂=u𝝂\lim_{N\to\infty}\tilde{u}_{N,{\boldsymbol{\nu}}}=u_{{\boldsymbol{\nu}}}

for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F}. Therefore

𝝂Λβ𝝂(r,ϱ)u𝝂X2=limN𝝂Λβ𝝂(r,ϱ)u~N,𝝂X2C(𝒃)δ2.\sum_{{\boldsymbol{\nu}}\in\Lambda}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{{\boldsymbol{\nu}}}\|_{X}^{2}=\lim_{N\to\infty}\sum_{{\boldsymbol{\nu}}\in\Lambda}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|\tilde{u}_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\leq\frac{C({\boldsymbol{b}})}{\delta^{2}}.

Since Λ\Lambda\subset\mathcal{F} was arbitrary, this shows that

𝝂β𝝂(r,ϱ)u𝝂X2δ2C(𝒃)<.\sum_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{X}^{2}\leq\delta^{2}C({\boldsymbol{b}})<\infty.

Finally, due to 𝒃p(){\boldsymbol{b}}\in\ell^{p}({\mathbb{N}}), with

ϱj=bjp1ξ4r!𝒃p\varrho_{j}=b_{j}^{p-1}\frac{\xi}{4\sqrt{r!}\|{\boldsymbol{b}}\|_{\ell^{p}}}

as in Theorem 4.8 we have (ϱj1)jp/(1p)()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{p/(1-p)}(\mathbb{N}). By Lemma 3.11, it holds

(β𝝂(r,ϱ)1/2)𝝂p/(1p)().(\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{p/(1-p)}(\mathcal{F}). (4.12)

The relation (4.11) is proven. Hölder’s inequality can be used to show that (4.12) gives

(u𝝂X)𝝂2p/(2p)()(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{2p/(2-p)}(\mathcal{F})

(by a similar calculation as at the end of the proof of Theorem 4.8 with q=p/(1p)q=p/(1-p)). ∎

Remark 4.10.

We establish the convergence rate of best nn-term approximation of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions based on the p\ell^{p}-summability. Let uu be (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝒃p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) and some p(0,1)p\in(0,1) as in Theorem 4.8. By Theorem 4.9 we then have (u𝝂X)𝝂2p2p(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{\frac{2p}{2-p}}.

Let Λn\Lambda_{n}\subseteq\mathcal{F} be a set of cardinality nn\in\mathbb{N} containing nn multiindices 𝝂{\boldsymbol{\nu}}\in\mathcal{F} such that u𝝁Xu𝝂X\|u_{\boldsymbol{\mu}}\|_{X}\leq\|u_{\boldsymbol{\nu}}\|_{X} whenever 𝝂Λn{\boldsymbol{\nu}}\in\Lambda_{n} and 𝝁Λn{\boldsymbol{\mu}}\notin\Lambda_{n}. Then, by Theorem 4.9, for the truncated the Wiener-Hermite PC expansion we have the error bound

u(𝒚)𝝂Λnu𝝂H𝝂(𝒚)L2(U,X;γ)2=𝝂\Λnu𝝂X2sup𝝂\Λnu𝝂X22p2p𝝁\Λnu𝝁X2p2p.\left\|u({\boldsymbol{y}})-\sum_{{\boldsymbol{\nu}}\in\Lambda_{n}}u_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\right\|_{L^{2}(U,X;\gamma)}^{2}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{n}}\|u_{\boldsymbol{\nu}}\|_{X}^{2}\leq\sup_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{n}}\|u_{\boldsymbol{\nu}}\|_{X}^{2-\frac{2p}{2-p}}\sum_{{\boldsymbol{\mu}}\in\mathcal{F}\backslash\Lambda_{n}}\|u_{\boldsymbol{\mu}}\|_{X}^{\frac{2p}{2-p}}.

For a nonnegative, monotonically decreasing sequence (xj)jq()(x_{j})_{j\in\mathbb{N}}\in\ell^{q}(\mathbb{N}) with q>0q>0 we have

xnq1nj=1nxjqx_{n}^{q}\leq\frac{1}{n}\sum_{j=1}^{n}x_{j}^{q}

and thus

xnn1q(xj)jq().x_{n}\leq n^{-\frac{1}{q}}\|(x_{j})_{j\in\mathbb{N}}\|_{\ell^{q}(\mathbb{N})}.

With q=2p2pq=\frac{2p}{2-p} this implies

(sup𝝂\Λnu𝝂X)22p2p(n2p2p(𝝂u𝝂X2p2p)2p2p)22p2p=𝒪(n2p+2).\left(\sup_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{n}}\|u_{\boldsymbol{\nu}}\|_{X}\right)^{2-\frac{2p}{2-p}}\leq\left(n^{-\frac{2-p}{2p}}\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\|u_{\boldsymbol{\nu}}\|_{X}^{\frac{2p}{2-p}}\right)^{\frac{2-p}{2p}}\right)^{2-\frac{2p}{2-p}}={\mathcal{O}}(n^{-\frac{2}{p}+2}).

Hence, by truncating the Wiener-Hermite PC expansion (4.5) after nn largest terms yields the best nn-term convergence rate

u(𝒚)𝝂Λnu𝝂H𝝂(𝒚)L2(U,X;γ)=𝒪(n1p+1)as n.\left\|u({\boldsymbol{y}})-\sum_{{\boldsymbol{\nu}}\in\Lambda_{n}}u_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\right\|_{L^{2}(U,X;\gamma)}={\mathcal{O}}(n^{-\frac{1}{p}+1})\qquad\text{as }n\to\infty. (4.13)

4.2 (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-Holomorphy of composite functions

We now show that certain composite functions of the type

u(𝒚)=𝒰(exp(jyjψj))u({\boldsymbol{y}})={\mathcal{U}}\bigg{(}\exp\bigg{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\bigg{)}\bigg{)} (4.14)

are (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic under certain conditions.

The significance of such functions is the following: if we think for example of 𝒰{\mathcal{U}} as the solution operator 𝒮\mathcal{S} in (3.5) (for a fixed ff) which maps the diffusion coefficient aL(D)a\in L^{\infty}({{D}}) to the solution 𝒰(a)H01(D){\mathcal{U}}(a)\in H_{0}^{1}({{D}}) of an elliptic PDE on some domain Dd{{D}}\subseteq\mathbb{R}^{d}, then 𝒰(exp(jyjψj)){\mathcal{U}}\big{(}\exp\big{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\big{)}\big{)} is exactly the parametric solution discussed in Sections 3.13.6. We explain this in more detail in Section 4.3.1. The presently developed, abstract setting allows, however, to consider 𝒰{\mathcal{U}} as a solution operator of other, structurally similar PDEs with log-Gaussian random input data. Furthermore, if 𝒢{\mathcal{G}} is another map with suitable holomorphy properties, the composition 𝒢(𝒰(exp(jyjψj))){\mathcal{G}}\big{(}{\mathcal{U}}\big{(}\exp\big{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\big{)}\big{)}\big{)} is again of the general type 𝒰~(exp(jyjψj))\tilde{\mathcal{U}}\big{(}\exp\big{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\big{)}\big{)} with 𝒰~=𝒢𝒰\tilde{\mathcal{U}}={\mathcal{G}}\circ{\mathcal{U}}.

This will allow to apply the ensuing results on convergence rates of deterministic collocation and quadrature algorithms to a wide range of PDEs with GRF inputs and functionals on their random solutions. As a particular case in point, we apply our results to posterior densities in Bayesian inversion, as we explain subsequently in Section 5. As a result, the concept of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy is fairly broad and covers a large range of parametric PDEs depending on log-Gaussian distributed data.

To formalize all of this, we now provide sufficient conditions on the solution operator 𝒰{\mathcal{U}} and the sequence (ψj)j(\psi_{j})_{j\in\mathbb{N}} guaranteeing (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy. Let dd\in\mathbb{N}, Dd{D}\subseteq\mathbb{R}^{d} be an open set and E{E} a complex Banach space which is continuously embedded into L(D;)L^{\infty}({D};\mathbb{C}), and finally let XX be another complex Banach space. Additionally, suppose that there exists CE>0C_{E}>0 such that for all ψ1,ψ2E\psi_{1},\psi_{2}\in{E} and some mm\in\mathbb{N}

exp(ψ1)exp(ψ2)ECEψ1ψ2Emax{exp(mψ1E);exp(mψ2E)}.\|\exp(\psi_{1})-\exp(\psi_{2})\|_{E}\leq C_{E}\|\psi_{1}-\psi_{2}\|_{E}\max\Big{\{}\exp\big{(}m\|\psi_{1}\|_{E}\big{)};\,\exp\big{(}m\|\psi_{2}\|_{E}\big{)}\Big{\}}. (4.15)

This inequality covers in particular the Sobolev spaces Wk(D;)W^{k}_{\infty}({D};\mathbb{C}), k0k\in\mathbb{N}_{0}, on bounded Lipschitz domains Dd{{D}}\subseteq\mathbb{R}^{d}, but also the Kondrat’ev spaces 𝒲k(D;){\mathcal{W}}_{\infty}^{k}({D};\mathbb{C}) on polygonal domains D2{{D}}\subseteq\mathbb{R}^{2}, cp. Lemma 3.28.

For a function ψEL(D;)\psi\in{E}\subseteq L^{\infty}({D};\mathbb{C}) we will write (ψ)L(D;)L(D;)\Re(\psi)\in L^{\infty}({D};\mathbb{R})\subseteq L^{\infty}({D};\mathbb{C}) to denote its real part and (ψ)L(D;)L(D;)\Im({\psi})\in L^{\infty}({D};\mathbb{R})\subseteq L^{\infty}({D};\mathbb{C}) its imaginary part so that ψ=(ψ)+i(ψ)\psi=\Re(\psi)+{\rm i}\Im(\psi). Recall that the quantity ρ(a)\rho(a) is defined in (3.25) for aL(D;)a\in L^{\infty}({D};\mathbb{C}).

Theorem 4.11.

Let 0<δ<δmax0<\delta<\delta_{\rm max}, K>0K>0, η>0\eta>0 and mm\in\mathbb{N}. Let the inequality (4.15) hold for the space EE. Assume that for an open set OEO\subseteq{E} containing

{exp(ψ):ψE,(ψ)Eη},\{\exp(\psi)\,:\,\psi\in{E},\leavevmode\nobreak\ \|\Im(\psi)\|_{E}\leq\eta\},

it holds

  1. (i)

    𝒰:OX{\mathcal{U}}:O\to X is holomorphic,

  2. (ii)

    for all aOa\in O

    𝒰(a)Xδ(1+aEmin{1,ρ(a)})m,\|{\mathcal{U}}(a)\|_{X}\leq\delta\left(\frac{1+\|a\|_{E}}{\min\{1,{\rho(a)}\}}\right)^{m},
  3. (iii)

    for all aa, bOb\in O

    𝒰(a)𝒰(b)XK(1+max{aE,bE}min{1,ρ(a),ρ(b)})mabE,\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{X}\leq K\left(\frac{1+\max\{\|a\|_{E},\|b\|_{E}\}}{\min\{1,{\rho(a)},{\rho(b)}\}}\right)^{m}\|a-b\|_{E},
  4. (iv)

    (ψj)jEL(D)(\psi_{j})_{j\in\mathbb{N}}\subseteq{E}\cap L^{\infty}({D}) and with bj:=ψjEb_{j}:=\|\psi_{j}\|_{E} it holds 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}).

Then there exists ξ>0\xi>0 and for every δmax>0\delta_{\rm max}>0 there exists C~\tilde{C} depending on 𝐛{\boldsymbol{b}}, δmax\delta_{\rm max}, CEC_{E} and mm but independent of δ(0,δmax)\delta\in(0,\delta_{\rm max}), such that with

uN((yj)j=1N)=𝒰(exp(j=1Nyjψj))(yj)j=1NN,u_{N}\left((y_{j})_{j=1}^{N}\right)={\mathcal{U}}\Bigg{(}\exp\bigg{(}\sum_{j=1}^{N}y_{j}\psi_{j}\bigg{)}\Bigg{)}\qquad\forall(y_{j})_{j=1}^{N}\in\mathbb{R}^{N},

and u~N(𝐲)=uN(y1,,yN)\tilde{u}_{N}({\boldsymbol{y}})=u_{N}(y_{1},\dots,y_{N}) for 𝐲U{\boldsymbol{y}}\in U, the function

u:=limNu~NL2(U,X;γ)u:=\lim_{N\to\infty}\tilde{u}_{N}\in L^{2}(U,X;\gamma)

is well-defined and (𝐛,ξ,δC~,X)({\boldsymbol{b}},\xi,\delta\tilde{C},X)-holomorphic.

Proof.

Step 1. Choosing ψ20\psi_{2}\equiv 0 in (4.15) with ψ1=ψ\psi_{1}=\psi, we obtain

exp(ψ)ECEexp((m+1)ψE).\|\exp(\psi)\|_{E}\leq C_{E}^{\prime}\exp\big{(}(m+1)\|\psi\|_{E}\big{)}. (4.16)

for some positive constant CEC_{E}^{\prime}. Indeed,

exp(ψ1)E\displaystyle\|\exp(\psi_{1})\|_{E} 1E+CEψ1Eexp(m1E+mψ1E)\displaystyle\leq\|1\|_{E}+C_{E}\|\psi_{1}\|_{E}\exp\big{(}m\|1\|_{E}+m\|\psi_{1}\|_{E}\big{)}
CE(1+ψ1E)exp(mψ1E)\displaystyle\leq C_{E}^{\prime}\big{(}1+\|\psi_{1}\|_{E}\big{)}\exp\big{(}m\|\psi_{1}\|_{E}\big{)}
CEexp((m+1)ψ1E).\displaystyle\leq C_{E}^{\prime}\exp\big{(}(m+1)\|\psi_{1}\|_{E}\big{)}.

We show that uNL2(N,X;γN)u_{N}\in L^{2}(\mathbb{R}^{N},X;\gamma_{N}) for every NN\in\mathbb{N}. To this end we recall that for any s>0s>0 (see, e.g., [73, Appendix B], [9, (38)] for a proof)

exp(s|y|)dγ1(y)exp(s22+2sπ).\int_{\mathbb{R}}\exp(s|y|)\,\mathrm{d}\gamma_{1}(y)\leq\exp\bigg{(}\frac{s^{2}}{2}+\frac{\sqrt{2}s}{\pi}\bigg{)}. (4.17)

Since E{E} is continuously embedded into L(D;)L^{\infty}({D};\mathbb{C}), there exists C0>0C_{0}>0 such that

ψL(D)C0ψEψE.\displaystyle\|\psi\|_{L^{\infty}({{D}})}\leq C_{0}\|\psi\|_{E}\qquad\forall\psi\in E.

Using (ii), (4.16), and

1essinf𝒙D(exp(j=1Nyjψj(𝒙)))\displaystyle\frac{1}{\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{D}}\big{(}\exp\big{(}\sum_{j=1}^{N}y_{j}\psi_{j}({\boldsymbol{x}})\big{)}\big{)}} exp(j=1Nyjψj)L\displaystyle\leq\bigg{\|}\exp\Big{(}-\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{L^{\infty}}
exp(j=1NyjψjL)exp(C0j=1N|yj|ψjE),\displaystyle\leq\exp\bigg{(}\bigg{\|}\sum_{j=1}^{N}y_{j}\psi_{j}\bigg{\|}_{L^{\infty}}\bigg{)}\leq\exp\bigg{(}C_{0}\sum_{j=1}^{N}|y_{j}|\big{\|}\psi_{j}\big{\|}_{E}\bigg{)},

we obtain the bound

uN(𝒚)X\displaystyle\|u_{N}({\boldsymbol{y}})\|_{X} δ(1+exp(j=1Nyjψj)E)mexp(C0mj=1N|yj|ψjE)\displaystyle\leq\delta\bigg{(}1+\bigg{\|}\exp\bigg{(}\sum_{j=1}^{N}y_{j}\psi_{j}\bigg{)}\bigg{\|}_{E}\bigg{)}^{m}\exp\bigg{(}C_{0}m\sum_{j=1}^{N}|y_{j}|\big{\|}\psi_{j}\big{\|}_{E}\bigg{)}
δ(1+CEexp((m+1)j=1N|yj|ψjE))mexp(C0mj=1N|yj|ψjE)\displaystyle\leq\delta\Bigg{(}1+C_{E}^{\prime}\exp\bigg{(}(m+1)\sum_{j=1}^{N}|y_{j}|\left\|\psi_{j}\right\|_{E}\bigg{)}\Bigg{)}^{m}\exp\bigg{(}C_{0}m\sum_{j=1}^{N}|y_{j}|\big{\|}\psi_{j}\big{\|}_{E}\bigg{)}
C1exp((2+C0)m2j=1N|yj|ψjE)\displaystyle\leq C_{1}\exp\bigg{(}(2+C_{0})m^{2}\sum_{j=1}^{N}|y_{j}|\left\|\psi_{j}\right\|_{E}\bigg{)}

for some constant C1>0C_{1}>0 depending on δ,CE\delta,C_{E} and mm. Hence, by (4.17) we have

NuN(𝒚)X2dγN(𝒚)\displaystyle\int_{\mathbb{R}^{N}}\|u_{N}({\boldsymbol{y}})\|_{X}^{2}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}}) C1Nexp((2+C0)m2)j=1N|yj|ψjE)dγN(𝒚)\displaystyle\leq C_{1}\int_{\mathbb{R}^{N}}\exp\bigg{(}(2+C_{0})m^{2})\sum_{j=1}^{N}|y_{j}|\|\psi_{j}\|_{E}\bigg{)}\,\mathrm{d}\gamma_{N}({\boldsymbol{y}})
C1exp((2+C0)2m42j=1Nbj2+2(2+C0)m2πj=1Nbj)<.\displaystyle\leq C_{1}\exp\Bigg{(}\frac{(2+C_{0})^{2}m^{4}}{2}\sum_{j=1}^{N}b_{j}^{2}+\frac{\sqrt{2}(2+C_{0})m^{2}}{\pi}\sum_{j=1}^{N}b_{j}\Bigg{)}<\infty.

Step 2. We show that (u~N)N(\tilde{u}_{N})_{N\in\mathbb{N}} which is defined as u~N(𝒚):=uN(y1,,yN)\tilde{u}_{N}({\boldsymbol{y}}):=u_{N}(y_{1},\dots,y_{N}) for 𝒚U{\boldsymbol{y}}\in U, is a Cauchy sequence in L2(U,X;γ)L^{2}(U,X;\gamma). For any N<MN<M by (iii)

u~Mu~NL2(U,X;γ)2\displaystyle\|\tilde{u}_{M}-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}^{2} =U𝒰(exp(j=1Myjψj))𝒰(exp(j=1Nyjψj))X2dγ(𝒚)\displaystyle=\int_{U}\bigg{\|}{\mathcal{U}}\bigg{(}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}\bigg{)}-{\mathcal{U}}\bigg{(}\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{)}\bigg{\|}_{X}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})
KU[(1+exp(j=1Myjψj)E+exp(j=1Nyjψj)E)m\displaystyle\quad\leq K\int_{U}\Bigg{[}\Bigg{(}1+\bigg{\|}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E}+\bigg{\|}\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E}\Bigg{)}^{m}
×exp(C0mj=1M|yj|ψjE)exp(j=1Myjψj)exp(j=1Nyjψj)E]dγ(𝒚).\displaystyle\quad\times\exp\bigg{(}C_{0}m\sum_{j=1}^{M}|y_{j}|\big{\|}\psi_{j}\big{\|}_{E}\bigg{)}\cdot\bigg{\|}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}-\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E}\Bigg{]}\,\mathrm{d}\gamma({\boldsymbol{y}}).

Using (4.16) again we can estimate

u~Mu~NL2(U,X;γ)2KU[(1+2CEexp((m+1)j=1M|yj|ψjE))m\displaystyle\|\tilde{u}_{M}-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}^{2}\leq K\int_{U}\Bigg{[}\Bigg{(}1+2C_{E}^{\prime}\exp\bigg{(}(m+1)\sum_{j=1}^{M}|y_{j}|\left\|\psi_{j}\right\|_{E}\bigg{)}\Bigg{)}^{m}
×exp(C0mj=1M|yj|ψjE)exp(j=1Myjψj)exp(j=1Nyjψj)E]dγ(𝒚)\displaystyle\times\exp\bigg{(}C_{0}m\sum_{j=1}^{M}|y_{j}|\big{\|}\psi_{j}\big{\|}_{E}\bigg{)}\cdot\bigg{\|}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}-\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E}\Bigg{]}\,\mathrm{d}\gamma({\boldsymbol{y}})
C2U[exp((2+C0)m2)j=1M|yj|ψjE)exp(j=1Myjψj)exp(j=1Nyjψj)E]dγ(𝒚)\displaystyle\leq C_{2}\int_{U}\Bigg{[}\exp\bigg{(}(2+C_{0})m^{2})\sum_{j=1}^{M}|y_{j}|\left\|\psi_{j}\right\|_{E}\bigg{)}\bigg{\|}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}-\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E}\Bigg{]}\,\mathrm{d}\gamma({\boldsymbol{y}})\,

for C2>0C_{2}>0 depending only on K,CEK,C_{E}, and mm. Now, employing (4.15) we obtain

exp(j=1Myjψj)exp(j=1Nyjψj)E\displaystyle\bigg{\|}\exp\Big{(}\sum_{j=1}^{M}y_{j}\psi_{j}\Big{)}-\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{\|}_{E} CEj=N+1M|yj|ψjEexp(mj=1M|yj|ψjE)\displaystyle\leq C_{E}\sum_{j=N+1}^{M}|y_{j}|\|\psi_{j}\|_{E}\exp\Bigg{(}m\sum_{j=1}^{M}|y_{j}|\|\psi_{j}\|_{E}\Bigg{)}
CEj=N+1M|yj|ψjEexp(m2j=1M|yj|ψjE).\displaystyle\leq C_{E}\sum_{j=N+1}^{M}|y_{j}|\|\psi_{j}\|_{E}\exp\bigg{(}m^{2}\sum_{j=1}^{M}|y_{j}|\|\psi_{j}\|_{E}\bigg{)}.

Therefore for a constant C3C_{3} depending on CEC_{E} and δ\delta (but independent of NN), using |yj|exp(|yj|)|y_{j}|\leq\exp(|y_{j}|),

u~Mu~NL2(U,X;γ)2\displaystyle\|\tilde{u}_{M}-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}^{2}
C3j=N+1MψjEM|yj|exp((3+C0)m2i=1M|yi|ψiE)dγM((yi)i=1M)\displaystyle\leq C_{3}\sum_{j=N+1}^{M}\|\psi_{j}\|_{E}\int_{\mathbb{R}^{M}}|y_{j}|\exp\bigg{(}(3+C_{0})m^{2}\sum_{i=1}^{M}|y_{i}|\|\psi_{i}\|_{E}\bigg{)}\,\mathrm{d}\gamma_{M}((y_{i})_{i=1}^{M})
C3j=N+1MψjEMexp(|yj|+(3+C0)m2i=1M|yi|ψiE)dγM((yi)i=1M)\displaystyle\leq C_{3}\sum_{j=N+1}^{M}\|\psi_{j}\|_{E}\int_{\mathbb{R}^{M}}\exp\bigg{(}|y_{j}|+(3+C_{0})m^{2}\sum_{i=1}^{M}|y_{i}|\|\psi_{i}\|_{E}\bigg{)}\,\mathrm{d}\gamma_{M}((y_{i})_{i=1}^{M})
C3(j=N+1Mbj)(exp(12+2π+(3+C0)2m42i=1Mbj2+2(3+C0)m2πj=1Mbj)),\displaystyle\leq C_{3}\bigg{(}\sum_{j=N+1}^{M}b_{j}\bigg{)}\Bigg{(}\exp\bigg{(}\frac{1}{2}+\frac{\sqrt{2}}{\pi}+\frac{(3+C_{0})^{2}m^{4}}{2}\sum_{i=1}^{M}b_{j}^{2}+\frac{\sqrt{2}(3+C_{0})m^{2}}{\pi}\sum_{j=1}^{M}b_{j}\bigg{)}\Bigg{)},

where we used (4.17) and in the last inequality. Since 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}) the last term is bounded by C4(j=N+1bj)C_{4}\Big{(}\sum_{j=N+1}^{\infty}b_{j}\Big{)} for a constant C4C_{4} depending on CEC_{E}, KK and 𝒃{\boldsymbol{b}} but independent of NN, MM. Due to 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}), it also holds

j=N+1bj0asN.\sum_{j=N+1}^{\infty}b_{j}\to 0\ \ {\rm as}\ \ N\to\infty.

Since N<MN<M are arbitrary, we have shown that (u~N)N(\tilde{u}_{N})_{N\in\mathbb{N}} is a Cauchy sequence in the Banach space L2(U,X;γ)L^{2}(U,X;\gamma). This implies that there is a function

u:=limNu~NL2(U,X;γ).u:=\lim_{N\to\infty}\tilde{u}_{N}\in L^{2}(U,X;\gamma).

Step 3. To show that uu is (𝒃,ξ,δC~,X)({\boldsymbol{b}},\xi,\delta\tilde{C},X) holomorphic, we provide constants ξ>0\xi>0 and C~>0\tilde{C}>0 independent of δ\delta so that uNu_{N} admits holomorphic extensions as in Definition 4.1. This concludes the proof.

Let ξ:=π/(4C0)\xi:=\pi/(4C_{0}). Fix NN\in\mathbb{N} and assume

j=1Nbjϱj<ξ\sum_{j=1}^{N}b_{j}\varrho_{j}<\xi

(i.e. (ϱj)j=1N(\varrho_{j})_{j=1}^{N} is (𝒃,δ1)({\boldsymbol{b}},\delta_{1})-admissible). Then for zj=yj+iζjz_{j}=y_{j}+{\rm i}\zeta_{j}\in\mathbb{C} such that |(zj)|=|ζj|<ϱj|\Im(z_{j})|=|\zeta_{j}|<\varrho_{j} for all jj,

ρ(exp(j=1Nzjψj(𝒙)))=essinf𝒙D(exp(j=1Nyjψj(𝒙)))cos(j=1Nζjψj(𝒙)).\rho\Bigg{(}\exp\Big{(}\sum_{j=1}^{N}z_{j}\psi_{j}({\boldsymbol{x}})\Big{)}\Bigg{)}=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,\Bigg{(}\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}({\boldsymbol{x}})\Big{)}\Bigg{)}\cos\Bigg{(}\sum_{j=1}^{N}\zeta_{j}\psi_{j}({\boldsymbol{x}})\Bigg{)}.

Due to

esssup𝒙D|j=1Nζjψj(𝒙)|j=1NϱjψjLj=1NC0ϱjψjV=j=1NC0ϱjbjπ4,\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,sup}}\,\Bigg{|}\sum_{j=1}^{N}\zeta_{j}\psi_{j}({\boldsymbol{x}})\Bigg{|}\leq\sum_{j=1}^{N}\varrho_{j}\|\psi_{j}\|_{L^{\infty}}\leq\sum_{j=1}^{N}C_{0}\varrho_{j}\|\psi_{j}\|_{V}=\sum_{j=1}^{N}C_{0}\varrho_{j}b_{j}\leq\frac{\pi}{4},

we obtain for such (zj)j=1N(z_{j})_{j=1}^{N}

ρ(exp(j=1Nzjψj(𝒙)))exp(j=1N|yj|ψjL)cos(π4)>0.\rho\Bigg{(}\exp\Big{(}\sum_{j=1}^{N}z_{j}\psi_{j}({\boldsymbol{x}})\Big{)}\Bigg{)}\geq\exp\Bigg{(}-\sum_{j=1}^{N}|y_{j}|\|\psi_{j}\|_{L^{\infty}}\Bigg{)}\cos\Bigg{(}\frac{\pi}{4}\Bigg{)}>0. (4.18)

This shows that for every ϱ=(ϱj)j=1N(0,)N{\boldsymbol{\varrho}}=(\varrho_{j})_{j=1}^{N}\in(0,\infty)^{N} such that j=1Nbjϱj<ξ\sum_{j=1}^{N}b_{j}\varrho_{j}<\xi, it holds

j=1NzjψjO𝒛𝒮(ϱ).\sum_{j=1}^{N}z_{j}\psi_{j}\in O\quad\forall{\boldsymbol{z}}\in{\mathcal{S}}({\boldsymbol{\varrho}}).

Since 𝒰:OX{\mathcal{U}}:O\to X is holomorphic, the function

uN((yj)j=1N)=𝒰(exp(j=1Nyjψj))u_{N}\left((y_{j})_{j=1}^{N}\right)={\mathcal{U}}\Bigg{(}\exp\bigg{(}\sum_{j=1}^{N}y_{j}\psi_{j}\bigg{)}\Bigg{)}

can be holomorphically extended to arguments (zj)j=1N𝒮(ϱ)(z_{j})_{j=1}^{N}\in{\mathcal{S}}({\boldsymbol{\varrho}}).

Finally we fix again NN\in\mathbb{N} and provide a function φNL2(U;γ)\varphi_{N}\in L^{2}(U;\gamma) as in Definition 4.1. Fix 𝒚N{\boldsymbol{y}}\in\mathbb{R}^{N} and 𝒛ϱ{\boldsymbol{z}}\in{\mathcal{B}}_{\boldsymbol{\varrho}} and set

a:=j=1N(yj+zj)ψj.a:=\sum_{j=1}^{N}(y_{j}+z_{j})\psi_{j}.

By (ii), (4.18) and because bj=ψjEb_{j}=\|\psi_{j}\|_{E} and

j=1Nbjϱjξ,\sum_{j=1}^{N}b_{j}\varrho_{j}\leq\xi,

we have that

uN((yj+zj)j=1N)X\displaystyle\|u_{N}((y_{j}+z_{j})_{j=1}^{N})\|_{X} δ(1+aEmin{1,ρ(a)})m\displaystyle\leq\delta\left(\frac{1+\|a\|_{E}}{\min\{1,\rho(a)\}}\right)^{m}
δ(1+CEexp((m+1)j=1N(|yj|+|zj|)ψjE)exp(C0j=1N(|yj|+|zj|)ψjE)cos(π4))m\displaystyle\leq\delta\left(\frac{1+C_{E}^{\prime}\exp\big{(}(m+1)\sum_{j=1}^{N}(|y_{j}|+|z_{j}|)\|\psi_{j}\|_{E}\big{)}}{\exp(-C_{0}\sum_{j=1}^{N}(|y_{j}|+|z_{j}|)\|\psi_{j}\|_{E})\cos(\frac{\pi}{4})}\right)^{m}
δ(1+CEexp((m+1)j=1N|yj|bj)exp((m+1)ξ)exp(C0j=1N|yj|bj)exp(C0ξ)cos(π4))m\displaystyle\leq\delta\left(\frac{1+C_{E}^{\prime}\exp\big{(}(m+1)\sum_{j=1}^{N}|y_{j}|b_{j}\big{)}\exp((m+1)\xi)}{\exp(-C_{0}\sum_{j=1}^{N}|y_{j}|b_{j})\exp(-C_{0}\xi)\cos(\frac{\pi}{4})}\right)^{m}
δLexp((2+C0)m2j=1N|yj|bj)\displaystyle\leq\delta L\exp\bigg{(}(2+C_{0})m^{2}\sum_{j=1}^{N}|y_{j}|b_{j}\bigg{)}

for some LL depending only on CE,C0C_{E},C_{0} and mm. Let us define the last quantity as φN((yj)j=1N)\varphi_{N}\left((y_{j})_{j=1}^{N}\right). Then by (4.17) and because γN\gamma_{N} is a probability measure on N\mathbb{R}^{N},

φNL2(N;γN)\displaystyle\|\varphi_{N}\|_{L^{2}(\mathbb{R}^{N};\gamma_{N})} δLexp(j=1N(2+C0)2m4bj22+(2+C0)m22bjπ)\displaystyle\leq\delta L\exp\bigg{(}\sum_{j=1}^{N}\frac{(2+C_{0})^{2}m^{4}b_{j}^{2}}{2}+(2+C_{0})m^{2}\frac{\sqrt{2}b_{j}}{\pi}\bigg{)}
δLexp(j(2+C0)2m4bj22+(2+C0)m2bjπ)\displaystyle\leq\delta L\exp\bigg{(}\sum_{j\in\mathbb{N}}\frac{(2+C_{0})^{2}m^{4}b_{j}^{2}}{2}+(2+C_{0})m\frac{\sqrt{2}b_{j}}{\pi}\bigg{)}
δC~(𝒃,C0,CE,m),\displaystyle\leq\delta\tilde{C}({\boldsymbol{b}},C_{0},C_{E},m),

for some constant C~(𝒃,C0,CE,m)(0,)\tilde{C}({\boldsymbol{b}},C_{0},C_{E},m)\in(0,\infty) because 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}). In all, we have shown that uu satisfies (𝒃,ξ,δC~,X)({\boldsymbol{b}},\xi,\delta\tilde{C},X)-holomorphy as in Definition 4.1. ∎

4.3 Examples of holomorphic data-to-solution maps

We revisit the example of linear elliptic divergence-form PDE with diffusion coefficient introduced in Section 3. Its coefficient-to-solution map SS from (3.5) for a fixed fXf\in X^{\prime}, gives rise to parametric maps which are parametric-holomorphic. This kind of function will, on the one hand, arise as generic model of Banach-space valued uncertain inputs of PDEs, and on the other hand as model of solution manifolds of PDEs. The connection is made through preservation of holomorphy under composition with inversion of boundedly invertible differential operators.

Let fXf\in X^{\prime} be given. If A(a)is(X,X)A(a)\in\mathcal{L}_{{\rm is}}(X,X^{\prime}) is an isomorphism depending (locally) holomorphically on aEa\in E, then

𝒰:EX:a(invA(a))f{\mathcal{U}}:E\to X:a\mapsto(\mathrm{inv}\circ A(a))f

is also locally holomorphic as a function of aEa\in{E}. Here inv\mathrm{inv} denotes the inversion map. This is a consequence of the fact that the inv:is(X,X)is(X,X)\mathrm{inv}:\mathcal{L}_{{\rm is}}(X,X^{\prime})\to\mathcal{L}_{{\rm is}}(X^{\prime},X) is holomorphic, see e.g. [111, Example 1.2.38]. This argument can be used to show that the solution operator corresponding to the solution of certain PDEs is holomorphic in the parameter. We informally discuss this for some parametric PDEs and refer to [111, Chapter 1 and 5] for more details.

4.3.1 Linear elliptic divergence-form PDE with parametric diffusion coefficient

Let us again consider the model linear elliptic PDE

div(a𝒰(a))=fin D,𝒰(a)=0on D-\operatorname{div}(a\nabla{\mathcal{U}}(a))=f\;\;\text{in }{{D}}\;,\quad{\mathcal{U}}(a)=0\;\;\text{on }\partial{{D}} (4.19)

where dd\in\mathbb{N}, Dd{{D}}\subseteq\mathbb{R}^{d} is a bounded Lipschitz domain, X:=H01(D;)X:=H_{0}^{1}({{D}};\mathbb{C}), fH1(D;):=(H01(D;))f\in H^{-1}({{D}};\mathbb{C}):=(H_{0}^{1}({{D}};\mathbb{C}))^{\prime} and aE:=L(D;)a\in{E}:=L^{\infty}({{D}};\mathbb{C}). Then the solution operator 𝒰:OX{\mathcal{U}}:O\to X maps the coefficient function aa to the weak solution 𝒰(a){\mathcal{U}}(a), where

O:={aL(D;):ρ(a)>0},O:=\{a\in L^{\infty}({{D}};\mathbb{C})\,:\,\rho(a)>0\},

with ρ(a)\rho(a) defined in (3.25) for aL(D;)a\in L^{\infty}({D};\mathbb{C}). With A(a)A(a) denoting the differential operator div(a)L(X,X)-\operatorname{div}(a\nabla\cdot)\in L(X,X^{\prime}) we can also write 𝒰(a)=A(a)1f{\mathcal{U}}(a)=A(a)^{-1}f. We now check assumptions (i)(iii) of Theorem 4.11.

  1. (i)

    As mentioned above, complex Fréchet differentiability (i.e. holomorphy) of 𝒰:OX{\mathcal{U}}:O\to X is satisfied because the operation of inversion of linear operators is holomorphic on the set of boundedly invertible linear operators, AA depends boundedly and linearly (thus holomorphically) on aa, and therefore, the map

    aA(a)1f=𝒰(a)a\mapsto A(a)^{-1}f={\mathcal{U}}(a)

    is a composition of holomorphic functions. We refer once more to [111, Example 1.2.38] for more details.

  2. (ii)

    For aOa\in O, it holds

    𝒰(a)X2ρ(a)|D𝒰(a)a𝒰(a)¯d𝒙|=|f,𝒰(a)¯|fX𝒰(a)X.\|{\mathcal{U}}(a)\|_{X}^{2}\,\rho(a)\leq\left|\int_{{D}}\nabla{\mathcal{U}}(a)^{\top}a\overline{\nabla{\mathcal{U}}(a)}\,\mathrm{d}{\boldsymbol{x}}\right|=\left|\left\langle f,\overline{{\mathcal{U}}(a)}\right\rangle\right|\leq\|f\|_{X^{\prime}}\|{\mathcal{U}}(a)\|_{X}.

    Here ,\left\langle\cdot,\cdot\right\rangle denotes the dual product between XX^{\prime} and XX. This gives the usual a-priori bound

    𝒰(a)XfXρ(a).\|{\mathcal{U}}(a)\|_{X}\leq\frac{\|f\|_{X^{\prime}}}{\rho(a)}. (4.20)
  3. (iii)

    For aa, bOb\in O and with w:=𝒰(a)𝒰(b)w:={\mathcal{U}}(a)-{\mathcal{U}}(b), we have that

    wX2ρ(a)\displaystyle\frac{\|w\|_{X}^{2}}{\rho(a)} |Dwaw¯d𝒙|\displaystyle\leq\left|\int_{{D}}\nabla w^{\top}a\overline{\nabla w}\,\mathrm{d}{\boldsymbol{x}}\right|
    =|D𝒰(a)aw¯d𝒙D𝒰(b)bw¯d𝒙D𝒰(b)(ab)w¯d𝒙|\displaystyle=\left|\int_{{D}}\nabla{\mathcal{U}}(a)^{\top}a\overline{\nabla w}\,\mathrm{d}{\boldsymbol{x}}-\int_{{D}}\nabla{\mathcal{U}}(b)^{\top}b\overline{\nabla w}\,\mathrm{d}{\boldsymbol{x}}-\int_{{D}}\nabla{\mathcal{U}}(b)^{\top}(a-b)\overline{\nabla w}\,\mathrm{d}{\boldsymbol{x}}\right|
    𝒰(b)XwXabE\displaystyle\leq\|{\mathcal{U}}(b)\|_{X}\|w\|_{X}\|a-b\|_{E}
    fXρ(b)wXabE,\displaystyle\leq\frac{\|f\|_{X^{\prime}}}{\rho(b)}\|w\|_{X}\|a-b\|_{E},

    and thus

    𝒰(a)𝒰(b)XfXaEρ(b)abE.\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{X}\leq\|f\|_{X^{\prime}}\frac{\|a\|_{E}}{\rho(b)}\|a-b\|_{E}. (4.21)

Hence, if (ψj)jE(\psi_{j})_{j\in\mathbb{N}}\subset{E} such that with bj:=ψjEb_{j}:=\|\psi_{j}\|_{E} it holds 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}), then the solution

u(𝒚)=limN𝒰(exp(j=1Nyjψj))L2(U,X;γ)u({\boldsymbol{y}})=\lim_{N\to\infty}{\mathcal{U}}\left(\exp\left(\sum_{j=1}^{N}y_{j}\psi_{j}\right)\right)\in L^{2}(U,X;\gamma)

is well-defined and (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic by Theorem 4.11.

This example can easily be generalized to spaces of higher-regularity, e.g., if Dd{{D}}\subseteq\mathbb{R}^{d} is a bounded Cs1C^{s-1} domain for some ss\in\mathbb{N}, s2s\geq 2, then we may set X:=H01(D;)Hs(D;)X:=H_{0}^{1}({{D}};\mathbb{C})\cap H^{s}({{D}};\mathbb{C}) and E:=Ws(D;){E}:=W^{s}_{\infty}({{D}};\mathbb{C}) and repeat the above calculation.

4.3.2 Linear parabolic PDE with parametric coefficient

Let 0<T<0<T<\infty denote a finite time-horizon and let D{{D}} be a bounded domain with Lipschitz boundary D\partial{{D}} in d{\mathbb{R}}^{d}. We define I:=(0,T)I:=(0,T) and consider the initial boundary value problem (IBVP for short) for the linear parabolic PDE

{u(t,𝒙)tdiv(a(𝒙)u(t,𝒙))=f(t,𝒙),(t,𝒙)I×D,u|D×I=0,u|t=0=u0(𝒙).\begin{cases}\frac{\partial u(t,{\boldsymbol{x}})}{\partial t}-\operatorname{div}\big{(}a({\boldsymbol{x}})\nabla u(t,{\boldsymbol{x}})\big{)}=f(t,{\boldsymbol{x}}),\qquad(t,{\boldsymbol{x}})\in I\times{{D}},\\ u|_{\partial{{D}}\times I}=0,\\ u|_{t=0}=u_{0}({\boldsymbol{x}}).\end{cases} (4.22)

In this section, we prove that the solution to this problem satisfies the assumptions of Theorem 4.11 for certain spaces EE and XX. We first review results on the existence and uniqueness of solutions to the equation (4.22). We refer to [97] and the references there for proofs and more detailed discussion.

We denote V:=H01(D;)V:=H_{0}^{1}({{D}};\mathbb{C}) and V:=H1(D;)V^{\prime}:=H^{-1}({{D}};\mathbb{C}). The parabolic IBPV given by equation (4.22) is a well-posed operator equation in the intersection space of Bochner spaces (e.g. [97, Appendix], and e.g. [110, 53] for the definition of spaces)

X:=L2(I,V)H1(I,V)=(L2(I)V)(H1(I)V)X:=L^{2}(I,V)\cap H^{1}(I,V^{\prime})=\big{(}L^{2}(I)\otimes V\big{)}\cap\big{(}H^{1}(I)\otimes V^{\prime}\big{)}

equipped with the sum norm

uX:=(uL2(I,V)2+uH1(I,V)2)1/2,uX,\|u\|_{X}:=\Big{(}\|u\|_{L^{2}(I,V)}^{2}+\|u\|_{H^{1}(I,V^{\prime})}^{2}\Big{)}^{1/2},\qquad u\in X,

where

uL2(I,V)2=Iu(t,)V2dt,\|u\|_{L^{2}(I,V)}^{2}=\int_{I}\|u(t,\cdot)\|_{V}^{2}\,\,\mathrm{d}t\,,

and

uH1(I,V)2=Itu(t,)V2dt.\|u\|_{H^{1}(I,V^{\prime})}^{2}=\int_{I}\|\partial_{t}u(t,\cdot)\|_{V^{\prime}}^{2}\,\,\mathrm{d}t\,.

To state a space-time variational formulation and to specify the data space for (4.22), we introduce the test-function space

Y=L2(I,V)×L2(D)=(L2(I)V)×L2(D)Y=L^{2}(I,V)\times L^{2}({D})=\big{(}L^{2}(I)\otimes V\big{)}\times L^{2}({D})

which we endow with the norm

vY=(v1L2(I,V)2+v2L2(D)2)1/2,v=(v1,v2)Y.\|v\|_{Y}=\Big{(}\|v_{1}\|_{L^{2}(I,V)}^{2}+\|v_{2}\|_{L^{2}({D})}^{2}\Big{)}^{1/2},\qquad v=(v_{1},v_{2})\in Y\,.

Given a time-independent diffusion coefficient aL(D;)a\in L^{\infty}({{D}};\mathbb{C}) and (f,u0)Y(f,u_{0})\in Y^{\prime}, the continuous sesquilinear and antilinear forms corresponding to the parabolic problem (4.22) reads for uXu\in X and v=(v1,v2)Yv=(v_{1},v_{2})\in Y as

B(u,v;a):=IDtuv1¯d𝒙dt+IDauv1¯d𝒙dt+Du0v2¯d𝒙\begin{split}B(u,v;a)&:=\int_{I}\int_{{D}}\partial_{t}u\,\overline{v_{1}}\,\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}t+\int_{I}\int_{{D}}a\nabla u\cdot\overline{\nabla v_{1}}\,\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}t+\int_{{D}}u_{0}\,\overline{v_{2}}\,\mathrm{d}{\boldsymbol{x}}\end{split}

and

L(v):=If(t,),v1(t,)dt+Du0v2¯d𝒙,\begin{split}L(v):=\ \int_{I}\big{\langle}f(t,\cdot),v_{1}(t,\cdot)\big{\rangle}\,\mathrm{d}t+\int_{{D}}u_{0}\,\overline{v_{2}}\,\mathrm{d}{\boldsymbol{x}},\end{split}

where ,\langle\cdot,\cdot\rangle is the anti-duality pairing between VV^{\prime} and VV. Then the space-time variational formulation of equation (4.22) is: Find 𝒰(a)X{\mathcal{U}}(a)\in X such that

B(𝒰(a),v;a)=L(v),vY.B({\mathcal{U}}(a),v;a)=L(v),\quad\forall v\in Y\,. (4.23)

The existence and uniqueness of solution to the equation (4.23) was proved in [97] which reads as follows.

Proposition 4.12.

Assume that (f,u0)Y(f,u_{0})\in Y^{\prime} and that

0<ρ(a):=essinf𝒙D(a(𝒙))|a(𝒙)|aL<,𝒙D0<\rho(a):=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,\Re(a({\boldsymbol{x}}))\leq|a({\boldsymbol{x}})|\leq\|a\|_{L^{\infty}}<\infty,\qquad{\boldsymbol{x}}\in{D} (4.24)

Then the parabolic operator (X,Y)\mathcal{B}\in\mathcal{L}(X,Y^{\prime}) defined by

(u)(v)=B(u,v;a),(\mathcal{B}u)(v)=B(u,v;a),

is an isomorphism and 1:YX\mathcal{B}^{-1}:Y\to X has the norm

11β(a),\|\mathcal{B}^{-1}\|\leq\frac{1}{\beta(a)},

where

β(a):=min(ρ(a)aL2,ρ(a))2max(ρ(a)2,1)+ϑ2andϑ:=supw0,wXw(0,)L2(D)wX.\beta(a):=\frac{\min\big{(}\rho(a)\|a\|_{L^{\infty}}^{-2},\rho(a)\big{)}}{\sqrt{2\max(\rho(a)^{-2},1)+\vartheta^{2}}}\qquad\text{and}\qquad\vartheta:=\sup_{w\not=0,w\in X}\frac{\|w(0,\cdot)\|_{L^{2}({D})}}{\|w\|_{X}}\,.

The constant ϑ\vartheta depends only on TT.

The data space for the equation (4.22) for complex-valued data is E:=L(D,)E:=L^{\infty}({D},{\mathbb{C}}). With the set of admissible diffusion coefficients in the data space

O:={aL(D,):ρ(a)>0},O:=\{a\in L^{\infty}({{D}},\mathbb{C})\,:\,\rho(a)>0\},

from the above proposition we immediately deduce that for given (f,u0)Y(f,u_{0})\in Y^{\prime}, the map

𝒰:OX:a𝒰(a){\mathcal{U}}:O\to X:a\ \mapsto\ {\mathcal{U}}(a)

is well-defined.

Furthermore, there holds the a-priori estimate

𝒰(a)X1β(a)(fL2(I,V)2+u0L22)1/2.\|{\mathcal{U}}(a)\|_{X}\leq\frac{1}{\beta(a)}\Big{(}\|f\|_{L^{2}(I,V^{\prime})}^{2}+\|u_{0}\|_{L^{2}}^{2}\Big{)}^{1/2}\,. (4.25)

This bound is a consequence of the following result which states that the data-to-solution map a𝒰(a)a\to{\mathcal{U}}(a) is locally Lipschitz continuous.

Lemma 4.13.

Let (f,u0)Y(f,u_{0})\in Y^{\prime}. Assume that 𝒰(a){\mathcal{U}}(a) and 𝒰(b){\mathcal{U}}(b) be solutions to (4.23) with coefficients aa, bb satisfying (4.24), respectively.

Then, with the function β()\beta(\cdot) in variable aa as in Proposition 4.12, we have

𝒰(a)𝒰(b)X1β(a)β(b)abL(fL2(I,V)2+u0L22)1/2.\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{X}\leq\frac{1}{\beta(a)\beta(b)}\|a-b\|_{L_{\infty}}\Big{(}\|f\|_{L^{2}(I,V^{\prime})}^{2}+\|u_{0}\|_{L^{2}}^{2}\Big{)}^{1/2}\,.
Proof.

From (4.23) we find that for w:=𝒰(a)𝒰(b)w:={\mathcal{U}}(a)-{\mathcal{U}}(b),

IDtwv1¯d𝒙dt+IDawv1¯d𝒙dt+Dw|t=0v2¯d𝒙=ID(ab)𝒰(b)v1¯d𝒙dt.\begin{split}\int_{I}\int_{{D}}\partial_{t}w\,\overline{v_{1}}\,\mathrm{d}{\boldsymbol{x}}\,\,\mathrm{d}t&+\int_{I}\int_{{D}}a\nabla w\cdot\overline{\nabla v_{1}}\,\,\mathrm{d}{\boldsymbol{x}}\,\,\mathrm{d}t+\int_{{D}}w\big{|}_{t=0}\overline{v_{2}}\,\mathrm{d}{\boldsymbol{x}}\\ &=-\int_{I}\int_{{D}}\big{(}a-b\big{)}\nabla{\mathcal{U}}(b)\cdot\overline{\nabla v_{1}}\,\,\mathrm{d}{\boldsymbol{x}}\,\,\mathrm{d}t\,.\end{split}

This is a parabolic equation in the variational form with (f~,0)Y(\tilde{f},0)\in Y^{\prime} where f~:L2(I,V)\tilde{f}:L^{2}(I,V)\to\mathbb{C} is given by

f~(v1):=ID(ab)𝒰(b)v1¯d𝒙dt,v1L2(I,V).\tilde{f}(v_{1}):=-\int_{I}\int_{{D}}\big{(}a-b\big{)}\nabla{\mathcal{U}}(b)\cdot\overline{\nabla v_{1}}\,\,\mathrm{d}{\boldsymbol{x}}\,\,\mathrm{d}t\,,\qquad v_{1}\in L^{2}(I,V).

Now applying Proposition 4.12 we find

𝒰(a)𝒰(b)Xf~L2(I,V)β(a).\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{X}\leq\frac{\|\tilde{f}\|_{L^{2}(I,V^{\prime})}}{\beta(a)}. (4.26)

We also have

f~L2(I,V)=supv1L2(I,V)=1|f~(v1)|abL𝒰(b)L2(I,V)v1L2(I,V)abL1β(b)(fL2(I,V)2+u0L22)1/2,\begin{split}\|\tilde{f}\|_{L^{2}(I,V^{\prime})}=\sup_{\|v_{1}\|_{L^{2}(I,V)}=1}|\tilde{f}(v_{1})|&\leq\|a-b\|_{L_{\infty}}\|{\mathcal{U}}(b)\|_{L^{2}(I,V)}\|v_{1}\|_{L^{2}(I,V)}\\ &\leq\|a-b\|_{L_{\infty}}\frac{1}{\beta(b)}\Big{(}\|f\|_{L^{2}(I,V^{\prime})}^{2}+\|u_{0}\|_{L^{2}}^{2}\Big{)}^{1/2},\end{split}

where in the last estimate we used again Proposition 4.12. Inserting this into (4.26) we obtain the desired result. ∎

We are now in position to verify the assumptions (i)(iii) of Theorem 4.11 for the data-to-solution map a𝒰(a)a\mapsto{\mathcal{U}}(a) to the equation (4.22).

  1. (i)

    For the first condition, it has been shown that the weak solution to the linear parabolic PDEs (4.22) depends holomorphically on the data aOa\in O by the Ladyzhenskaya-Babuška-Brezzi theorem in Hilbert spaces over \mathbb{C}, see e.g. [37, Pages 26, 27].

  2. (ii)

    Let aOa\in O. Using the elementary estimate a+baba+b\leq ab with a,b2a,b\geq 2, we get

    2max(ρ(a)2,1)+ϑ22max(ρ(a)2,1)+max(ϑ2,2)2max(ρ(a)2,1)max(ϑ2,2)max(ϑ2,2)(ρ(a)1+1).\begin{split}\sqrt{2\max(\rho(a)^{-2},1)+\vartheta^{2}}&\leq\sqrt{2\max(\rho(a)^{-2},1)+\max(\vartheta^{2},2)}\\ &\leq\sqrt{2\max(\rho(a)^{-2},1)\max(\vartheta^{2},2)}\leq\max(\vartheta\sqrt{2},2)({\rho(a)}^{-1}+1)\,.\end{split}

    Hence, from (4.25) we can bound

    𝒰(a)XC0(ρ(a)1+1)min(ρ(a)aL2,ρ(a))=C0(1+ρ(a))ρ(a)2min(aL2,1)C0(1+aL)aL2min(ρ(a)4,1)C0(1+aLmin(ρ(a),1))4,\begin{split}\|{\mathcal{U}}(a)\|_{X}\leq\frac{C_{0}(\rho(a)^{-1}+1)}{\min\big{(}\rho(a)\|a\|_{L^{\infty}}^{-2},\rho(a)\big{)}}&=\frac{C_{0}(1+\rho(a))}{\rho(a)^{2}\min\big{(}\|a\|_{L^{\infty}}^{-2},1\big{)}}\\ &\leq\frac{C_{0}(1+\|a\|_{L^{\infty}})\|a\|_{L^{\infty}}^{2}}{\min\big{(}\rho(a)^{4},1\big{)}}\leq C_{0}\bigg{(}\frac{1+\|a\|_{L^{\infty}}}{\min\big{(}\rho(a),1\big{)}}\bigg{)}^{4}\,,\end{split} (4.27)

    where

    C0=max(ϑ2,2)(fL2(I,V)2+u0L22)1/2.C_{0}=\max(\vartheta\sqrt{2},2)\big{(}\|f\|_{L^{2}(I,V^{\prime})}^{2}+\|u_{0}\|_{L^{2}}^{2}\big{)}^{1/2}.
  3. (iii)

    The third assumption follows from Lemma 4.13 and the part (ii), i.e., for a,bOa,b\in O holds

    𝒰(a)𝒰(b)XC(1+aLmin(ρ(a),1))4(1+bLmin(ρ(b),1))4abL,\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{X}\leq C\bigg{(}\frac{1+\|a\|_{L^{\infty}}}{\min\big{(}\rho(a),1\big{)}}\bigg{)}^{4}\bigg{(}\frac{1+\|b\|_{L^{\infty}}}{\min\big{(}\rho(b),1\big{)}}\bigg{)}^{4}\|a-b\|_{L_{\infty}}, (4.28)

    for some C>0C>0 depending on ff, u0u_{0} and TT.

In conclusion, if (ψj)jL(D)(\psi_{j})_{j\in\mathbb{N}}\subset L^{\infty}({{D}}) such that with bj:=ψjLb_{j}:=\|\psi_{j}\|_{L^{\infty}} it holds 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}), then the solution

u(𝒚)=limN𝒰(exp(j=1Nyjψj))u({\boldsymbol{y}})=\lim_{N\to\infty}{\mathcal{U}}\left(\exp\left(\sum_{j=1}^{N}y_{j}\psi_{j}\right)\right) (4.29)

belonging to L2(U,X;γ)L^{2}(U,X;\gamma) is well-defined and (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic by Theorem 4.11.

We continue studying the holomorphy of the solution map to the equation (4.22) in function space of higher-regularity. Denote by H1(I,L2(D))H^{1}(I,L^{2}({D})) the space of all functions v(t,𝒙)L2(I,L2(D))v(t,{\boldsymbol{x}})\in L^{2}(I,L^{2}({D})) such that the norm

vH1(I,L2):=(vL(I,L2)2+tvL2(I,L2)2)1/2\|v\|_{H^{1}(I,L^{2})}:=\Big{(}\|v\|_{L(I,L^{2})}^{2}+\|\partial_{t}v\|_{L^{2}(I,L^{2})}^{2}\Big{)}^{1/2}

is finite. We put

Z:=L2(I,W)H1(I,L2(D)),W:={vV:ΔvL2(D)},Z:=L^{2}(I,W)\cap H^{1}(I,L^{2}({D})),\quad W:=\big{\{}v\in V:\ \Delta v\in L^{2}({D})\big{\}},

and

vZ:=(vH1(I,L2)2+vL2(I,W)2)1/2.\|v\|_{Z}:=\Big{(}\|v\|_{H^{1}(I,L^{2})}^{2}+\|v\|_{L^{2}(I,W)}^{2}\Big{)}^{1/2}.

In the following the constant CC and CC^{\prime} may change their values from line to line.

Lemma 4.14.

Assume that aW1(D)Oa\in W^{1}_{\infty}({D})\cap O and fL2(I,L2(D))f\in L^{2}(I,L^{2}({D})) and u0Vu_{0}\in V. Suppose further that 𝒰(a)X{\mathcal{U}}(a)\in X is the weak solution to the equation (4.22). Then 𝒰(a)L2(I,W)H1(I,L2(D)){\mathcal{U}}(a)\in L^{2}(I,W)\cap H^{1}(I,L^{2}({D})). Furthermore,

t𝒰(a)L2(I,L2)(1+aW1min(ρ(a),1))4(u0V+fL2(I,L2))1/2,\displaystyle\|\partial_{t}{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})}\leq\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min(\rho(a),1)}\bigg{)}^{4}\big{(}\|u_{0}\|_{V}+\|f\|_{L^{2}(I,L^{2})}\big{)}^{1/2},

and

Δ𝒰(a)L2(I,L2)\displaystyle\|\Delta{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})} C(1+aW1min(ρ(a),1))5(u0V2+fL2(I,L2)2)1/2,\displaystyle\leq C\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2},

where C>0C>0 independent of ff and u0u_{0}. Therefore,

𝒰(a)ZC(1+aW1min(ρ(a),1))5(u0V2+fL2(I,L2)2)1/2.\|{\mathcal{U}}(a)\|_{Z}\leq C\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}.
Proof.

The argument follows along the lines of, e.g., [53, Section 7.1.3] by separation of variables. Let (ωk)kV(\omega_{k})_{k\in{\mathbb{N}}}\subset V be an orthogonal basis which is orthonormal basis of L2(D)L^{2}({D}), [eigenbasis in polygon generally not smooth], see, e.g. [53, Page 353]. Let further, for mm\in\mathbb{N},

𝒰m(a)=k=1mdmk(t)ωkVm{\mathcal{U}}_{m}(a)=\sum_{k=1}^{m}d_{m}^{k}(t)\omega_{k}\in V_{m}

be a Galerkin approximation to 𝒰(a){\mathcal{U}}(a) on Vm:=span{ωk,k=1,,m}V_{m}:={\rm span}\{\omega_{k},\ k=1,\ldots,m\}.

Then we have

t𝒰m(a)=k=1mddtdmk(t)ωkVm.\partial_{t}{\mathcal{U}}_{m}(a)=\sum_{k=1}^{m}\frac{d}{dt}d_{m}^{k}(t)\omega_{k}\in V_{m}.

Multiplying both sides with t𝒰m(a)\partial_{t}{\mathcal{U}}_{m}(a) we get

Dt𝒰m(a)t𝒰m(a)¯d𝒙+Da𝒰m(a)t𝒰m(a)¯d𝒙=Dft𝒰m(a)¯d𝒙.\displaystyle\int_{{D}}\partial_{t}{\mathcal{U}}_{m}(a)\partial_{t}\overline{{\mathcal{U}}_{m}(a)}\,\mathrm{d}{\boldsymbol{x}}+\int_{{D}}a\nabla{\mathcal{U}}_{m}(a)\cdot\partial_{t}\overline{\nabla{\mathcal{U}}_{m}(a)}\,\,\mathrm{d}{\boldsymbol{x}}=\int_{{D}}f\partial_{t}\overline{{\mathcal{U}}_{m}(a)}\,\,\mathrm{d}{\boldsymbol{x}}.

The conjugate equation is given by

Dt𝒰m(a)t𝒰m(a)¯d𝒙\displaystyle\int_{{D}}\partial_{t}{\mathcal{U}}_{m}(a)\partial_{t}\overline{{\mathcal{U}}_{m}(a)}\,\mathrm{d}{\boldsymbol{x}} +Da¯𝒰m(a)¯t𝒰m(a)d𝒙=Df¯t𝒰m(a)d𝒙.\displaystyle+\int_{{D}}\overline{a}\,\overline{\nabla{\mathcal{U}}_{m}(a)}\cdot\partial_{t}\nabla{\mathcal{U}}_{m}(a)\,\,\mathrm{d}{\boldsymbol{x}}=\int_{{D}}\bar{f}\partial_{t}{{\mathcal{U}}_{m}(a)}\,\,\mathrm{d}{\boldsymbol{x}}.

Consequently we obtain

2t𝒰m(a)L22\displaystyle 2\|\partial_{t}{\mathcal{U}}_{m}(a)\|_{L^{2}}^{2} +ddtD(a)|𝒰m(a)|2d𝒙=Dft𝒰m(a)¯d𝒙+Df¯t𝒰m(a)d𝒙.\displaystyle+\frac{\,\mathrm{d}}{\,\mathrm{d}t}\int_{{D}}\Re(a)|\nabla{\mathcal{U}}_{m}(a)|^{2}\,\mathrm{d}{\boldsymbol{x}}=\int_{{D}}f\partial_{t}\overline{{\mathcal{U}}_{m}(a)}\,\,\mathrm{d}{\boldsymbol{x}}+\int_{{D}}\bar{f}\partial_{t}{{\mathcal{U}}_{m}(a)}\,\,\mathrm{d}{\boldsymbol{x}}.

Integrating both sides with respect to tt on II and using the Cauchy-Schwarz inequality we arrive at

2t𝒰m(a)L2(I,L2)2\displaystyle 2\|\partial_{t}{\mathcal{U}}_{m}(a)\|_{L^{2}(I,L^{2})}^{2} +D(a)|𝒰m(a)|t=T|2d𝒙\displaystyle+\int_{{D}}\Re(a)\big{|}\nabla{\mathcal{U}}_{m}(a)\big{|}_{t=T}\big{|}^{2}\,\mathrm{d}{\boldsymbol{x}}
D(a)|𝒰m(a)|t=0|2d𝒙+fL2(I,L2)2+t𝒰m(a)L2(I,L2)2,\displaystyle\leq\int_{{D}}\Re(a)\big{|}\nabla{\mathcal{U}}_{m}(a)\big{|}_{t=0}\big{|}^{2}\,\mathrm{d}{\boldsymbol{x}}+\|f\|_{L^{2}(I,L^{2})}^{2}+\|\partial_{t}{\mathcal{U}}_{m}(a)\|_{L^{2}(I,L^{2})}^{2},

which implies

t𝒰m(a)L2(I,L2)2\displaystyle\|\partial_{t}{\mathcal{U}}_{m}(a)\|_{L^{2}(I,L^{2})}^{2} D(a)|𝒰m(a)|t=0|2d𝒙+fL2(I,L2)2\displaystyle\leq\int_{{D}}\Re(a)\big{|}\nabla{\mathcal{U}}_{m}(a)\big{|}_{t=0}\big{|}^{2}\,\mathrm{d}{\boldsymbol{x}}+\|f\|_{L^{2}(I,L^{2})}^{2} (4.30)
aL𝒰m(a)|t=0L22+fL2(I,L2)2\displaystyle\leq\|a\|_{L^{\infty}}\big{\|}\nabla{\mathcal{U}}_{m}(a)\big{|}_{t=0}\big{\|}_{L^{2}}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}
aLu0V2+fL2(I,L2)2,\displaystyle\leq\|a\|_{L^{\infty}}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2},

where we used the bounds 𝒰m(a)|t=0L2u0V\|\nabla{\mathcal{U}}_{m}(a)\big{|}_{t=0}\|_{L^{2}}\leq\|u_{0}\|_{V}, see [53, Page 362].

Passing to limits we deduce that

t𝒰(a)L2(I,L2)\displaystyle\|\partial_{t}{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})} (aLu0V2+fL2(I,L2)2)1/2\displaystyle\leq\big{(}\|a\|_{L^{\infty}}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
(aL+1)1/2(u0V2+fL2(I,L2)2)1/2\displaystyle\leq(\|a\|_{L^{\infty}}+1)^{1/2}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
C(1+aW1min(ρ(a),1))4(u0V2+fL2(I,L2)2)1/2.\displaystyle\leq C\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{4}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}.

We also have from (4.25) and (4.27) that

𝒰(a)L2(I,L2)C𝒰(a)L2(I,V)\displaystyle\|{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})}\leq C\|{\mathcal{U}}(a)\|_{L^{2}(I,V)} Cβ(a)(fL2(I,V)2+u0L22)1/2\displaystyle\leq\frac{C}{\beta(a)}\big{(}\|f\|_{L^{2}(I,V^{\prime})}^{2}+\|u_{0}\|_{L^{2}}^{2}\big{)}^{1/2} (4.31)
Cβ(a)(u0V2+fL2(I,L2)2)1/2\displaystyle\leq\frac{C}{\beta(a)}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
C(1+aLmin(ρ(a),1))4(u0V2+fL2(I,L2)2)1/2\displaystyle\leq C\bigg{(}\frac{1+\|a\|_{L^{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{4}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
C(1+aW1min(ρ(a),1))4(u0V2+fL2(I,L2)2)1/2.\displaystyle\leq C\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{4}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}.

We now estimate Δ𝒰(a)L2(I,L2)\|\Delta{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})}. From the identity (valid in L2(I,L2(D))L^{2}(I,L^{2}({D})))

Δ𝒰(a)=1a[a𝒰(a)+ft𝒰(a)],\displaystyle-\Delta{\mathcal{U}}(a)=\frac{1}{a}\big{[}\nabla a\cdot\nabla{\mathcal{U}}(a)+f-\partial_{t}{\mathcal{U}}(a)\big{]},

and (4.30), (4.31) we obtain that

Δ𝒰(a)L2(I,L2)\displaystyle\|\Delta{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})} 1ρ(a)[aW1𝒰(a)L2(I,V)+fL2(I,L2)+t𝒰(a)L2(I,L2)]\displaystyle\leq\frac{1}{\rho(a)}\bigg{[}\|a\|_{W^{1}_{\infty}}\|{\mathcal{U}}(a)\|_{L^{2}(I,V)}+\|f\|_{L^{2}(I,L^{2})}+\|\partial_{t}{\mathcal{U}}(a)\|_{L^{2}(I,L^{2})}\bigg{]}
CaW1ρ(a)(1+aLmin(ρ(a),1))4(u0V2+fL2(I,L2)2)1/2\displaystyle\leq C\frac{\|a\|_{W^{1}_{\infty}}}{{\rho(a)}}\bigg{(}\frac{1+\|a\|_{L^{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{4}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
C(1+aW1min(ρ(a),1))5(u0V2+fL2(I,L2)2)1/2,\displaystyle\leq C\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2},

with C>0C>0 independent of ff and u0u_{0}. Combining this and (4.30), (4.31), the desired result follows. ∎

Lemma 4.15.

Assume fL2(I,L2(D))f\in L^{2}(I,L^{2}({D})) and u0Vu_{0}\in V. Let 𝒰(a){\mathcal{U}}(a) and 𝒰(b){\mathcal{U}}(b) be the solutions to (4.23) with a,bW1(D)Oa,b\in W^{1}_{\infty}({D})\cap O, respectively. Then we have

𝒰(a)𝒰(b)Z\displaystyle\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{Z} C(1+aW1min(ρ(a),1))5(1+bW1min(ρ(b),1))5abW1,\displaystyle\leq C^{\prime}\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\bigg{(}\frac{1+\|b\|_{W^{1}_{\infty}}}{\min({\rho(b)},1)}\bigg{)}^{5}\|a-b\|_{W^{1}_{\infty}},

with C>0C^{\prime}>0 depending on ff and u0u_{0}.

Proof.

Denote w:=𝒰(a)𝒰(b)w:={\mathcal{U}}(a)-{\mathcal{U}}(b). Then ww is the solution to the equation

{twdiv(aw)=(ab)𝒰(b)+(ab)Δ𝒰(b),w|D×I=0,w|t=0=0.\begin{cases}{\partial_{t}w-\operatorname{div}\big{(}a\nabla w\big{)}=\nabla(a-b)\cdot\nabla{\mathcal{U}}(b)+(a-b)\Delta{\mathcal{U}}(b),}\\ w|_{\partial{{D}}\times I}=0,\\ w|_{t=0}=0.\end{cases} (4.32)

Hence

Δw=1a[aw+(ab)𝒰(b)+(ab)Δ𝒰(b)tw]-\Delta w=\frac{1}{a}\bigg{[}\nabla a\cdot\nabla w+\nabla(a-b)\cdot\nabla{\mathcal{U}}(b)+(a-b)\Delta{\mathcal{U}}(b)-\partial_{t}w\bigg{]}

which leads to

ΔwL2(I,L2)\displaystyle\|\Delta w\|_{L^{2}(I,L^{2})} 1ρ(a)[aW1wL2(I,V)+twL2(I,L2)\displaystyle\leq\frac{1}{{\rho(a)}}\bigg{[}\|a\|_{W^{1}_{\infty}}\|w\|_{L^{2}(I,V)}+\|\partial_{t}w\|_{L^{2}(I,L^{2})}
+abW1(𝒰(b)L2(I,W)+𝒰(b)L2(I,V))].\displaystyle\ \ +\|a-b\|_{W^{1}_{\infty}}\big{(}\|{\mathcal{U}}(b)\|_{L^{2}(I,W)}+\|{\mathcal{U}}(b)\|_{L^{2}(I,V)}\big{)}\bigg{]}.

Lemma 4.14 gives that

twL2(I,L2)\displaystyle\|\partial_{t}w\|_{L^{2}(I,L^{2})} (1+aW1min(ρ(a),1))5((ab)𝒰(b)L2(I,L2)2+(ab)Δ𝒰(b)L2(I,L2)2)1/2\displaystyle\leq\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\big{(}\|\nabla(a-b)\cdot\nabla{\mathcal{U}}(b)\|_{L^{2}(I,L^{2})}^{2}+\|(a-b)\Delta{\mathcal{U}}(b)\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2}
(1+aW1min(ρ(a),1))5abW1(𝒰(b)L2(I,W)2+𝒰(b)L2(I,V)2)1/2,\displaystyle\leq\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\|a-b\|_{W^{1}_{\infty}}\big{(}\|{\mathcal{U}}(b)\|_{L^{2}(I,W)}^{2}+\|{\mathcal{U}}(b)\|_{L^{2}(I,V)}^{2}\big{)}^{1/2},

and

𝒰(b)L2(I,W)+𝒰(b)L2(I,V)\displaystyle\|{\mathcal{U}}(b)\|_{L^{2}(I,W)}+\|{\mathcal{U}}(b)\|_{L^{2}(I,V)} C(1+bW1min(ρ(b),1))5(u0V2+fL2(I,L2)2)1/2,\displaystyle\leq C\bigg{(}\frac{1+\|b\|_{{W^{1}_{\infty}}}}{\min({\rho(b)},1)}\bigg{)}^{5}\big{(}\|u_{0}\|_{V}^{2}+\|f\|_{L^{2}(I,L^{2})}^{2}\big{)}^{1/2},

which implies

abW1(𝒰(b)L2(I,W)+𝒰(b)L2(I,V))+twL2(I,L2)\displaystyle\|a-b\|_{W^{1}_{\infty}}\big{(}\|{\mathcal{U}}(b)\|_{L^{2}(I,W)}+\|{\mathcal{U}}(b)\|_{L^{2}(I,V)}\big{)}+\|\partial_{t}w\|_{L^{2}(I,L^{2})}
C(1+aW1min(ρ(a),1))5(1+bW1min(ρ(b),1))5abW1\displaystyle\leq C^{\prime}\bigg{(}\frac{1+\|a\|_{W^{1}_{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{5}\bigg{(}\frac{1+\|b\|_{{W^{1}_{\infty}}}}{\min({\rho(b)},1)}\bigg{)}^{5}\|a-b\|_{W^{1}_{\infty}}

We also have

wL2(I,V)1β(a)β(b)abLC(1+aLmin(ρ(a),1))4(1+bLmin(ρ(b),1))4abW1,\|w\|_{L^{2}(I,V)}\leq\frac{1}{\beta(a)\beta(b)}\|a-b\|_{L_{\infty}}\leq C^{\prime}\bigg{(}\frac{1+\|a\|_{L^{\infty}}}{\min({\rho(a)},1)}\bigg{)}^{4}\bigg{(}\frac{1+\|b\|_{L^{\infty}}}{\min({\rho(b)},1)}\bigg{)}^{4}\|a-b\|_{W^{1}_{\infty}},

see (4.28). Hence

ΔwL2(I,L2)\displaystyle\|\Delta w\|_{L^{2}(I,L^{2})} C(1+aW1min(ρ(a),1))5(1+bW1min(ρ(b),1))5abW1.\displaystyle\leq C^{\prime}\bigg{(}\frac{1+\|a\|_{{W^{1}_{\infty}}}}{\min({\rho(a)},1)}\bigg{)}^{5}\bigg{(}\frac{1+\|b\|_{{W^{1}_{\infty}}}}{\min({\rho(b)},1)}\bigg{)}^{5}\|a-b\|_{{W^{1}_{\infty}}}. (4.33)

Since the terms twL2(I,L2)\|\partial_{t}w\|_{L^{2}(I,L^{2})} and wL2(I,L2)\|w\|_{L^{2}(I,L^{2})} are also bounded by the right side of (4.33), we arrive at

wZ\displaystyle\|w\|_{Z} =(ΔwL2(I,L2)2+twL2(I,L2)2+wL2(I,L2)2)1/2\displaystyle=\Big{(}\|\Delta w\|_{L^{2}(I,L^{2})}^{2}+\|\partial_{t}w\|_{L^{2}(I,L^{2})}^{2}+\|w\|_{L^{2}(I,L^{2})}^{2}\Big{)}^{1/2}
C(1+aW1min(ρ(a),1))5(1+bW1min(ρ(b),1))5abW1\displaystyle\leq C^{\prime}\bigg{(}\frac{1+\|a\|_{{W^{1}_{\infty}}}}{\min(\rho(a),1)}\bigg{)}^{5}\bigg{(}\frac{1+\|b\|_{{W^{1}_{\infty}}}}{\min(\rho(b),1)}\bigg{)}^{5}\|a-b\|_{{W^{1}_{\infty}}}

which is the claim. ∎

From Lemma 4.15, by the same argument as in the proof of [71, Proposition 4.5] we can verify that the solution map a𝒰(a)a\mapsto{\mathcal{U}}(a) from W1(D)O{W^{1}_{\infty}}({D})\cap O to ZZ is holomorphic. If we assume further that (ψj)jW1(D)(\psi_{j})_{j\in\mathbb{N}}\subseteq{W^{1}_{\infty}}({D}) and with bj:=ψjW1b_{j}:=\|\psi_{j}\|_{W^{1}_{\infty}}, it holds 𝒃1(){\boldsymbol{b}}\in\ell^{1}(\mathbb{N}) and all the conditions in Theorem 4.11 are satisfied. Therefore, u(𝒚)u({\boldsymbol{y}}) given by the formula (4.29) is (𝒃,ξ,δ,Z)({\boldsymbol{b}},\xi,\delta,Z)-holomorphic with appropriate ξ\xi and δ\delta.

Remark 4.16.

For s>1s>1, let

Zs:=k=0sHk(I,H2s2k(D))Z^{s}:=\bigcap_{k=0}^{s}H^{k}(I,H^{2s-2k}({D}))

with the norm

vZs=(k=0sdkvdtkL2(I,H2s2k)2)1/2.\|v\|_{Z^{s}}=\Bigg{(}\sum_{k=0}^{s}\bigg{\|}\frac{\,\mathrm{d}^{k}v}{\,\mathrm{d}t^{k}}\bigg{\|}_{L^{2}(I,H^{2s-2k})}^{2}\Bigg{)}^{1/2}.

Assume that aW2s1(D)Oa\in W^{2s-1}_{\infty}({D})\cap O. At present we do not know whether the solution map a𝒰(a)a\mapsto{\mathcal{U}}(a) from W2s1(D)OW^{2s-1}_{\infty}({D})\cap O to ZsZ^{s} is holomorphic. To obtain the holomorphy of the solution map, we need a result similar to that in Lemma 4.15. In order for this to hold, higher-order regularity and compatibility of the data for equation (4.32) is required, i.e,

g0=0V,g1=h(0)Lg0V,,gs=ds1hdts1(0)Lgs1V,g_{0}=0\in V,\quad g_{1}=h(0)-Lg_{0}\in V,\ldots,g_{s}=\frac{\,\mathrm{d}^{s-1}h}{\,\mathrm{d}t^{s-1}}(0)-Lg_{s-1}\in V,

where

h=(ab)𝒰(b)+(ab)Δ𝒰(b),L=tdiv(a).h=\nabla(a-b)\cdot\nabla{\mathcal{U}}(b)+(a-b)\Delta{\mathcal{U}}(b),\quad L=\partial_{t}\cdot-\operatorname{div}\big{(}a\nabla\cdot\big{)}.

See e.g. [110, Theorem 27.2]. It is known that without such compatibility, the solution will develop spatial singularities at the corners and edges of D{D}, and temporal singularities as t0t\downarrow 0; see e.g. [81].

In general the compatibility condition does not hold when we only assume that

u0H2s1(D)VanddkfdtkL2(I,H2s2k2(D))u_{0}\in H^{2s-1}({D})\cap V\ \ {\rm and}\ \ \frac{\,\mathrm{d}^{k}f}{\,\mathrm{d}t^{k}}\in L^{2}(I,H^{2s-2k-2}({D}))

for k=0,,s1k=0,\ldots,s-1.

4.3.3 Linear elastostatics with log-Gaussian modulus of elasticity

We illustrate the foregoing abstract setting of Section 4.1 for another class of boundary value problems. In computational mechanics, one is interested in the numerical approximation of deformations of elastic bodies. We refer to e.g. [108] for an accessible exposition of the mathematical foundations and assumptions. In linearized elastostatics one is concerned with small (in a suitable sense, see [108] for details) deformations.

We consider an elastic body occupying the domain Dd{D}\subset\mathbb{R}^{d}, d=2,3d=2,3 (the physically relevant case naturally is d=3d=3, we include d=2d=2 to cover the so-called model of “plane-strain” which is widely used in engineering, and has governing equations with the same mathematical structure). In the linear theory, small deformations of the elastic body occupying D{D}, subject to, e.g., body forces 𝒇:Dd{\boldsymbol{f}}:{D}\to\mathbb{R}^{d} such as gravity are modeled in terms of the displacement field 𝒖:Dd{\boldsymbol{u}}:{D}\to\mathbb{R}^{d}, describing the displacement of a material point 𝒙D{\boldsymbol{x}}\in{D} (see [108] for a discussion of axiomatics related to this mathematical concept). Importantly, unlike the scale model problem considered up to this point, modeling now involves vector fields of data (e.g., 𝒇{\boldsymbol{f}}) and solution (i.e., 𝒖{\boldsymbol{u}}).

Governing equations for the mathematical model of linearly elastic deformation, subject to homogeneous Dirichlet boundary conditions on D\partial{D}, read: to find 𝒖:Dd{\boldsymbol{u}}:{D}\to\mathbb{R}^{d} such that

div𝝈[𝒖]+𝒇=0inD,𝒖=0onD.\begin{array}[]{rcl}{\rm div}{\boldsymbol{\sigma}}[{\boldsymbol{u}}]+{\boldsymbol{f}}&=&0\quad\mbox{in}\;\;{D}\;,\\ {\boldsymbol{u}}&=&0\quad\mbox{on}\;\;\partial{D}\;.\end{array} (4.34)

Here 𝝈:Dd×dsym{\boldsymbol{\sigma}}:{D}\to\mathbb{R}^{d\times d}_{{\rm sym}} is a symmetric matrix function, the so-called stress tensor. It depends on the displacement field uu via the so-called (linearized) strain tensor ϵ[𝒖]:Dd×dsym{\boldsymbol{\epsilon}}[{\boldsymbol{u}}]:{{D}}\to\mathbb{R}^{d\times d}_{{\rm sym}}, which is given by

ϵ[𝒖]:=12(grad𝒖+(grad𝒖)),(ϵ[𝒖])ij:=12(jui+iuj),i,j=1,,d.{\boldsymbol{\epsilon}}[{\boldsymbol{u}}]:=\frac{1}{2}\left({\rm grad}{\boldsymbol{u}}+({\rm grad}{\boldsymbol{u}})^{\top}\right)\;,\;\;({\boldsymbol{\epsilon}}[{\boldsymbol{u}}])_{ij}:=\frac{1}{2}(\partial_{j}u_{i}+\partial_{i}u_{j})\;,i,j=1,...,d\;. (4.35)

In the linearized theory, the tensors 𝝈{\boldsymbol{\sigma}} and ϵ{\boldsymbol{\epsilon}} in (4.34), (4.35) are related by the linear constitutive stress-strain relation (“Hooke’s law”)

𝝈=𝙰ϵ.{\boldsymbol{\sigma}}={\tt A}{\boldsymbol{\epsilon}}\;. (4.36)

In (4.36), 𝙰{\tt A} is a fourth order tensor field, i.e.

𝙰={𝙰ijkl:i,j,k,l=1,,d},{\tt A}=\{{\tt A}_{ijkl}:i,j,k,l=1,...,d\},

with certain symmetries that must hold among its d4d^{4} components independent of the particular material constituting the elastic body (see, e.g., [108] for details). Thus, (4.36) reads in components as σij=𝙰ijklϵkl\sigma_{ij}={\tt A}_{ijkl}\epsilon_{kl} with summation over repeated indices implied. Let us now fix d=3d=3. Symmetry implies that ϵ\epsilon and σ\sigma are characterized by 66 components. If, in addition, the material constituting the elastic body is isotropic, the tensor 𝙰{\tt A} can in fact be characterized by only two independent coefficient functions. We adopt here the Poisson ratio, denoted ν\nu, and the modulus of elasticity 𝙴{\tt E}. With these two parameters, the stress-strain law (4.36) can be expressed in the component form

(σ11σ22σ33σ12σ13σ23)=𝙴(1+ν)(12ν)(1ννν000ν1νν000νν1ν00000012ν00000012ν00000012ν)(ϵ11ϵ22ϵ33ϵ12ϵ13ϵ23).\left(\begin{array}[]{c}\sigma_{11}\\ \sigma_{22}\\ \sigma_{33}\\ \sigma_{12}\\ \sigma_{13}\\ \sigma_{23}\end{array}\right)=\frac{{\tt E}}{(1+\nu)(1-2\nu)}\left(\begin{array}[]{cccccc}1-\nu&\nu&\nu&0&0&0\\ \nu&1-\nu&\nu&0&0&0\\ \nu&\nu&1-\nu&0&0&0\\ 0&0&0&1-2\nu&0&0\\ 0&0&0&0&1-2\nu&0\\ 0&0&0&0&0&1-2\nu\end{array}\right)\left(\begin{array}[]{c}\epsilon_{11}\\ \epsilon_{22}\\ \epsilon_{33}\\ \epsilon_{12}\\ \epsilon_{13}\\ \epsilon_{23}\end{array}\right)\;. (4.37)

We see from (4.37) that for isotropic elastic materials, the tensor 𝙰{\tt A} is proportional to the modulus 𝙴>0{\tt E}>0, with the Poisson ratio ν[0,1/2)\nu\in[0,1/2). We remark that for common materials, ν1/2\nu\uparrow 1/2 arises in the so-called incompressible limit. In that case, (4.34) can be described by the Stokes equations.

With the constitutive law (4.36), we may cast the governing equation (4.34) into the so-called “primal”, or “displacement-formulation”: find 𝒖:Dd{\boldsymbol{u}}:{D}\to\mathbb{R}^{d} such that

div(𝙰ϵ[𝒖])=𝒇inD,𝒖|D=0.-{\rm div}({\tt A}{\boldsymbol{\epsilon}}[{\boldsymbol{u}}])=\boldsymbol{f}\quad\mbox{in}\;\;{D}\;,\qquad{\boldsymbol{u}}|_{\partial{D}}=0\;. (4.38)

This form is structurally identical to the scalar diffusion problem (3.1).

Accordingly, we fix ν[0,1/2)\nu\in[0,1/2) and model uncertainty in the elastic modulus 𝙴>0{\tt E}>0 in (4.37) by a log-Gaussian random field

𝙴(𝒚)(𝒙):=exp(b(𝒚))(𝒙),𝒙D,𝒚U.{\tt E}({\boldsymbol{y}})({\boldsymbol{x}}):=\exp(b({\boldsymbol{y}}))({\boldsymbol{x}})\;,\quad{\boldsymbol{x}}\in{D}\;,\;\;{\boldsymbol{y}}\in U\;. (4.39)

Here, b(𝒚)b({\boldsymbol{y}}) is a Gaussian series representation of the GRF b(Y(ω))b(Y(\omega)) as discussed in Section 2.5. The log-Gaussian ansatz 𝙴=exp(b){\tt E}=\exp(b) ensures

Emin(𝒚):=essinf𝒙D𝙴(𝒚)(𝒙)>0γ-a.e.𝒚U,E_{\min}({\boldsymbol{y}}):={\rm ess}\inf_{{\boldsymbol{x}}\in{D}}{\tt E}({\boldsymbol{y}})({\boldsymbol{x}})>0\qquad\mbox{$\gamma$-a.e.}\ {\boldsymbol{y}}\in U\;,

i.e., the γ\gamma-almost sure positivity of (realizations of) the elastic modulus 𝙴{\tt E}. Denoting the 3×33\times 3 matrix relating the stress and strain components in (4.37) also by 𝙰{\tt A} (this slight abuse of notation should, however, not cause confusion in the following), we record that for 0ν<1/20\leq\nu<1/2, the matrix 𝙰{\tt A} is invertible:

𝙰1=1𝙴(1νν000ν1ν000νν10000001+ν0000001+ν0000001+ν).{\tt A}^{-1}=\frac{1}{{\tt E}}\left(\begin{array}[]{cccccc}1&-\nu&-\nu&0&0&0\\ -\nu&1&-\nu&0&0&0\\ -\nu&-\nu&1&0&0&0\\ 0&0&0&1+\nu&0&0\\ 0&0&0&0&1+\nu&0\\ 0&0&0&0&0&1+\nu\end{array}\right)\;. (4.40)

It readily follows from this explicit expression that due to

𝙴1(𝒚)(𝒙)=exp(b(𝒚)(𝒙)),{\tt E}^{-1}({\boldsymbol{y}})({\boldsymbol{x}})=\exp(-b({\boldsymbol{y}})({\boldsymbol{x}})),

by the Gerschgorin theorem invertibility holds for γ\gamma-a.e. 𝒚U{\boldsymbol{y}}\in U. Also, the components of 𝙰1{\tt A}^{-1} are GRFs (which are, however, fully correlated for deterministic ν\nu).

Occasionally, instead of the constants 𝙴{\tt E} and ν\nu, one finds the (equivalent) so-called Lamé-constants λ\lambda, μ\mu. They are related to 𝙴{\tt E} and ν\nu by

λ=𝙴ν(1+ν)(12ν),μ=𝙴2(1+ν).\lambda=\frac{{\tt E}\nu}{(1+\nu)(1-2\nu)}\;,\quad\mu=\frac{{\tt E}}{2(1+\nu)}\;. (4.41)

For GRF models (4.39) of 𝙴{\tt E}, (4.41) shows that for each fixed ν(0,1/2)\nu\in(0,1/2), also the Lamé-constants are GRFs which are fully correlated. This implies, in particular, that “large” realizations of the GRF (4.39) do not cause so-called “volume locking” in the equilibrium equation (4.34): this effect is related to the elastic material described by the constitutive equation (4.36) being nearly incompressible. Incompressibility here arises as either ν1/2\nu\uparrow 1/2 at fixed 𝙴{\tt E} or, equivalently, as λ\lambda\to\infty at fixed μ\mu.

Parametric weak solutions of (4.38) with (4.39) are within the scope of the abstract theory developed up to this point. To see this, we provide a variational formulation of (4.38). Assuming for convenience homogeneous Dirichlet boundary conditions, we multiply (4.38) by a test displacement field 𝒗X:=Vd{\boldsymbol{v}}\in X:=V^{d} with V:=H01(D)V:=H_{0}^{1}({{D}}), and integrate by parts, to obtain the weak formulation: find 𝒖X{\boldsymbol{u}}\in X such that, for all 𝒗X{\boldsymbol{v}}\in X holds (in the matrix-vector notation (4.37))

Dϵ[𝒗]𝙰ϵ[𝒖]d𝒙=2μ(ϵ[𝒖],ϵ[𝒗])+λ(div𝒖,div𝒗)=(𝒇,𝒗).\int_{{D}}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\cdot{\tt A}{\boldsymbol{\epsilon}}[{\boldsymbol{u}}]\,\mathrm{d}{\boldsymbol{x}}=2\mu({\boldsymbol{\epsilon}}[{\boldsymbol{u}}],{\boldsymbol{\epsilon}}[{\boldsymbol{v}}])+\lambda(\operatorname{div}{\boldsymbol{u}},\operatorname{div}{\boldsymbol{v}})=({\boldsymbol{f}},{\boldsymbol{v}})\;. (4.42)

The variational form (4.42) suggests that, as λ\lambda\to\infty for fixed μ\mu, the “volume-preservation” constraint div𝒖L2=0\|\operatorname{div}{\boldsymbol{u}}\|_{L^{2}}=0 is imposed for 𝒗=𝒖{\boldsymbol{v}}={\boldsymbol{u}} in (4.42).

Unique solvability of (4.42) follows upon verifying coercivity of the corresponding bilinear form on the left-hand side of (4.42). It follows from (4.37) and (4.40) that

𝒗H1(D)d:𝙴cmin(ν)ϵ[𝒗]2L2Dϵ[𝒗]𝙰ϵ[𝒗]d𝒙𝙴cmax(ν)ϵ[𝒗]2L2.\forall{\boldsymbol{v}}\in H^{1}({D})^{d}:\quad{\tt E}c_{\min}(\nu)\|{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\|^{2}_{L^{2}}\leq\int_{{D}}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\cdot{\tt A}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\,\mathrm{d}{\boldsymbol{x}}\leq{\tt E}c_{\max}(\nu)\|{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\|^{2}_{L^{2}}\;.

Here, the constants cmin,cmaxc_{\min},c_{\max} are positive and bounded for 0<ν<1/20<\nu<1/2 and independent of 𝙴{\tt E}.

For the log-Gaussian model (4.39) of the elastic modulus 𝙴{\tt E}, the relations (4.41) show in particular, that the volume-locking effect arises as in the deterministic setting only if ν1/2\nu\simeq 1/2, independent of the realization of 𝙴(𝐲){\tt E}({\boldsymbol{y}}). Let us consider well-posedness of the variational formulation (4.42), for log-Gaussian, parametric elastic modulus 𝙴(𝒚){\tt E}({\boldsymbol{y}}) as in (4.39). To this end, with 𝙰1{\tt A}_{1} denoting the matrix 𝙰{\tt A} in (4.37) with 𝙴=1{\tt E}=1, we introduce in (4.42) the parametric bilinear forms

b(𝒖,𝒗;𝒚):=𝙴(𝒚)Dϵ[𝒗]𝙰1ϵ[𝒖]d𝒙=𝙴(𝒚)1+ν((ϵ[𝒖],ϵ[𝒗])+ν12ν(div𝒖,div𝒗)).b({\boldsymbol{u}},{\boldsymbol{v}};{\boldsymbol{y}}):={\tt E}({\boldsymbol{y}})\int_{{D}}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\cdot{\tt A}_{1}{\boldsymbol{\epsilon}}[{\boldsymbol{u}}]\,\mathrm{d}{\boldsymbol{x}}=\frac{{\tt E}({\boldsymbol{y}})}{1+\nu}\left(({\boldsymbol{\epsilon}}[{\boldsymbol{u}}],{\boldsymbol{\epsilon}}[{\boldsymbol{v}}])+\frac{\nu}{1-2\nu}(\operatorname{div}{\boldsymbol{u}},\operatorname{div}{\boldsymbol{v}})\right)\;.

Let us verify continuity and coercivity of the parametric bilinear forms

{b(,;𝒚):X×X:𝒚U},\{b(\cdot,\cdot;{\boldsymbol{y}}):X\times X\to\mathbb{R}:{\boldsymbol{y}}\in U\}, (4.43)

where we recall that U:=U:={\mathbb{R}}^{\infty}. With 𝙰1{\tt A}_{1} as defined above, we write for arbitrary 𝒗X=H10(D)d{\boldsymbol{v}}\in X=H^{1}_{0}({D})^{d}, d=2,3d=2,3, and for all 𝒚U0U{\boldsymbol{y}}\in U_{0}\subset U where the set U0U_{0} is as in (3.21),

b(𝒗,𝒗;𝒚)=Dϵ[𝒗](𝙰ϵ[𝒗])d𝒙=DE(𝒚)(ϵ[𝒗](𝙰1ϵ[𝒗]))d𝒙c(ν)DE(𝒚)ϵ[𝒗]22d𝒙c(ν)exp(b(𝒚)L)Dϵ[𝒗]22d𝒙c(ν)2amin(𝒚)|𝒗|H12CPc(ν)2amin(𝒚)𝒗H12.\begin{array}[]{rcl}b({\boldsymbol{v}},{\boldsymbol{v}};{\boldsymbol{y}})&=&\displaystyle\int_{{D}}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\cdot({\tt A}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}])\,\mathrm{d}{\boldsymbol{x}}=\int_{{D}}E({\boldsymbol{y}})\left({\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\cdot({\tt A}_{1}{\boldsymbol{\epsilon}}[{\boldsymbol{v}}])\right)\,\mathrm{d}{\boldsymbol{x}}\\ &\geq&\displaystyle c(\nu)\int_{{D}}E({\boldsymbol{y}})\|{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\|_{2}^{2}\,\mathrm{d}{\boldsymbol{x}}\\ &\geq&\displaystyle c(\nu)\exp(-\|b({\boldsymbol{y}})\|_{L^{\infty}})\int_{{D}}\|{\boldsymbol{\epsilon}}[{\boldsymbol{v}}]\|_{2}^{2}\,\mathrm{d}{\boldsymbol{x}}\\ &\geq&\displaystyle\frac{c(\nu)}{2}a_{\min}({\boldsymbol{y}})|{\boldsymbol{v}}|_{H^{1}}^{2}\\ &\geq&\displaystyle C_{P}\frac{c(\nu)}{2}a_{\min}({\boldsymbol{y}})\|{\boldsymbol{v}}\|_{H^{1}}^{2}\;.\end{array}

Here, in the last two steps we employed the first Korn’s inequality, and the Poincaré inequality, respectively. The lower bound E(𝒚)exp(b(𝒚)L)E({\boldsymbol{y}})\geq\exp(-\|b({\boldsymbol{y}})\|_{L^{\infty}}) is identical to (3.20) in the scalar diffusion problem.

In a similar fashion, continuity of the bilinear forms (4.43) may be established: there exists a constant c(ν)>0c^{\prime}(\nu)>0 such that

𝒖,𝒗X,𝒚U0:|b(𝒖,𝒗;𝒚)|c(ν)exp(b(𝒚)L)𝒖H1𝒗H1.\forall{\boldsymbol{u}},{\boldsymbol{v}}\in X,\;\forall{\boldsymbol{y}}\in U_{0}:\quad|b({\boldsymbol{u}},{\boldsymbol{v}};{\boldsymbol{y}})|\leq c^{\prime}(\nu)\exp(\|b({\boldsymbol{y}})\|_{L^{\infty}})\|{\boldsymbol{u}}\|_{H^{1}}\|{\boldsymbol{v}}\|_{H^{1}}.

With continuity and coercivity of the parametric forms (4.43) verified for 𝒚U0{\boldsymbol{y}}\in U_{0}, the Lax-Milgram lemma ensures for given 𝒇L2(D)d{\boldsymbol{f}}\in L^{2}({D})^{d} the existence of the parametric solution family

{𝒖(𝒚)X:b(𝒖,𝒗;𝒚)=(𝒇,𝒗)𝒗X,𝒚U0}.\{{\boldsymbol{u}}({\boldsymbol{y}})\in X:b({\boldsymbol{u}},{\boldsymbol{v}};{\boldsymbol{y}})=({\boldsymbol{f}},{\boldsymbol{v}})\ \forall{\boldsymbol{v}}\in X,{\boldsymbol{y}}\in U_{0}\}\;. (4.44)

Similar to the scalar case discussed in Proposition 3.7, the following result on almost everywhere existence and measurability holds.

Proposition 4.17.

Under Assumption 3.6, γ(U0)=1\gamma(U_{0})=1. For all kk\in{\mathbb{N}} there holds, with 𝔼()\mathbb{E}(\cdot) denoting expectation with respect to γ\gamma,

𝔼(exp(kb()L))<.\mathbb{E}\left(\exp(k\|b(\cdot)\|_{L^{\infty}})\right)<\infty\;.

The parametric solution family (4.44) of the parametric elliptic boundary value problem (4.42) with log-Gaussian modulus E(𝐲)E({\boldsymbol{y}}) as in (4.39) is in Lk(U,V;γ)L^{k}(U,V;\gamma) for every finite kk\in{\mathbb{N}}.

For the parametric solution family (4.44), analytic continuations into complex parameter domains, and parametric regularity results may be developed in analogy to the development in Sections 3.7 and 3.8. The key result for bootstrapping to higher order regularity is, in the case of smooth boundaries D\partial{D}, classical elliptic regularity for linear, Agmon-Douglis-Nirenberg elliptic systems which comprise (4.38). In the polygonal (for d=2d=2) or polyhedral (d=3d=3) case, weighted regularity shifts in Kondrat’ev type spaces are available in [62, Theorem 5.2] (for d=2d=2) and in [103] (for both, d=2,3d=2,3).

4.3.4 Maxwell equations with log-Gaussian permittivity

Similar models are available for time-harmonic, electromagnetic waves in dielectric media with uncertain conductivity. We refer to [76], where log-Gaussian models are employed. There, also the parametric regularity analysis of the parametric electric and magnetic fields is discussed, albeit by real-variable methods. The setting in [76] is, however, so that the presently developed, complex variable methods can be brought to bear on it. We refrain from developing the details.

4.3.5 Linear parametric elliptic systems and transmission problems

In Section 3.8.1, Theorem 3.29 we obtained parameter-explicit elliptic regularity shifts for a scalar, linear second order parametric elliptic divergence-form PDE in polygonal domain D2{D}\subset\mathbb{R}^{2}. A key feature of these estimates in the subsequent analysis of sparsity of gpc expansions was the polynomial dependence on the parameter in the bounds on parametric solutions in corner-weighted Sobolev spaces of Kondrat’ev type. Such a-priori bounds are not limited to the particular setting considered in Section 3.8.1, but hold for rather general, linear elliptic PDEs in smooth domains Dd{D}\subset\mathbb{R}^{d} of space dimension d2d\geq 2, with parametric differential and boundary operators of general integer order. In particular, for example, for linear, anisotropic elastostatics in 3\mathbb{R}^{3}, for parametric fourth order PDEs in 2\mathbb{R}^{2} which arise in dimensionally reduced models of elastic continua (plates, shells, etc.). We refer to [80] for statements of results and proofs.

In the results in Section 3.8.1, we admitted inhomogeneous coefficients which are regular in all of D{D}. In many applications, transmission problems with parametric, inhomogeneous coefficients with are piecewise regular on a given, fixed (i.e. non-parametric) partition of D{D} is of interest. Also in these cases, corresponding a-priori estimates of parametric solution families with norm bounds which are polynomial with respect to the parameters hold. We refer to [95] for such results, in smooth domains D{D}, with smooth interfaces.

5 Parametric posterior analyticity and sparsity in BIPs

We have investigated the parametric analyticity of the forward solution maps of linear PDEs with uncertain parametric inputs which typically arise from GRF models for these inputs. We have also provided an analysis of sparsity in the Wiener-Hermite PC expansion of the corresponding parametric solution families.

We now explore the notion of parametric holomorphy in the context of BIPs for linear PDEs. For these PDEs we adopt the Bayesian setting as outlined, e.g., in [48] and the references there. This Bayesian setting is briefly recapitulated in Section 5.1. With a suitable version of Bayes’ theorem, the main result is a (short) proof of parametric (𝒃,ξ,δ,)({\boldsymbol{b}},\xi,\delta,\mathbb{C})-holomorphy of the Bayesian posterior density for unbounded parameter ranges. This implies sparsity of the coefficients in Wiener-Hermite PC expansions of the Bayesian posterior density, which can be leveraged to obtain higher-order approximation rates that are free from the curse of dimensionality for various deterministic approximation methods of the Bayesian expectations, for several classes of function space priors modelled by product measures on the parameter sequences 𝒚{\boldsymbol{y}}. In particular, the construction of Gaussian priors described in Section 2.2 is applicable. Concerning related previous works, we remark the following. In [98] holomorphy for a bounded parameter domain (in connection with uniform prior measure) has been addressed by complex variable arguments in the same fashion. In [96], MC and QMC integration has been analyzed by real-variable arguments for such Gaussian priors. In [66], corresponding results have been obtained also for so-called Besov priors, again by real-variable arguments for the parametric posterior. Since the presently developed, quantified parametric holomorphy results are independent of the particular measure placed upon the unbounded parameter domain {\mathbb{R}}^{\infty}. The sparsity and approximation rate bounds for the parametric deterministic posterior densities will imply approximate rate bounds also for prior constructions beyond the Gaussian ones.

5.1 Formulation and well-posedness

With E{E} and XX denoting separable Banach and Hilbert spaces over \mathbb{C}, respectively, we consider a forward solution map 𝒰:EX{\mathcal{U}}:{E}\to X and an observation map 𝓞:Xm{\boldsymbol{{\mathcal{O}}}}:X\to{\mathbb{R}}^{m}. In the context of the previous sections, 𝒰{\mathcal{U}} could denote again the map which associates with a diffusion coefficient aE:=L(D;)a\in{E}:=L^{\infty}({{D}};\mathbb{C}) the solution 𝒰(a)X:=H01(D;){\mathcal{U}}(a)\in X:=H_{0}^{1}({{D}};\mathbb{C}) of the equation (5.7) below. We assume the map 𝒰{\mathcal{U}} to be Borel measurable.

The inverse problem consists in determining the (expected value of an) uncertain input datum aEa\in{E} from noisy observation data 𝖉m{\boldsymbol{\mathfrak{d}}}\in{\mathbb{R}}^{m}. Here, the observation noise 𝜼m{\boldsymbol{\eta}}\in{\mathbb{R}}^{m} is assumed additive centered Gaussian, i.e., the observation data 𝖉{\boldsymbol{\mathfrak{d}}} for input aa is

𝖉=𝓞𝒰(a)+𝜼,{\boldsymbol{\mathfrak{d}}}={\boldsymbol{{\mathcal{O}}}}\circ{\mathcal{U}}(a)+{\boldsymbol{\eta}}\;,

where 𝜼𝒩(0,𝚪){\boldsymbol{\eta}}\sim{\mathcal{N}}(0,{\boldsymbol{\Gamma}}). We assume the observation noise covariance 𝚪m×m{\boldsymbol{\Gamma}}\in\mathbb{R}^{m\times m} is symmetric positive definite.

In the so-called Bayesian setting of the inverse problem, one assumes that the uncertain input aa is modelled as RV which is distributed according to a prior measure π0\pi_{0} on E{E}. Then, under suitable conditions, which are made precise in Theorem 5.2 below, the posterior distribution π(|𝖉)\pi(\cdot|{\boldsymbol{\mathfrak{d}}}) on the conditioned RV 𝒰|𝖉{\mathcal{U}}|{\boldsymbol{\mathfrak{d}}} is absolutely continuous w.r.t. the prior measure π0\pi_{0} on E{E} and there holds Bayes’ theorem in the form

dπ(|𝖉)dπ0(a)=1ZΘ(a).\frac{\,\mathrm{d}\pi(\cdot|{\boldsymbol{\mathfrak{d}}})}{\,\mathrm{d}\pi_{0}}(a)=\frac{1}{Z}\Theta(a). (5.1)

In (5.1), the posterior density Θ\Theta and the normalization constant ZZ are given by

Θ(a)=exp(Φ(𝖉;a)),Φ(𝖉;a)=12𝚪1/2(𝖉𝓞(𝒰(a)))22,Z=𝔼π0[Θ()].\Theta(a)=\exp(-\Phi({\boldsymbol{\mathfrak{d}}};a)),\qquad\Phi({\boldsymbol{\mathfrak{d}}};a)=\frac{1}{2}\|{\boldsymbol{\Gamma}}^{-1/2}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}({\mathcal{U}}(a)))\|_{2}^{2},\qquad Z=\mathbb{E}_{\pi_{0}}[\Theta(\cdot)]\;. (5.2)

Additional conditions ensure that the posterior measure π(|𝖉)\pi(\cdot|{\boldsymbol{\mathfrak{d}}}) is well-defined and that (5.1) holds according to the following result from [48].

Proposition 5.1.

Assume that 𝓞𝒰:Em{\boldsymbol{{\mathcal{O}}}}\circ{\mathcal{U}}:{E}\to{\mathbb{R}}^{m} is continuous and that π0(E)=1\pi_{0}({E})=1. Then the posterior π(|𝖉)\pi(\cdot|{\boldsymbol{\mathfrak{d}}}) is absolutely continuous with respect to π0\pi_{0}, and (5.2) holds.

The condition π0(E)=1\pi_{0}({E})=1 can in fact be weakened to π0(E)>0\pi_{0}({E})>0 (e.g. [48, Theorem 3.4]).

The solution of the BIP amounts to the evaluation of the posterior expectation 𝔼μ𝖉[]\mathbb{E}_{\mu^{\boldsymbol{\mathfrak{d}}}}[\cdot] of a continuous linear map ϕ:XQ\phi:X\to Q of the map 𝒰(a){\mathcal{U}}(a), where QQ is a suitable Hilbert space over \mathbb{C}. Solving the Bayesian inverse problem is thus closely related to the numerical approximation of the posterior expectation

𝔼π(|𝖉)[ϕ(𝒰())]Q.\mathbb{E}_{\pi(\cdot|{\boldsymbol{\mathfrak{d}}})}[\phi({\mathcal{U}}(\cdot))]\in Q.

For computational purposes, and to facilitate Wiener-Hermite PC approximation of the density Θ\Theta in (5.1), one parametrizes the input data a=a(𝒚)Ea=a({\boldsymbol{y}})\in E by a Gaussian series as discussed in Section 2.5. Inserting into Θ(a)\Theta(a) in (5.1), (5.2) this results in a countably-parametric density U𝒚Θ(a(𝒚))U\ni{\boldsymbol{y}}\mapsto\Theta(a({\boldsymbol{y}})), for 𝒚U{\boldsymbol{y}}\in U, and the Gaussian reference measure π0\pi_{0} on EE in (5.1) is pushed forward into a countable product γ\gamma of the sequence of Gaussian measures {γ1,n}n\{\gamma_{1,n}\}_{n\in{\mathbb{N}}} on {\mathbb{R}}: using (5.1) and choosing a Gaussian prior (e.g. [48, Section 2.4] or [66, 85])

π0=γ=jγ1,n\pi_{0}=\gamma=\bigotimes_{j\in{\mathbb{N}}}\gamma_{1,n}

on UU (see Example 2.17), the Bayesian estimate, i.e., the posterior expectation, can then be written as a (countably) iterated integral [98, 48, 96] with respect to the product GM γ\gamma, i.e.

𝔼π(|𝖉)[ϕ(𝒰(a()))]=1ZUϕ(𝒰(a(𝒚)))Θ(a(𝒚))dγ(𝒚)Q,Z=UΘ(a(𝒚))dγ(𝒚).\mathbb{E}_{\pi(\cdot|{\boldsymbol{\mathfrak{d}}})}[\phi({\mathcal{U}}(a(\cdot)))]=\frac{1}{Z}\int_{U}\phi({\mathcal{U}}(a({\boldsymbol{y}})))\Theta(a({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})\in Q,\quad Z=\int_{U}\Theta(a({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})\in{\mathbb{R}}. (5.3)

The parametric density UU\to{\mathbb{R}} in (5.3) which arises in Bayesian PDE inversion under Gaussian prior and also under more general, so-called Besov prior measures on UU, see, e.g. [48, Section 2.3], [66, 85]. The parametric density

𝒚ϕ(𝒰(a(𝒚)))Θ(a(𝒚)),{\boldsymbol{y}}\mapsto\phi({\mathcal{U}}(a({\boldsymbol{y}})))\Theta(a({\boldsymbol{y}}))\;,

inherits sparsity from the forward map 𝒚𝒰(a(𝒚)){\boldsymbol{y}}\mapsto{\mathcal{U}}(a({\boldsymbol{y}})), whose sparsity is expressed as before in terms of p\ell^{p}-summability and weighted 2\ell^{2}-summability of Wiener-Hermite PC expansion coefficients. We employ the parametric holomorphy of the forward map a𝒰(a)a\mapsto{\mathcal{U}}(a) to quantify the sparsity of the parametric posterior densities 𝒚Θ(a(𝒚)){\boldsymbol{y}}\mapsto\Theta(a({\boldsymbol{y}})) and 𝒚ϕ(𝒰(a(𝒚)))Θ(a(𝒚)){\boldsymbol{y}}\mapsto\phi({\mathcal{U}}(a({\boldsymbol{y}})))\Theta(a({\boldsymbol{y}})) in (5.3).

5.2 Posterior parametric holomorphy

With a Gaussian series in the data space EE, for the resulting parametric data-to-solution map

u:UX:𝒚𝒰(a(𝒚)),u:U\to X:{\boldsymbol{y}}\mapsto{\mathcal{U}}(a({\boldsymbol{y}})),

we now prove that under certain conditions both, the corresponding parametric posterior density

𝒚exp((𝖉𝓞(u(𝒚)))𝚪1(𝖉𝓞(u(𝒚)))){\boldsymbol{y}}\mapsto\exp\left(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))\right) (5.4)

in (5.2), and the integrand

𝒚ϕ(u(𝒚))exp((𝖉𝓞(u(𝒚)))𝚪1(𝖉𝓞(u(𝒚)))){\boldsymbol{y}}\mapsto\phi(u({\boldsymbol{y}}))\exp\left(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))\right) (5.5)

in (5.3) are (𝒃,ξ,δ,)({\boldsymbol{b}},\xi,\delta,\mathbb{C})-holomorphic and (𝒃,ξ,δ,Q)({\boldsymbol{b}},\xi,\delta,Q)-holomorphic, respectively.

Theorem 5.2.

Let r>0r>0. Assume that the map u:UXu:U\to X is (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic with constant functions φNr\varphi_{N}\equiv r, NN\in\mathbb{N}, in Definition 4.1. Let the observation noise covariance matrix 𝚪m×m{\boldsymbol{\Gamma}}\in\mathbb{R}^{m\times m} be symmetric positive definite.

Then, for any bounded linear quantity of interest ϕL(X,Q)\phi\in L(X,Q), and for any observable 𝓞(X)m{\boldsymbol{{\mathcal{O}}}}\in(X^{\prime})^{m} with arbitrary, finite mm, the function in (5.4) is (𝐛,ξ,δ,)({\boldsymbol{b}},\xi,\delta,\mathbb{C})-holomorphic and the function in (5.5) is (𝐛,ξ,δ,Q)({\boldsymbol{b}},\xi,\delta,Q)-holomorphic.

Proof.

We only show the statement for the parametric integrand in (5.5), as the argument for the posterior density in (5.4) is completely analogous.

Consider the map

Ξ:{vX:vXr}Q:vϕ(v)exp((𝖉𝓞(v))𝚪1(𝖉𝓞(v))).\Xi:\{v\in X\,:\,\|v\|_{X}\leq r\}\to Q:v\mapsto\phi(v)\exp(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(v)){\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(v))).

This function is well-defined. We have |𝓞(v)|𝓞Xr|{\boldsymbol{{\mathcal{O}}}}(v)|\leq\|{\boldsymbol{{\mathcal{O}}}}\|_{X^{\prime}}r and |ϕ(v)|ϕL(X;Q)r|\phi(v)|\leq\|\phi\|_{L(X;Q)}r for all vXv\in X with vXr\|v\|_{X}\leq r. Since exp:\exp:\mathbb{C}\to\mathbb{C} is Lipschitz continuous on compact subsets of \mathbb{C} and since ϕL(X;Q)\phi\in L(X;Q) is bounded linear map (and thus Lipschitz continuous), we find that

supvXrΞ(v)Q=:r~<\sup_{\|v\|_{X}\leq r}\|\Xi(v)\|_{Q}=:\tilde{r}<\infty

and that

Ξ:{vX:vXr}\Xi:\{v\in X\,:\,\|v\|_{X}\leq r\}\to\mathbb{C}

is Lipschitz continuous with some Lipschitz constant L>0L>0.

Let us recall that the (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy of u:UXu:U\to X, implies the existence of (continuous) functions uNL2(N,X;γN)u_{N}\in L^{2}(\mathbb{R}^{N},X;\gamma_{N}) such that with u~N(𝒚)=uN(y1,,yN)\tilde{u}_{N}({\boldsymbol{y}})=u_{N}(y_{1},\dots,y_{N}) it holds limNu~N=u\lim_{N\to\infty}\tilde{u}_{N}=u in the sense of L2(U,X;γ)L^{2}(U,X;\gamma). Furthermore, if

j=1Nbjϱjδ\sum_{j=1}^{N}b_{j}\varrho_{j}\leq\delta

(i.e. ϱ=(ϱj)j=1N{\boldsymbol{\varrho}}=(\varrho_{j})_{j=1}^{N} is (𝒃,ξ)({\boldsymbol{b}},\xi)-admissible in the sense of Definition 4.1), then uNu_{N} allows a holomorphic extension

uN:𝒮ϱXu_{N}:{\mathcal{S}}_{\boldsymbol{\varrho}}\to X

such that for all 𝒚N{\boldsymbol{y}}\in\mathbb{R}^{N}

sup𝒛ϱuN(𝒚+𝒛)XφN(𝒚)=r𝒚N,\sup_{{\boldsymbol{z}}\in{\mathcal{B}}_{\boldsymbol{\varrho}}}\|u_{N}({\boldsymbol{y}}+{\boldsymbol{z}})\|_{X}\leq\varphi_{N}({\boldsymbol{y}})=r\qquad\forall{\boldsymbol{y}}\in\mathbb{R}^{N}, (5.6)

see (4.1) for the definition of 𝒮ϱ{\mathcal{S}}_{\boldsymbol{\varrho}} and ϱ{\mathcal{B}}_{\boldsymbol{\varrho}}.

We want to show that f(𝒚):=Ξ(u(𝒚))f({\boldsymbol{y}}):=\Xi(u({\boldsymbol{y}})) is well-defined in L2(U,Q;γ)L^{2}(U,Q;\gamma), and given as the limit of the functions

f~N(𝒚)=fN((yj)j=1N)\tilde{f}_{N}({\boldsymbol{y}})=f_{N}((y_{j})_{j=1}^{N})

for all 𝒚U{\boldsymbol{y}}\in U and NN\in\mathbb{N}, where

fN((yj)j=1N)=Ξ(uN((yj)j=1N).f_{N}((y_{j})_{j=1}^{N})=\Xi(u_{N}((y_{j})_{j=1}^{N}).

Note at first that fN:NQf_{N}:\mathbb{R}^{N}\to Q is well-defined. In the case

j=1Nbjϱjδ,\sum_{j=1}^{N}b_{j}\varrho_{j}\leq\delta,

fNf_{N} allows a holomorphic extension fN:𝒮ϱXf_{N}:{\mathcal{S}}_{\boldsymbol{\varrho}}\to X given through ΞuN\Xi\circ u_{N}. Using (5.6), this extension satisfies for any NN\in\mathbb{N} and any (𝒃,ξ)({\boldsymbol{b}},\xi)-admissible ϱ(0,)N\varrho\in(0,\infty)^{N}

sup𝒛ϱ|fN(𝒚+𝒛)|supvXr|Ξ(v)|=r~𝒚N.\sup_{{\boldsymbol{z}}\in{\mathcal{B}}_{\boldsymbol{\varrho}}}|f_{N}({\boldsymbol{y}}+{\boldsymbol{z}})|\leq\sup_{\|v\|_{X}\leq r}|\Xi(v)|=\tilde{r}\qquad\forall{\boldsymbol{y}}\in\mathbb{R}^{N}.

This shows assumptions (i)-(ii) of Definition 4.1 for fN:NQf_{N}:\mathbb{R}^{N}\to Q.

Finally we show assumption (iii) of Definition 4.1. By assumption it holds limNu~N=u\lim_{N\to\infty}\tilde{u}_{N}=u in the sense of L2(U,X;γ)L^{2}(U,X;\gamma). Thus for f=Ξuf=\Xi\circ u and with fN=ΞuNf_{N}=\Xi\circ u_{N}

Uf(𝒚)fN(𝒚)Q2dγ(𝒚)\displaystyle\int_{U}\|f({\boldsymbol{y}})-f_{N}({\boldsymbol{y}})\|_{Q}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}}) =UΞ(u(𝒚))Ξ(uN(𝒚))Q2dγ(𝒚)\displaystyle=\int_{U}\|\Xi(u({\boldsymbol{y}}))-\Xi(u_{N}({\boldsymbol{y}}))\|_{Q}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}})
L2Uu(𝒚)uN(𝒚)X2dγ(𝒚),\displaystyle\leq L^{2}\int_{U}\|u({\boldsymbol{y}})-u_{N}({\boldsymbol{y}})\|_{X}^{2}\,\mathrm{d}\gamma({\boldsymbol{y}}),

which tends to 0 as NN\to\infty. Here we used that LL is a Lipschitz constant of Ξ\Xi. ∎

Let us now discuss which functions satisfy the requirements of Theorem 5.2. Additional to (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy, we had to assume boundedness of the holomorphic extensions in Definition 4.1. For functions of the type as in Theorem 4.11 u(𝒚)=limN𝒰(exp(j=1Nyjψj)),u({\boldsymbol{y}})=\lim_{N\to\infty}{\mathcal{U}}\left(\exp\left(\sum_{j=1}^{N}y_{j}\psi_{j}\right)\right), the following result gives sufficient conditions such that the assumptions of Theorem 5.2 are satisfied for the forward map.

Corollary 5.3.

Assume that 𝒰:OX{\mathcal{U}}:O\to X and (ψj)jE(\psi_{j})_{j\in\mathbb{N}}\subset{E} satisfy Assumptions (i), (iii) and (iv) of Theorem 4.11 and additionally for some r>0r>0

  1. (ii)

    𝒰(a)Xr\|{\mathcal{U}}(a)\|_{X}\leq r for all aOa\in O.

Then

u(𝒚)=limN𝒰(exp(j=1Nyjψj))L2(U,X;γ)u({\boldsymbol{y}})=\lim_{N\to\infty}{\mathcal{U}}\left(\exp\left(\sum_{j=1}^{N}y_{j}\psi_{j}\right)\right)\in L^{2}(U,X;\gamma)

is (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic with constant functions φNr\varphi_{N}\equiv r, NN\in\mathbb{N}, in Definition 4.1.

Proof.

By Theorem 4.11, uu is (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic. Recalling the construction of φN:N\varphi_{N}:\mathbb{R}^{N}\to\mathbb{R} in Step 3 of the proof of Theorem 4.11, we observe that φN\varphi_{N} can be chosen as φNr\varphi_{N}\equiv r. ∎

5.3 Example: parametric diffusion coefficient

We revisit the example of the diffusion equation with parametric log-Gaussian coefficient as introduced in Section 3.5 and used in Section 4.3.1. With the Lipschitz continuity of the data-to-solution map established in Section 4.3.1, we verify the well-posedness of the corresponding BIP.

We fix the dimension dd\in\mathbb{N} of the physical domain Dd{{D}}\subseteq\mathbb{R}^{d}, being a bounded Lipschitz domain, and choose E=L(D;)E=L^{\infty}({{D}};\mathbb{C}) and X=H01(D;)X=H_{0}^{1}({{D}};\mathbb{C}). We assume that fXf\in X^{\prime} and a0Ea_{0}\in E with

ρ(a0)>0.\rho(a_{0})>0.

For

aO:={aE:ρ(a)>0},a\in O:=\{a\in E\,:\,\rho(a)>0\},

let 𝒰(a){\mathcal{U}}(a) be the solution to the equation

div((a0+a)𝒰(a))=f in D,𝒰(a)=0 on D,\displaystyle-\operatorname{div}((a_{0}+a)\nabla{\mathcal{U}}(a))=f\text{ in }{{D}},\;\;{\mathcal{U}}(a)=0\text{ on }\partial{{D}}, (5.7)

for some fixed fXf\in X^{\prime}.

Due to

ρ(a0+a)ρ(a0)>0,\rho(a_{0}+a)\geq\rho(a_{0})>0,

for every aOa\in O, as in (4.20) we find that 𝒰(a){\mathcal{U}}(a) is well-defined and it holds

𝒰(a)XfXρ(a0)=:raO.\|{\mathcal{U}}(a)\|_{X}\leq\frac{\|f\|_{X^{\prime}}}{{\rho(a_{0})}}=:r\qquad\forall a\in O.

This shows assumption (ii) in Corollary 5.3. Slightly adjusting the arguments in Section 4.3.1 one observes that 𝒰:OX{\mathcal{U}}:O\to X satisfies assumptions (i) and (iii) in Theorem 4.11. Fix a representation system (ψj)jV(\psi_{j})_{j\in\mathbb{N}}\subseteq V such that with bj:=ψjEb_{j}:=\|\psi_{j}\|_{E} it holds (bj)j1()(b_{j})_{j\in\mathbb{N}}\in\ell^{1}(\mathbb{N}). Then Corollary 5.3 implies that the forward map

u(𝒚)=limN𝒰(exp(j=1Nyjψj))u({\boldsymbol{y}})=\lim_{N\to\infty}{\mathcal{U}}\bigg{(}\exp\Big{(}\sum_{j=1}^{N}y_{j}\psi_{j}\Big{)}\bigg{)}

satisfies the assumptions of Theorem 5.2. Theorem 5.2 in turn implies that the posterior density for this model is (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic. We shall prove in Section 6 that sparse-grid quadratures can be constructed which achieve higher order convergence for the integrands in (5.4) and (5.5), with the convergence rate being a decreasing function of p(0,4/5)p\in(0,4/5) such that 𝒃p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}), see Theorem 6.16. Furthermore, Theorem 4.9 implies a certain sparsity for the family of Wiener-Hermite PC expansion coefficients of the parametric maps in (5.4) and (5.5).

6 Smolyak sparse-grid interpolation and quadrature

Theorem 4.9 shows that if vv is (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝒃p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) and some p(0,1)p\in(0,1), then (v𝝂X)𝝂2p/(2p)().(\|v_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{2p/(2-p)}(\mathcal{F}). In Remark 4.10, based on this summability of the Wiener-Hermite PC expansion coefficients, we derived the convergence rate of best nn-term approximation as in (4.13). This approximation is not linear since the approximant is taken accordingly to the NN largest terms v𝝂X\|v_{\boldsymbol{\nu}}\|_{X}. To construct a linear approximation which gives the same convergence rate it is suitable to use the stronger weighted 2\ell^{2}-summability result (4.11) in Theorem 4.9.

In Theorem 4.9 of Section 4, we have obtained the weighted 2\ell^{2}-summability

𝝂β𝝂(r,ϱ)u𝝂X2<with(β𝝂(r,ϱ)1/2)𝝂p/(1p)(),\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{X}^{2}<\infty\ \ \ \text{with}\ \ \ \big{(}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{p/(1-p)}({\mathcal{F}}), (6.1)

for the norms of the Wiener-Hermite PC expansion coefficients of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions uu if 𝒃p(){\boldsymbol{b}}\in\ell^{p}({\mathbb{N}}) for some 0<p<10<p<1. In Section 4.2 and Section 5 we saw that solutions to certain parametric PDEs as well as posterior densities satisfy (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy.

The goal of this section is in a constructive way to sharpen and improve these results in a form more suitable for numerical implementation by using some ideas from [43, 45, 114]. We shall construct a new weight family (c𝝂)𝝂(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} based on (β𝝂(r,ϱ))𝝂(\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}))_{{\boldsymbol{\nu}}\in\mathcal{F}}, such that (6.1) with β𝝂(r,ϱ)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) replaced by c𝝂c_{\boldsymbol{\nu}}, and its generalization of the form (3.43) for σ𝝂=c𝝂1/2\sigma_{\boldsymbol{\nu}}=c_{\boldsymbol{\nu}}^{1/2} hold. Once a suitable family (c𝝂)𝝂(c_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} has been identified, we obtain a multiindex set Λε\Lambda_{\varepsilon}\subseteq\mathcal{F} for ε>0\varepsilon>0 via

Λε:={𝝂:c𝝂1ε},\Lambda_{\varepsilon}:=\{{\boldsymbol{\nu}}\in\mathcal{F}\,:\,c_{\boldsymbol{\nu}}^{-1}\geq\varepsilon\}, (6.2)

The set Λε\Lambda_{\varepsilon} will then serve as an index set to define interpolation operators 𝐈Λε\mathbf{I}_{\Lambda_{\varepsilon}} and quadrature operators 𝐐Λε\mathbf{Q}_{\Lambda_{\varepsilon}}. As the sequence (c𝝂)𝝂(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} is used to construct sets of multiindices, it should possess certain features, including each c𝝂c_{\boldsymbol{\nu}} to be easily computable for 𝝂{\boldsymbol{\nu}}\in\mathcal{F}, and for the resulting numerical algorithm to be efficient.

6.1 Smolyak sparse-grid interpolation and quadrature

6.1.1 Smolyak sparse-grid interpolation

Recall that for every n0n\in\mathbb{N}_{0} denote by (χn,j)j=0n(\chi_{n,j})_{j=0}^{n}\subseteq\mathbb{R} the Gauss-Hermite points in one dimension (in particular, χ0,0=0\chi_{0,0}=0), that is, the roots of Hermite polynomial Hn+1H_{n+1}. Let

In:C0()C0()I_{n}:C^{0}(\mathbb{R})\to C^{0}(\mathbb{R})

be the univariate polynomial Lagrange interpolation operator defined by

(Inu)(y):=j=0nu(χn,j)i=0ijnyχn,iχn,jχn,i,y,(I_{n}u)(y):=\sum_{j=0}^{n}u(\chi_{n,j})\prod_{\begin{subarray}{c}i=0\\ i\neq j\end{subarray}}^{n}\frac{y-\chi_{n,i}}{\chi_{n,j}-\chi_{n,i}},\qquad y\in\mathbb{R},

with convention that I1:C0()C0()I_{-1}:C^{0}(\mathbb{R})\to C^{0}(\mathbb{R}) is defined as the constant 0 operator.

For any multi-index 𝝂{\boldsymbol{\nu}}\in\mathcal{F}, introduce the tensorized operators 𝐈𝝂\mathbf{I}_{\boldsymbol{\nu}} by

𝐈𝟎u:=u((χ0,0)j),\mathbf{I}_{\boldsymbol{0}}u:=u((\chi_{0,0})_{j\in\mathbb{N}}),

and for 𝝂𝟎{\boldsymbol{\nu}}\neq{\boldsymbol{0}} via

𝐈𝝂:=jIνj,\mathbf{I}_{\boldsymbol{\nu}}:=\bigotimes_{j\in\mathbb{N}}I_{\nu_{j}}, (6.3)

i.e.,

𝐈𝝂u(𝒚)={𝝁:𝝁𝝂}u((χνj,μj)j)ji=0iμjνjyjχνj,iχνj,μjχνj,i,𝒚U.\mathbf{I}_{\boldsymbol{\nu}}u({\boldsymbol{y}})=\sum_{\{{\boldsymbol{\mu}}\in\mathcal{F}\,:\,{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}\}}u((\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}})\prod_{j\in\mathbb{N}}\prod_{\begin{subarray}{c}i=0\\ i\neq\mu_{j}\end{subarray}}^{\nu_{j}}\frac{y_{j}-\chi_{\nu_{j},i}}{\chi_{\nu_{j},\mu_{j}}-\chi_{\nu_{j},i}},\quad{\boldsymbol{y}}\in U.

The operator 𝐈𝝂\mathbf{I}_{\boldsymbol{\nu}} can thus be applied to functions uu which are pointwise defined at each (χνj,μj)jU(\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}}\in U. Via Remark 4.4, we can apply it in particular to (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. Observe that the product over jj\in\mathbb{N} in (6.3) is a finite product, since for every jj with νj=0\nu_{j}=0, the inner product over i{0,,μj1,μj+1,,νj}i\in\{0,\dots,\mu_{j}-1,\mu_{j}+1,\dots,\nu_{j}\} is over an empty set, and therefore equal to one by convention. Then for a finite set Λ\Lambda\subseteq\mathcal{F}

𝐈Λ:=𝝂Λj(IνjIνj1).\mathbf{I}_{\Lambda}:=\sum_{{\boldsymbol{\nu}}\in\Lambda}\bigotimes_{j\in\mathbb{N}}(I_{\nu_{j}}-I_{\nu_{j}-1}). (6.4)

Expanding all tensor product operators, we get

𝐈Λ=𝝂ΛσΛ;𝝂𝐈𝝂whereσΛ;𝝂:={𝒆{0,1}:𝝂+𝒆Λ}(1)|𝒆|.\mathbf{I}_{\Lambda}=\sum_{{\boldsymbol{\nu}}\in\Lambda}\sigma_{\Lambda;{\boldsymbol{\nu}}}\mathbf{I}_{\boldsymbol{\nu}}\qquad\text{where}\qquad\sigma_{\Lambda;{\boldsymbol{\nu}}}:=\sum_{\{{\boldsymbol{e}}\in\{0,1\}^{\infty}\,:\,{\boldsymbol{\nu}}+{\boldsymbol{e}}\in\Lambda\}}(-1)^{|{\boldsymbol{e}}|}. (6.5)
Definition 6.1.

An index set Λ\Lambda\subseteq\mathcal{F} is called downward closed, if it is finite and if for every 𝛎Λ{\boldsymbol{\nu}}\in\Lambda it holds 𝛍Λ{\boldsymbol{\mu}}\in\Lambda whenever 𝛍𝛎{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}. Here, the ordering “\leq” between two indices 𝛍=(μj)j{\boldsymbol{\mu}}=(\mu_{j})_{j\in{\mathbb{N}}} and 𝛎=(νj)j{\boldsymbol{\nu}}=(\nu_{j})_{j\in{\mathbb{N}}} in \mathcal{F} expresses that for all jj\in{\mathbb{N}} holds μjνj\mu_{j}\leq\nu_{j} with strict inequality for at least one index jj.

As is well-known, 𝐈Λ\mathbf{I}_{\Lambda} possesses the following crucial property, see for example [111, Lemma 1.3.3].

Lemma 6.2.

Let Λ\Lambda\subseteq\mathcal{F} be downward closed. Then 𝐈Λf=f\mathbf{I}_{\Lambda}f=f for all fspan{𝐲𝛎:𝛎Λ}f\in{\rm span}\{{\boldsymbol{y}}^{\boldsymbol{\nu}}\,:\,{\boldsymbol{\nu}}\in\Lambda\}.

The reason to choose the collocation points (χn,j)j=0n(\chi_{n,j})_{j=0}^{n} as the Gauss-Hermite points, is that it was recently shown that the interpolation operators InI_{n} then satisfy the following stability estimate, see [52, Lemma 3.13].

Lemma 6.3.

For every n0n\in\mathbb{N}_{0} and every mm\in\mathbb{N} it holds

In(Hm)L2(;γ1)42m1.\|I_{n}(H_{m})\|_{L^{2}(\mathbb{R};\gamma_{1})}\leq 4\sqrt{2m-1}.

With the presently adopted normalization of the GM γ1\gamma_{1}, it holds H01H_{0}\equiv 1 and therefore In(H0)=H0I_{n}(H_{0})=H_{0} for all n0n\in\mathbb{N}_{0} (since the interpolation operator InI_{n} exactly reproduces all polynomials of degree n0n\in\mathbb{N}_{0}). Hence

In(H0)L2(;γ1)=H0L2(;γ1)=1\|I_{n}(H_{0})\|_{L^{2}(\mathbb{R};\gamma_{1})}=\|H_{0}\|_{L^{2}(\mathbb{R};\gamma_{1})}=1

for all n0n\in\mathbb{N}_{0}. Noting that 42m1(1+m)24\sqrt{2m-1}\leq(1+m)^{2} for all mm\in\mathbb{N}, we get

In(Hm)L2(;γ1)(1+m)2n,m0.\|I_{n}(H_{m})\|_{L^{2}(\mathbb{R};\gamma_{1})}\leq(1+m)^{2}\qquad\forall n,\leavevmode\nobreak\ m\in\mathbb{N}_{0}.

Consequently

𝐈𝝂(H𝝁)L2(U;γ)=jIνj(Hμj)L2(;γ1)j(1+μj)2𝝂,𝝁.\|\mathbf{I}_{\boldsymbol{\nu}}(H_{\boldsymbol{\mu}})\|_{L^{2}(U;\gamma)}=\prod_{j\in\mathbb{N}}\|I_{\nu_{j}}(H_{\mu_{j}})\|_{L^{2}(\mathbb{R};\gamma_{1})}\leq\prod_{j\in\mathbb{N}}(1+\mu_{j})^{2}\qquad\forall{\boldsymbol{\nu}},\leavevmode\nobreak\ {\boldsymbol{\mu}}\in\mathcal{F}. (6.6)

Recall that for 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}} and τ0\tau\geq 0, we denote

p𝝂(τ):=j(1+νj)τ.p_{\boldsymbol{\nu}}(\tau):=\prod_{j\in{\mathbb{N}}}(1+\nu_{j})^{\tau}.

If νj>μj\nu_{j}>\mu_{j} then (IνjIνj1)Hμj=0.(I_{\nu_{j}}-I_{\nu_{j}-1})H_{\mu_{j}}=0. Thus,

j(IνjIνj1)H𝝁=0,\bigotimes_{j\in\mathbb{N}}(I_{\nu_{j}}-I_{\nu_{j}-1})H_{\boldsymbol{\mu}}=0,

whenever there exists jj\in\mathbb{N} such that νj>μj\nu_{j}>\mu_{j}. Hence, for any downward closed set Λ\Lambda, it holds

𝐈Λ(H𝝁)L2(U;γ)p𝝁(3).\|\mathbf{I}_{\Lambda}(H_{\boldsymbol{\mu}})\|_{L^{2}(U;\gamma)}\leq p_{{\boldsymbol{\mu}}}(3). (6.7)

Indeed,

𝐈Λ(H𝝁)L2(U;γ){𝝂Λ:𝝂𝝁}p𝝁(2)|{𝝂Λ:𝝂𝝁}|p𝝁(2)=j(1+μj)p𝝁(2)=p𝝁(3).\|\mathbf{I}_{\Lambda}(H_{\boldsymbol{\mu}})\|_{L^{2}(U;\gamma)}\leq\sum_{\{{\boldsymbol{\nu}}\in\Lambda\,:\,{\boldsymbol{\nu}}\leq{\boldsymbol{\mu}}\}}p_{\boldsymbol{\mu}}(2)\leq|{\{{\boldsymbol{\nu}}\in\Lambda\,:\,{\boldsymbol{\nu}}\leq{\boldsymbol{\mu}}\}}|p_{\boldsymbol{\mu}}(2)=\prod_{j\in\mathbb{N}}(1+\mu_{j})p_{\boldsymbol{\mu}}(2)=p_{{\boldsymbol{\mu}}}(3).

6.1.2 Smolyak sparse-grid quadrature

Recall that analogously to InI_{n} we introduce univariate polynomial quadrature operators via

Qnu:=j=0nu(χn,j)ωn,j,ωn,j:=ijyχn,iχn,jχn,idγ1(y).Q_{n}u:=\sum_{j=0}^{n}u(\chi_{n,j})\omega_{n,j},\qquad\omega_{n,j}:=\int_{\mathbb{R}}\prod_{i\neq j}\frac{y-\chi_{n,i}}{\chi_{n,j}-\chi_{n,i}}\,\mathrm{d}\gamma_{1}(y).

Furthermore, we define

𝐐𝟎u:=u((χ0,0)j),\mathbf{Q}_{\boldsymbol{0}}u:=u((\chi_{0,0})_{j\in\mathbb{N}}),

and for 𝝂𝟎{\boldsymbol{\nu}}\neq{\boldsymbol{0}},

𝐐𝝂:=jQνj,\mathbf{Q}_{\boldsymbol{\nu}}:=\bigotimes_{j\in\mathbb{N}}Q_{\nu_{j}},

i.e.,

𝐐𝝂u={𝝁:𝝁𝝂}u((χνj,μj)j)jωνj,μj,\mathbf{Q}_{\boldsymbol{\nu}}u=\sum_{\{{\boldsymbol{\mu}}\in\mathcal{F}\,:\,{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}\}}u((\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}})\prod_{j\in\mathbb{N}}\omega_{\nu_{j},\mu_{j}},

and finally for a finite downward closed Λ\Lambda\subseteq\mathcal{F} with σΛ;𝝂\sigma_{\Lambda;{\boldsymbol{\nu}}} as in (6.5),

𝐐Λ:=𝝂ΛσΛ;𝝂Q𝝂.\mathbf{Q}_{\Lambda}:=\sum_{{\boldsymbol{\nu}}\in\Lambda}\sigma_{\Lambda;{\boldsymbol{\nu}}}Q_{\boldsymbol{\nu}}.

Again we emphasize that the above formulas are meaningful as long as point evaluations of uu at each (χνj,μj)j(\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}} are well defined, 𝝂{\boldsymbol{\nu}}\in\mathcal{F}, 𝝁𝝂{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}. Also note that

𝐐Λf=U𝐈Λf(𝒚)dγ(𝒚).\mathbf{Q}_{\Lambda}f=\int_{U}\mathbf{I}_{\Lambda}f({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}). (6.8)

Recall that the set 2\mathcal{F}_{2} is defined by

2:={𝝂:νj1j}.\mathcal{F}_{2}:=\{{\boldsymbol{\nu}}\in\mathcal{F}\,:\,\nu_{j}\neq 1\leavevmode\nobreak\ \forall j\}. (6.9)

We thus have 2\mathcal{F}_{2}\subsetneq\mathcal{F}. Similar to Lemma 6.2 we have the following lemma, which can be proven completely analogous to [111, Lemma 1.3.16] (also see [114, Remark 4.2]).

Lemma 6.4.

Let Λ\Lambda\subseteq\mathcal{F} be downward closed. Then

𝐐Λv=Uv(𝒚)dγ(𝒚)\mathbf{Q}_{\Lambda}v=\int_{U}v({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})

for all vspan{𝐲𝛎:𝛎Λ(\2)}v\in{\rm span}\{{\boldsymbol{y}}^{\boldsymbol{\nu}}\,:\,{\boldsymbol{\nu}}\in\Lambda\cup(\mathcal{F}\backslash\mathcal{F}_{2})\}.

With (6.6) it holds

|𝐐𝝂(H𝝁)|=|U𝐈𝝂(H𝝁)(𝒚)dγ(𝒚)|𝐈𝝂(H𝝁)L2(U;γ)j(1+μj)2𝝂,𝝁,|\mathbf{Q}_{\boldsymbol{\nu}}(H_{\boldsymbol{\mu}})|=\left|\int_{U}\mathbf{I}_{\boldsymbol{\nu}}(H_{\boldsymbol{\mu}})({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})\right|\leq\|\mathbf{I}_{\boldsymbol{\nu}}(H_{\boldsymbol{\mu}})\|_{L^{2}(U;\gamma)}\leq\prod_{j\in\mathbb{N}}(1+\mu_{j})^{2}\qquad\forall{\boldsymbol{\nu}},\leavevmode\nobreak\ {\boldsymbol{\mu}}\in\mathcal{F},

and similarly, using (6.7), we have the bound

|𝐐Λ(H𝝁)|p𝝁(3).|\mathbf{Q}_{\Lambda}(H_{\boldsymbol{\mu}})|\leq p_{{\boldsymbol{\mu}}}(3). (6.10)

6.2 Multiindex sets

In this section, we first recall some arguments from [43, 45, 114] which allow to bound the number of required function evaluations in the interpolation an quadrature algorithm. Subsequently, a construction of a suitable family (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} is provided for k{1,2}k\in\{1,2\}. The index kk determines whether the family will be used for a sparse-grid interpolation (k=1k=1) or a Smolyak-type sparse-grid quadrature (k=2k=2) algorithm. Finally, it is shown that the multiindex sets Λk,ε\Lambda_{k,\varepsilon} as in (6.2) based on (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, guarantee algebraic convergence rates for certain truncated Wiener-Hermite PC expansions. This will be exploited to verify convergence rates for interpolation in Section 6.3 and for quadrature in Section 6.4.

6.2.1 Number of function evaluations

In order to obtain a convergence rate in terms of the number of evaluations of uu, we need to determine the number of interpolation points used by the operator 𝐈Λ\mathbf{I}_{\Lambda} or 𝐐Λ\mathbf{Q}_{\Lambda}. Since the discussion of 𝐐Λ\mathbf{Q}_{\Lambda} is very similar, we concentrate here on 𝐈Λ\mathbf{I}_{\Lambda}.

Computing the interpolant 𝐈𝝂u\mathbf{I}_{{\boldsymbol{\nu}}}u in (6.3) requires knowledge of the function values of uu at each point in

{(χνj,μj)j:𝝁𝝂}.\{(\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}}\,:\,{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}\}.

The cardinality of this set is bounded by j(1+νj)=p𝝂(1)\prod_{j\in\mathbb{N}}(1+\nu_{j})=p_{\boldsymbol{\nu}}(1). Denote by

pts(Λ):={(χνj,μj)j:𝝁𝝂,𝝂Λ}{\rm pts}(\Lambda):=\{(\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}}\,:\,{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}},\leavevmode\nobreak\ {\boldsymbol{\nu}}\in\Lambda\} (6.11)

the set of interpolation points defining the interpolation operator 𝐈Λ\mathbf{I}_{\Lambda} (i.e.,  |pts(Λ)||{\rm pts}(\Lambda)| is the number of function evaluations of uu required to compute 𝐈Λu\mathbf{I}_{\Lambda}u). By (6.5) we obtain the bound

|pts(Λ)|{𝝂Λ:σΛ,𝝂0}j(1+νj)={𝝂Λ:σΛ,𝝂0}p𝝂(1).|{\rm pts}(\Lambda)|\leq\sum_{\{{\boldsymbol{\nu}}\in\Lambda\,:\,\sigma_{\Lambda,{\boldsymbol{\nu}}}\neq 0\}}\prod_{j\in\mathbb{N}}(1+\nu_{j})=\sum_{\{{\boldsymbol{\nu}}\in\Lambda\,:\,\sigma_{\Lambda,{\boldsymbol{\nu}}}\neq 0\}}p_{\boldsymbol{\nu}}(1). (6.12)

6.2.2 Construction of (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}

We are now in position to construct (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}. As mentioned above, we distinguish between the cases k=1k=1 and k=2k=2, which correspond to polynomial interpolation or quadrature. Note that in the next lemma we define ck,𝝂c_{k,{\boldsymbol{\nu}}} for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F}, but the estimate provided in the lemma merely holds for 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k}, k{1,2}k\in\{1,2\}, where 1:=\mathcal{F}_{1}:=\mathcal{F} and 2\mathcal{F}_{2} is defined in (6.9). Throughout what follows, empty products shall equal 11 by convention.

Lemma 6.5.

Assume that τ>0\tau>0, k{1,2}k\in\{1,2\} and r>max{τ,k}r>\max\{\tau,k\}. Let ϱ(0,){\boldsymbol{\varrho}}\in(0,\infty)^{\infty} be such that ϱj\varrho_{j}\to\infty as jj\to\infty.

Then there exist K>0K>0 and C0>0C_{0}>0 such that

ck,𝝂:=jsupp(𝝂)max{1,Kϱj}2kνjrτ,𝝂,c_{k,{\boldsymbol{\nu}}}:=\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\max\left\{1,K\varrho_{j}\right\}^{2k}\nu_{j}^{r-\tau},\quad{\boldsymbol{\nu}}\in\mathcal{F}, (6.13)

satisfies

C0ck,𝝂p𝝂(τ)β𝝂(r,ϱ)𝝂kC_{0}c_{k,{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(\tau)\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\quad\forall{\boldsymbol{\nu}}\in\mathcal{F}_{k} (6.14)

with β𝛎(r,ϱ)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) as in (3.36).

Proof.

Step 1. Fix 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k}, then jsupp(𝝂)j\in\operatorname{supp}({\boldsymbol{\nu}}) implies νjk\nu_{j}\geq k and thus min{r,νj}k\min\{r,\nu_{j}\}\geq k since r>kr>k by assumption. With s:=min{r,νj}νjs:=\min\{r,\nu_{j}\}\leq\nu_{j}, for all jj\in\mathbb{N} holds

(νjs)=νj!(νjs)!s!1s!(νjs+1)sνjs1s!ssνjs1r!rr=νjmin{νj,r}1r!rrνjr1r!r2r.\binom{\nu_{j}}{s}=\frac{\nu_{j}!}{(\nu_{j}-s)!s!}\geq\frac{1}{s!}(\nu_{j}-s+1)^{s}\geq\nu_{j}^{s}\frac{1}{s!s^{s}}\geq\nu_{j}^{s}\frac{1}{r!r^{r}}=\nu_{j}^{\min\{\nu_{j},r\}}\frac{1}{r!r^{r}}\geq\nu_{j}^{r}\frac{1}{r!r^{2r}}.

Furthermore, if jsupp(𝝂)j\in\operatorname{supp}({\boldsymbol{\nu}}), then due to s=min{νj,r}ks=\min\{\nu_{j},r\}\geq k, with ϱ0:=min{1,minjϱj}\varrho_{0}:=\min\{1,\min_{j\in\mathbb{N}}\varrho_{j}\} we have

ϱ02rmin{1,ϱj}2rϱj2(sk).\varrho_{0}^{2r}\leq\min\{1,\varrho_{j}\}^{2r}\leq\varrho_{j}^{2(s-k)}.

Thus

ϱjmin{νj,r}ϱ02rϱj2k\varrho_{j}^{\min\{\nu_{j},r\}}\geq\varrho_{0}^{2r}\varrho_{j}^{2k}

for all jj\in\mathbb{N}. In all, we conclude

β𝝂(r,ϱ)=j(l=0r(νjl)ϱj2l)jsupp(𝝂)(νjmin{νj,r})ϱj2min{νj,r}jsupp(𝝂)ϱ02rr!r2rϱj2kνjr.\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})=\prod_{j\in\mathbb{N}}\left(\sum_{l=0}^{r}\binom{\nu_{j}}{l}\varrho_{j}^{2l}\right)\geq\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\binom{\nu_{j}}{\min\{\nu_{j},r\}}\varrho_{j}^{2\min\{\nu_{j},r\}}\geq\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\frac{\varrho_{0}^{2r}}{r!r^{2r}}\varrho_{j}^{2k}\nu_{j}^{r}. (6.15)

Since 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k} was arbitrary, this estimate holds for all 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k}.

Step 2. Denote ϱ^j:=max{1,Kϱj}\hat{\varrho}_{j}:=\max\{1,K\varrho_{j}\}, where K>0K>0 is still at our disposal. We have

p𝝂(τ)jsupp(𝝂)2τνjτp_{\boldsymbol{\nu}}(\tau)\leq\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}2^{\tau}\nu_{j}^{\tau}

and thus

ck,𝝂p𝝂(τ)jsupp(𝝂)2τϱ^j2kνjr.c_{k,{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(\tau)\leq\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}2^{\tau}\hat{\varrho}_{j}^{2k}\nu_{j}^{r}. (6.16)

Again, this estimate holds for any 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k}.

With ϱ0:=min{1,minjϱj}\varrho_{0}:=\min\{1,\min_{j\in\mathbb{N}}\varrho_{j}\} denote

Cb:=(ϱ02rr!r2r)1/(2k)andCc:=(2τ)1/(2k).C_{b}:=\left(\frac{\varrho_{0}^{2r}}{r!r^{2r}}\right)^{1/(2k)}\qquad\text{and}\qquad C_{c}:=(2^{\tau})^{1/(2k)}.

Set

K:=CbCc,ϱ~j=KϱjK:=\frac{C_{b}}{C_{c}},\qquad\tilde{\varrho}_{j}=K\varrho_{j}

for all jj\in\mathbb{N}. Then

Cbϱj=Ccϱ~j=Ccϱ^j{1if Kϱj1,Kϱjif Kϱj<1.C_{b}\varrho_{j}=C_{c}\tilde{\varrho}_{j}=C_{c}\hat{\varrho}_{j}\begin{cases}1&\text{if }K\varrho_{j}\geq 1,\\ K\varrho_{j}&\text{if }K\varrho_{j}<1.\end{cases}

Let

C0:={j:Kϱj<1}(Kϱj)2kC_{0}:=\prod_{\{j\in\mathbb{N}\,:\,K\varrho_{j}<1\}}(K\varrho_{j})^{2k}

and note that this product is over a finite number of indices, since ϱj\varrho_{j}\to\infty as jj\to\infty. Then for any 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k}

jsupp(𝝂)Ccϱ~jC012kjsupp(𝝂)Ccϱ^j.\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}C_{c}\tilde{\varrho}_{j}\geq C_{0}^{\frac{1}{2k}}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}C_{c}\hat{\varrho}_{j}.

With (6.15) and (6.16) we thus obtain for every 𝝂k{\boldsymbol{\nu}}\in\mathcal{F}_{k},

β𝝂(r,ϱ)\displaystyle\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) jsupp(𝝂)(Cbϱj)2kνjr=jsupp(𝝂)(Ccϱ~j)2kνjr\displaystyle\geq\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}(C_{b}\varrho_{j})^{2k}\nu_{j}^{r}=\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}\left(C_{c}\tilde{\varrho}_{j}\right)^{2k}\nu_{j}^{r}
C0jsupp(𝝂)(Ccϱ^j)2kνjrC0ck,𝝂p𝝂(τ).\displaystyle\geq C_{0}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}(C_{c}\hat{\varrho}_{j})^{2k}\nu_{j}^{r}\geq C_{0}c_{k,{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(\tau).

6.2.3 Summability properties of the collection (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}

First we discuss the summability of the collection (ck,𝝂)𝝂(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}. We will require the following lemma which is a modification of [43, Lemma 6.2].

Lemma 6.6.

Let θ0\theta\geq 0. Let further k{1,2}k\in\{1,2\}, τ>0\tau>0, r>max{k,τ}r>\max\{k,\tau\} and q>0q>0 be such that (rτ)q/(2k)θ>1(r-\tau)q/(2k)-\theta>1. Assume that (ϱj)j(0,)(\varrho_{j})_{j\in\mathbb{N}}\in(0,\infty)^{\infty} satisfies (ϱj1)jq()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{q}(\mathbb{N}). Then with (ck,𝛎)𝛎(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} as in Lemma 6.5 it holds

𝝂p𝝂(θ)ck,𝝂q2k<.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(\theta)c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}<\infty.
Proof.

This lemma can be proven in the same way as the proof of [43, Lemma 6.2]. We provide a proof for completeness. With ϱ^j:=max{1,Kϱj}\hat{\varrho}_{j}:=\max\{1,K\varrho_{j}\} it holds (ϱ^j1)jq()(\hat{\varrho}_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{q}(\mathbb{N}). By definition of ck,𝝂c_{k,{\boldsymbol{\nu}}}, factorizing, we get

𝝂p𝝂(θ)ck,𝝂q2k=𝝂jsupp(𝝂)(1+νj)θ(ϱ^j2kνjrτ)q2kj(2θϱ^jqnnq(rτ)2knθ).\begin{split}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(\theta)c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}&=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\prod_{j\in\operatorname{supp}({\boldsymbol{\nu}})}(1+\nu_{j})^{\theta}\left(\hat{\varrho}_{j}^{2k}\nu_{j}^{r-\tau}\right)^{-\frac{q}{2k}}\leq\prod_{j\in\mathbb{N}}\left(2^{\theta}\hat{\varrho}_{j}^{-q}\sum_{n\in\mathbb{N}}n^{\frac{-q(r-\tau)}{2k}}n^{\theta}\right).\end{split}

The sum over nn equals some finite constant CC since by assumption q(rτ)/2kθ>1q(r-\tau)/2k-\theta>1. Using the inequality log(1+x)x\log(1+x)\leq x for all x>0x>0, we get

𝝂ck,𝝂q2kj(1+Cϱ^jq)=exp(jlog(1+Cϱ^jq))exp(jCϱ^jq),\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}\leq\prod_{j\in\mathbb{N}}\left(1+C\hat{\varrho}_{j}^{-q}\right)=\exp\left(\sum_{j\in\mathbb{N}}\log(1+C\hat{\varrho}_{j}^{-q})\right)\leq\exp\left(\sum_{j\in\mathbb{N}}C\hat{\varrho}_{j}^{-q}\right),

which is finite since (ϱ^j1)q()(\hat{\varrho}_{j}^{-1})\in\ell^{q}(\mathbb{N}). ∎

Based on (6.2), for ε>0\varepsilon>0 and k{1,2}k\in\{1,2\} let

Λk,ε:={𝝂:ck,𝝂1ε}.\Lambda_{k,\varepsilon}:=\{{\boldsymbol{\nu}}\in\mathcal{F}:c_{k,{\boldsymbol{\nu}}}^{-1}\geq\varepsilon\}\subseteq\mathcal{F}. (6.17)

The summability shown in Lemma 6.6 implies algebraic convergence rates of the tail sum as provided by the following proposition. This is well-known and follows by Stechkin’s lemma [102] which itself is a simple consequence of Hölder’s inequality.

Proposition 6.7.

Let k{1,2}k\in\{1,2\}, τ>0\tau>0, and q>0q>0. Let (ϱj1)jq()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{q}(\mathbb{N}) and r>max{k,τ}r>\max\{k,\tau\}, (rτ)q/(2k)>2(r-\tau)q/(2k)>2. Assume that (a𝛎)𝛎[0,)(a_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in[0,\infty)^{\infty} is such that

𝝂β𝝂(r,ϱ)a𝝂2<.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})a_{\boldsymbol{\nu}}^{2}<\infty. (6.18)

Then there exists a constant CC solely depending on (ck,𝛎)𝛎(c_{k,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} in (6.13) such that for all ε>0\varepsilon>0 it holds that

𝝂k\Λk,εp𝝂(τ)a𝝂C(𝝂β𝝂(r,ϱ)a𝝂2)12ε12q4k,\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\backslash\Lambda_{k,\varepsilon}}p_{\boldsymbol{\nu}}(\tau)a_{\boldsymbol{\nu}}\leq C\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})a_{\boldsymbol{\nu}}^{2}\right)^{\frac{1}{2}}\varepsilon^{\frac{1}{2}-\frac{q}{4k}},

and

|pts(Λk,ϵ)|Cεq2k.|{\rm pts}(\Lambda_{k,\epsilon})|\leq C\varepsilon^{-\frac{q}{2k}}. (6.19)
Proof.

We estimate

𝝂k\Λk,εp𝝂(τ)a𝝂(𝝂k\Λk,εp𝝂(τ)2a𝝂2ck,𝝂)1/2(𝝂k\Λk,εck,𝝂1)1/2.\begin{split}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\backslash\Lambda_{k,\varepsilon}}p_{\boldsymbol{\nu}}(\tau)a_{\boldsymbol{\nu}}&\leq\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\backslash\Lambda_{k,\varepsilon}}p_{\boldsymbol{\nu}}(\tau)^{2}a_{\boldsymbol{\nu}}^{2}c_{k,{\boldsymbol{\nu}}}\Bigg{)}^{1/2}\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\backslash\Lambda_{k,\varepsilon}}c_{k,{\boldsymbol{\nu}}}^{-1}\Bigg{)}^{1/2}.\end{split}

The first sum is finite by (6.18) and because C0p𝝂(τ)2ck,𝝂β𝝂(r,ϱ)C_{0}p_{\boldsymbol{\nu}}(\tau)^{2}c_{k,{\boldsymbol{\nu}}}\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}) according to (6.14). By Lemma 6.6 and (6.17) we obtain

𝝂k\Λk,εck,𝝂1=ck,𝝂1<εck,𝝂q2kck,𝝂1+q2kCε1q2k\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\backslash\Lambda_{k,\varepsilon}}c_{k,{\boldsymbol{\nu}}}^{-1}=\sum_{c_{k,{\boldsymbol{\nu}}}^{-1}<\varepsilon}c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}c_{k,{\boldsymbol{\nu}}}^{-1+\frac{q}{2k}}\leq C\varepsilon^{1-\frac{q}{2k}}

which proves the first statement. Moreover, for each 𝝂{\boldsymbol{\nu}}\in{\mathcal{F}}, the number of interpolation (quadrature) points is p𝝂(1)p_{\boldsymbol{\nu}}(1). Hence

|pts(Λk,ϵ)|=𝝂Λk,ϵp𝝂(1)=ck,𝝂1εp𝝂(1)ck,𝝂q2kck,𝝂q2kεq2k𝝂kp𝝂(1)ck,𝝂q2kCεq2k|{\rm pts}(\Lambda_{k,\epsilon})|=\sum_{{\boldsymbol{\nu}}\in\Lambda_{k,\epsilon}}p_{\boldsymbol{\nu}}(1)=\sum_{c_{k,{\boldsymbol{\nu}}}^{-1}\geq\varepsilon}p_{\boldsymbol{\nu}}(1)c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}c_{k,{\boldsymbol{\nu}}}^{\frac{q}{2k}}\leq\varepsilon^{-\frac{q}{2k}}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{k}}p_{\boldsymbol{\nu}}(1)c_{k,{\boldsymbol{\nu}}}^{-\frac{q}{2k}}\leq C\varepsilon^{-\frac{q}{2k}}

again by Lemma 6.6 and (6.17). ∎

6.2.4 Computing Λε\Lambda_{\varepsilon}

Having identified appropriate sequences (c𝝂)𝝂(c_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, in order to be able to implement the Smolyak sparse-grid interpolation operator 𝐈Λε\mathbf{I}_{\Lambda_{\varepsilon}} and the Smolyak sparse-grid quadrature operator 𝐐Λε\mathbf{Q}_{\Lambda_{\varepsilon}}, in practice it remains to compute the sets Λε=\Lambda_{\varepsilon}= in (6.2). We now recall Algorithm 2 in [111, Sec. 3.1.3] which achieves this in O(|Λε|)O(|\Lambda_{\varepsilon}|) work and memory. For the convenience of the reader we recall the main statement regarding the algorithm’s complexity below in Lemma 6.8. Additionally, we point to [19, Alg. 4.13] which presents an alternative approach—a recursive algorithm that also achieves linear computational complexity.

In the following denote 𝒆j:=(δij)j0{\boldsymbol{e}}_{j}:=(\delta_{ij})_{j\in\mathbb{N}}\in\mathbb{N}_{0}^{\infty}.

Algorithm 1 Lambda(ε,(c𝝂)𝝂)\varepsilon,(c_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}))
1:𝝂𝟎{\boldsymbol{\nu}}\leftarrow{\boldsymbol{0}}
2:if c𝝂<εc_{\boldsymbol{\nu}}<\varepsilon then
3:     Λ\Lambda\leftarrow\emptyset
4:     return Λ\Lambda
5:else
6:     Λ{𝝂}\Lambda\leftarrow\{{\boldsymbol{\nu}}\}
7:while True do
8:     d1d\leftarrow 1
9:     while a𝝂+𝒆d<εa_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{d}}<\varepsilon do
10:         if νd0\nu_{d}\neq 0 then\triangleright Reject 𝝂+𝒆d{\boldsymbol{\nu}}+{\boldsymbol{e}}_{d} where νd0\nu_{d}\neq 0
11:              νd0\nu_{d}\leftarrow 0
12:              dd+1d\leftarrow d+1
13:         else if 𝝂𝟎{\boldsymbol{\nu}}\neq{\boldsymbol{0}} then \triangleright Reject 𝝂+𝒆d{\boldsymbol{\nu}}+{\boldsymbol{e}}_{d} where νd=0\nu_{d}=0
14:              d=min{j:νj0}d=\min\{j\in\mathbb{N}\,:\,\nu_{j}\neq 0\}
15:         else\triangleright Reject 𝒆d{\boldsymbol{e}}_{d} \Rightarrow stop algorithm
16:              return Λ\Lambda               
17:     𝝂𝝂+𝒆d{\boldsymbol{\nu}}\leftarrow{\boldsymbol{\nu}}+{\boldsymbol{e}}_{d}
18:     ΛΛ{𝝂}\Lambda\leftarrow\Lambda\cup\{{\boldsymbol{\nu}}\}

The algorithm is of linear complexity in the following sense [111, 3.1.12]:

Lemma 6.8.

Let (c𝛎)𝛎[0,)(c_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subseteq[0,\infty) be a null-sequence such that (i) 𝛍𝛎{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}} implies c𝛍c𝛎c_{\boldsymbol{\mu}}\geq c_{{\boldsymbol{\nu}}} and (ii) if 𝛎{\boldsymbol{\nu}}\in\mathcal{F} and for some i<ji<j it holds νi=νj=0\nu_{i}=\nu_{j}=0, then c𝛎+𝐞ic𝛎+𝐞jc_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{i}}\geq c_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{j}}.

Then for any ε>0\varepsilon>0, Algorithm 1 terminates and returns Λε\Lambda_{\varepsilon} in (6.2). Moreover each line of Algorithm 1 is executed at most 4|Λε|+14|\Lambda_{\varepsilon}|+1 times.

6.3 Interpolation convergence rate

If XX is a Hilbert space, then the Wiener-Hermite PC expansion of u:UXu:U\to X converges in general only in L2(U,X;γ)L^{2}(U,X;\gamma). As mentioned before this creates some subtleties when working with interpolation and quadrature operators based on pointwise evaluations of the target function. To demonstrate this, we recall the following example from [40], which does not satisfy (𝐛,ξ,δ,)({\boldsymbol{b}},\xi,\delta,\mathbb{C})-holomorphy, since Definition 4.1 (iii) does not hold.

Example 6.9.

Define u:Uu:U\to\mathbb{C} pointwise by

u(𝒚):={1if |{j:yj0}|<0otherwise.u({\boldsymbol{y}}):=\begin{cases}1&\text{if }|\{j\in\mathbb{N}\,:\,y_{j}\neq 0\}|<\infty\\ 0&\text{otherwise.}\end{cases}

Then uu vanishes on the complement of the γ\gamma-null set

nn×{0}.\bigcup_{n\in\mathbb{N}}\mathbb{R}^{n}\times\{0\}^{\infty}.

Consequently uu is equal to the constant zero function in the sense of L2(U;γ)L^{2}(U;\gamma). Hence there holds the expansion u=𝛎0H𝛎u=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}0\cdot H_{\boldsymbol{\nu}} with convergence in L2(U;γ)L^{2}(U;\gamma). Now let Λ\Lambda\subseteq\mathcal{F} be nonempty, finite and downward closed. As explained in Section 6.1.1, the interpolation operator 𝐈Λ\mathbf{I}_{\Lambda} reproduces all polynomials in span{𝐲𝛎:𝛎Λ}{\rm span}\{{\boldsymbol{y}}^{\boldsymbol{\nu}}\,:\,{\boldsymbol{\nu}}\in\Lambda\}. Since any point (χνj,μj)j(\chi_{\nu_{j},\mu_{j}})_{j\in\mathbb{N}} with μjνj\mu_{j}\leq\nu_{j} is zero in all but finitely many coordinates (due to χ0,0=0\chi_{0,0}=0), we observe that

𝐈Λu10𝝂0𝐈ΛH𝝂.\mathbf{I}_{\Lambda}u\equiv 1\neq 0\equiv\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}0\cdot\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}}.

This is due to the fact that u=𝛎0H𝛎u=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}0\cdot H_{\boldsymbol{\nu}} only holds in the L2(U;γ)L^{2}(U;\gamma) sense, and interpolation or quadrature (which require pointwise evaluation of the function) are not meaningful for L2(U;γ)L^{2}(U;\gamma) functions.

The above example shows that if

u=𝝂u𝝂H𝝂L2(U;γ)u=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}u_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}}\in L^{2}(U;\gamma)

with Wiener-Hermite PC expansion coefficients (u𝝂)ν(u_{\boldsymbol{\nu}})_{\nu\in\mathcal{F}}\subset\mathbb{R}, then the formal equalities

𝐈Λu=𝝂u𝝂𝐈ΛH𝝂,\mathbf{I}_{\Lambda}u=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}u_{\boldsymbol{\nu}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}},

and

𝐐Λu=𝝂u𝝂𝐐ΛH𝝂\mathbf{Q}_{\Lambda}u=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}u_{\boldsymbol{\nu}}\mathbf{Q}_{\Lambda}H_{\boldsymbol{\nu}}

do in general not hold in L2(U;γ)L^{2}(U;\gamma). Our definition of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy allows to circumvent this by interpolating not uu itself but the approximations uNu_{N} to uu which are pointwise defined and only depend on finitely many variables, cp. Definition 4.1.

Our analysis starts with the following result about pointwise convergence. For k{1,2}k\in\{1,2\} and NN\in\mathbb{N} we introduce the notation

kN:={𝝂k:supp(𝝂){1,,N}}.\mathcal{F}_{k}^{N}:=\{{\boldsymbol{\nu}}\in\mathcal{F}_{k}\,:\,\operatorname{supp}({\boldsymbol{\nu}})\subseteq\{1,\dots,N\}\}.

These sets thus contain multiindices 𝝂{\boldsymbol{\nu}} for which νj=0\nu_{j}=0 for all j>Nj>N.

Lemma 6.10.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛(0,){\boldsymbol{b}}\in(0,\infty)^{\infty}. Let NN\in\mathbb{N}, and let u~N:UX\tilde{u}_{N}:U\to X be as in Definition 4.1. For 𝛎{\boldsymbol{\nu}}\in\mathcal{F} define

u~N,𝝂:=Uu~N(𝒚)H𝝂(𝒚)dγ(𝒚).\tilde{u}_{N,{\boldsymbol{\nu}}}:=\int_{U}\tilde{u}_{N}({\boldsymbol{y}})H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}).

Then,

u~N(𝒚)=𝝂1Nu~N,𝝂H𝝂(𝒚)\tilde{u}_{N}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N,{\boldsymbol{\nu}}}H_{\boldsymbol{\nu}}({\boldsymbol{y}}) (6.20)

with the equality and pointwise absolute convergence in XX for all 𝐲U{\boldsymbol{y}}\in U.

Proof.

From the Cramér bound

|H~n(x)|<2n/2n!exp(x2/2),|\tilde{H}_{n}(x)|<2^{n/2}\sqrt{n!}\exp(x^{2}/2),

see [74], and where H~n(x/2):=2n/2n!Hn(x)\tilde{H}_{n}(x/\sqrt{2}):=2^{n/2}\sqrt{n!}H_{n}(x), see [1, Page 787], we have for all n0n\in\mathbb{N}_{0}

supxexp(x2/4)|Hn(x)|1.\sup_{x\in\mathbb{R}}\exp(-x^{2}/4)|H_{n}(x)|\leq 1. (6.21)

By Theorem. 4.8 (u~N,𝝂)ν1()(\tilde{u}_{N,{\boldsymbol{\nu}}})_{\nu\in\mathcal{F}}\in\ell^{1}(\mathcal{F}). Note that for 𝝂1N{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}

u~N,𝝂=Uu~N(𝒚)H𝝂(𝒚)dγ(𝒚)=NuN(y1,,yN)j=1NHνj(yj)dγN((yj)j=1N)\tilde{u}_{N,{\boldsymbol{\nu}}}=\int_{U}\tilde{u}_{N}({\boldsymbol{y}})H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})=\int_{\mathbb{R}^{N}}u_{N}(y_{1},\dots,y_{N})\prod_{j=1}^{N}H_{\nu_{j}}(y_{j})\,\mathrm{d}\gamma_{N}((y_{j})_{j=1}^{N})

and thus u~N,𝝂\tilde{u}_{N,{\boldsymbol{\nu}}} coincides with the Wiener-Hermite PC expansion coefficient of uNu_{N} w.r.t. the multiindex (νj)j=1N0N(\nu_{j})_{j=1}^{N}\in\mathbb{N}_{0}^{N}. The summability of the collection

(uN,𝝂Xj=1NHνj(yj)L2(N;γN))𝝂1N\left(\|u_{N,{\boldsymbol{\nu}}}\|_{X}\|\prod_{j=1}^{N}H_{\nu_{j}}(y_{j})\|_{L^{2}(\mathbb{R}^{N};\gamma_{N})}\right)_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}

now implies in particular,

uN((yj)j=1N)=𝝂1NuN,𝝂j=1NHνj(yj)u_{N}((y_{j})_{j=1}^{N})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}u_{N,{\boldsymbol{\nu}}}\prod_{j=1}^{N}H_{\nu_{j}}(y_{j})

in the sense of L2(N;γN)L^{2}(\mathbb{R}^{N};\gamma_{N}).

Due to (6.21) and (uN,𝝂X)𝝂1N1(1N)(\|u_{N,{\boldsymbol{\nu}}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\in\ell^{1}(\mathcal{F}_{1}^{N}) we can define a continuous function

u^N:(yj)j=1N𝝂0NuN,𝝂j=1NHνj(yj)\hat{u}_{N}:(y_{j})_{j=1}^{N}\mapsto\sum_{{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{N}}u_{N,{\boldsymbol{\nu}}}\prod_{j=1}^{N}H_{\nu_{j}}(y_{j}) (6.22)

on N\mathbb{R}^{N}. By (6.21), for every fixed (yj)j=1NN(y_{j})_{j=1}^{N}\in\mathbb{R}^{N} we have the uniform bound |j=1NHνj(yj)|j=1Nexp(yj24)|\prod_{j=1}^{N}H_{\nu_{j}}(y_{j})|\leq\prod_{j=1}^{N}\exp(\frac{y_{j}^{2}}{4}) independent of 𝝂1N{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}. The summability of (uN,𝝂X)𝝂1N(\|u_{N,{\boldsymbol{\nu}}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}} implies the absolute convergence of the series in (6.22) for every fixed (yj)j=1NN(y_{j})_{j=1}^{N}\in\mathbb{R}^{N}.

Since they have the same Wiener-Hermite PC expansion, it holds u^N=uN\hat{u}_{N}=u_{N} in the sense of L2(N;γN)L^{2}(\mathbb{R}^{N};\gamma_{N}).

By Definition 4.1 the function u:NXu:\mathbb{R}^{N}\to X is in particular continuous (it even allows a holomorphic extension to some subset of N{\mathbb{C}}^{N} containing N\mathbb{R}^{N}). Now u^N\hat{u}_{N}, uN:NXu_{N}:\mathbb{R}^{N}\to X are two continuous functions which are equal in the sense of L2(N;γN)L^{2}(\mathbb{R}^{N};\gamma_{N}). Thus they coincide pointwise and it holds in XX for every 𝒚U{\boldsymbol{y}}\in U,

u~N(𝒚)=uN((yj)j=1N)=𝝂1Nu~N,𝝂H𝝂(𝒚).\tilde{u}_{N}({\boldsymbol{y}})=u_{N}((y_{j})_{j=1}^{N})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N,{\boldsymbol{\nu}}}H_{\boldsymbol{\nu}}({\boldsymbol{y}}).

The result on the pointwise absolute convergence in Lemma 6.10 is not sufficient for establishing the convergence rate of the interpolation approximation in the space L2(U,X;γ)L^{2}(U,X;\gamma). To this end, we need the result on convergence in the space L2(U,X;γ)L^{2}(U,X;\gamma) in the following lemma.

Lemma 6.11.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛(0,){\boldsymbol{b}}\in(0,\infty)^{\infty}. Let NN\in\mathbb{N}, and let u~N:UX\tilde{u}_{N}:U\to X be as in Definition 4.1 and u~N,𝛎\tilde{u}_{N,{\boldsymbol{\nu}}} as in Lemma 6.10. Let Λ1\Lambda\subset\mathcal{F}_{1} be a finite, downward closed set.

Then we have

𝐈Λu~N=𝝂1Nu~N;𝝂𝐈ΛH𝝂\displaystyle\mathbf{I}_{\Lambda}\tilde{u}_{N}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}} (6.23)

with the equality and unconditional convergence in the space L2(U,X;γ)L^{2}(U,X;\gamma).

Proof.

For a function v:UXv:U\to X we have

𝐈Λv(𝒚)=𝝂ΛσΛ;𝝂μ,𝝁𝝂v(χ𝝂,𝝁)L𝝂,𝝁(𝒚),\mathbf{I}_{\Lambda}v({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\Lambda}\sigma_{\Lambda;{\boldsymbol{\nu}}}\sum_{\mu\in{\mathcal{F}},{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}v(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}({\boldsymbol{y}}), (6.24)

where σΛ;𝝂\sigma_{\Lambda;{\boldsymbol{\nu}}} is defined in (6.5) and recall, χ𝝂,𝝁=(χνj,μj)j\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}=(\chi_{\nu_{j},\mu_{j}})_{j\in{\mathbb{N}}} and

L𝝂,𝝁(𝒚):=ji=0iμjνjyjχνj,iχνj,μjχνj,i,𝒚U.L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}({\boldsymbol{y}}):=\prod_{j\in\mathbb{N}}\prod_{\begin{subarray}{c}i=0\\ i\neq\mu_{j}\end{subarray}}^{\nu_{j}}\frac{y_{j}-\chi_{\nu_{j},i}}{\chi_{\nu_{j},\mu_{j}}-\chi_{\nu_{j},i}},\quad{\boldsymbol{y}}\in U. (6.25)

Since in a Banach space the absolute convergence implies the unconditional convergence, from Lemma 6.10 it follows that for any 𝒚U{\boldsymbol{y}}\in U,

u~N(𝒚)=𝝂1Nu~N,𝝂H𝝂(𝒚)\tilde{u}_{N}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N,{\boldsymbol{\nu}}}H_{\boldsymbol{\nu}}({\boldsymbol{y}}) (6.26)

with the equality and unconditional convergence in XX. Let {Fn}n1N\{F_{n}\}_{n\in{\mathbb{N}}}\subset{\mathcal{F}}_{1}^{N} be any sequence of finite sets in 1N{\mathcal{F}}_{1}^{N} exhausting 1N{\mathcal{F}}_{1}^{N}. Then

𝒚U:u~N(n)(𝒚):=𝝂Fnu~N,𝝂H𝝂(𝒚)u~N(𝒚),n,\forall{\boldsymbol{y}}\in U:\ \ \tilde{u}_{N}^{(n)}({\boldsymbol{y}}):=\sum_{{\boldsymbol{\nu}}\in F_{n}}\tilde{u}_{N,{\boldsymbol{\nu}}}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\ \to\ \tilde{u}_{N}({\boldsymbol{y}}),\ \ n\to\infty, (6.27)

with the sequence convergence in the space XX. Notice that the functions 𝐈Λu~N\mathbf{I}_{\Lambda}\tilde{u}_{N} and 𝝂Fnu~N,𝝂𝐈ΛH𝝂\sum_{{\boldsymbol{\nu}}\in F_{n}}\tilde{u}_{N,{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}} belong to the space L2(U,X;γ)L^{2}(U,X;\gamma). Hence we have that

𝐈Λu~N𝝂Fnu~N,𝝂𝐈ΛH𝝂L2(U,X;γ)=𝐈Λu~N𝐈Λu~N(n)L2(U,X;γ)=𝐈Λ(u~Nu~N(n))L2(U,X;γ)𝝂Λ|σΛ;𝝂|𝝁,𝝁𝝂u~N(χ𝝂,𝝁)u~N(n)(χ𝝂,𝝁)XU|L𝝂,𝝁(𝒚)|dγ(𝒚).\begin{split}&\bigg{\|}\mathbf{I}_{\Lambda}\tilde{u}_{N}-\sum_{{\boldsymbol{\nu}}\in F_{n}}\tilde{u}_{N,{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}}\bigg{\|}_{L^{2}(U,X;\gamma)}=\big{\|}\mathbf{I}_{\Lambda}\tilde{u}_{N}-\mathbf{I}_{\Lambda}\tilde{u}_{N}^{(n)}\|_{L^{2}(U,X;\gamma)}=\big{\|}\mathbf{I}_{\Lambda}\big{(}\tilde{u}_{N}-\tilde{u}_{N}^{(n)}\big{)}\|_{L^{2}(U,X;\gamma)}\\ &\leq\sum_{{\boldsymbol{\nu}}\in\Lambda}|\sigma_{\Lambda;{\boldsymbol{\nu}}}|\sum_{{\boldsymbol{\mu}}\in{\mathcal{F}},{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\big{\|}\tilde{u}_{N}(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})-\tilde{u}_{N}^{(n)}(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})\big{\|}_{X}\int_{U}|L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}({\boldsymbol{y}})|\,\mathrm{d}\gamma({\boldsymbol{y}}).\end{split} (6.28)

Observe that L𝝂,𝝁L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}} is a polynomial of order |𝝂||{\boldsymbol{\nu}}|. Since {𝝁:𝝁𝝂}\{{\boldsymbol{\mu}}\in{\mathcal{F}}:{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}\} and Λ\Lambda are finite sets, we can choose C:=C(Λ)>0C:=C(\Lambda)>0 so that

U|L𝝂,𝝁(𝒚)|dγ(𝒚)C\int_{U}|L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}({\boldsymbol{y}})|\,\mathrm{d}\gamma({\boldsymbol{y}})\leq C

for all 𝝁𝝂{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}} and 𝝂Λ{\boldsymbol{\nu}}\in\Lambda, and, moreover, by using (6.27) we can choose n0n_{0} so that

u~N(χ𝝂,𝝁)u~N(n)(χ𝝂,𝝁)Xε\|\tilde{u}_{N}(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})-\tilde{u}_{N}^{(n)}(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})\|_{X}\leq\varepsilon

for all nn0n\geq n_{0} and 𝝁𝝂{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}, 𝝂Λ{\boldsymbol{\nu}}\in\Lambda. Consequently, we have that for all nn0n\geq n_{0},

𝐈Λu~N𝝂Fnu~N,𝝂𝐈ΛH𝝂L2(U,X;γ)C𝝂Λ|σΛ;𝝂|𝝁𝝂ε=Cε𝝂Λ|σΛ;𝝂|p𝝂(1).\begin{split}\bigg{\|}\mathbf{I}_{\Lambda}\tilde{u}_{N}-\sum_{{\boldsymbol{\nu}}\in F_{n}}\tilde{u}_{N,{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}}\bigg{\|}_{L^{2}(U,X;\gamma)}\leq C\sum_{{\boldsymbol{\nu}}\in\Lambda}|\sigma_{\Lambda;{\boldsymbol{\nu}}}|\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\varepsilon=C\varepsilon\sum_{{\boldsymbol{\nu}}\in\Lambda}|\sigma_{\Lambda;{\boldsymbol{\nu}}}|p_{\boldsymbol{\nu}}(1).\end{split} (6.29)

Hence we derive the convergence in the space L2(U,X;γ)L^{2}(U,X;\gamma) of the sequence 𝝂Fnu~N,𝝂𝐈ΛH𝝂\sum_{{\boldsymbol{\nu}}\in F_{n}}\tilde{u}_{N,{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}} to 𝐈Λu~N\mathbf{I}_{\Lambda}\tilde{u}_{N} (nn\to\infty) for any sequence of finite sets {Fn}n1N\{F_{n}\}_{n\in{\mathbb{N}}}\subset{\mathcal{F}}_{1}^{N} exhausting 1N{\mathcal{F}}_{1}^{N}. This proves the lemma. ∎

Remark 6.12.

Under the assumption of Lemma 6.11, in a similar way, we can prove that for every 𝒚U{\boldsymbol{y}}\in U

𝐈Λu~N(𝒚)=𝝂1Nu~N;𝝂𝐈ΛH𝝂(𝒚)\mathbf{I}_{\Lambda}\tilde{u}_{N}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda}H_{\boldsymbol{\nu}}({\boldsymbol{y}}) (6.30)

with the equality and unconditional convergence in the space XX.

We arrive at the following convergence rate result, which improves the convergence rate in [52] (in terms of the number of function evaluations) by a factor 22 (for the case when the elements of the representation system are supported globally in D{D}). Additionally, we provide an explicit construction of suitable index sets. Recall that pointwise evaluations of a (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions are understood in the sense of Remark. 4.4.

Theorem 6.13.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) and some p(0,2/3)p\in(0,2/3). Let (c1,𝛎)𝛎(c_{1,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} be as in Lemma 6.5 with ϱ{\boldsymbol{\varrho}} as in Theorem 4.8.

Then there exist C>0C>0 and, for every nn\in{\mathbb{N}}, εn>0\varepsilon_{n}>0 such that |pts(Λ1,εn)|n|{\rm pts}(\Lambda_{1,\varepsilon_{n}})|\leq n (with Λ1,εn\Lambda_{1,\varepsilon_{n}} as in (6.17)) and

u𝐈Λ1,εnuL2(U,X;γ)Cn1p+32.\|u-\mathbf{I}_{\Lambda_{1,\varepsilon_{n}}}u\|_{L^{2}(U,X;\gamma)}\ \leq\ Cn^{{-\frac{1}{p}+\frac{3}{2}}}.
Proof.

For ε>0\varepsilon>0 small enough and satisfying |Λ1,ε|>0|\Lambda_{1,\varepsilon}|>0, take NN\in\mathbb{N} with

Nmax{jsupp(𝝂):𝝂Λ1,ε},N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,{\boldsymbol{\nu}}\in\Lambda_{1,\varepsilon}\},

so large that

uu~NL2(U,X;γ)ε12p4(1p),\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}\leq\varepsilon^{\frac{1}{2}-\frac{p}{4(1-p)}}, (6.31)

which is possible due to the (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy of uu (cp. Definition 4.1 (iii)). An appropriate value of ε\varepsilon depending on nn will be chosen below. In the following for 𝝂1N{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N} we denote by u~N,𝝂X\tilde{u}_{N,{\boldsymbol{\nu}}}\in X the PC coefficient of u~N\tilde{u}_{N} and for 𝝂{\boldsymbol{\nu}}\in\mathcal{F} as earlier u𝝂Xu_{\boldsymbol{\nu}}\in X is the PC coefficient of uu.

Because

Nmax{jsupp(𝝂):𝝂Λ1,ε}N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,{\boldsymbol{\nu}}\in\Lambda_{1,\varepsilon}\}

and χ0,0=0\chi_{0,0}=0, we have

𝐈Λ1,εu=𝐈Λ1,εu~N\mathbf{I}_{\Lambda_{1,\varepsilon}}u=\mathbf{I}_{\Lambda_{1,\varepsilon}}\tilde{u}_{N}

(cp. Remark 4.4). Hence by (6.31)

u𝐈Λ1,εuL2(U,X;γ)=u𝐈Λ1,εu~NL2(U,X;γ)ε12p4(1p)+u~N𝐈Λ1,εu~NL2(U,X;γ).\|u-\mathbf{I}_{\Lambda_{1,\varepsilon}}u\|_{L^{2}(U,X;\gamma)}=\|u-\mathbf{I}_{\Lambda_{1,\varepsilon}}\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}\leq\varepsilon^{\frac{1}{2}-\frac{p}{4(1-p)}}+\|\tilde{u}_{N}-\mathbf{I}_{\Lambda_{1,\varepsilon}}\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}. (6.32)

We now give a bound of the second term on the right side of (6.32). By Lemma 6.11 we can write

𝐈Λ1,εu~N=𝝂1Nu~N;𝝂𝐈Λ1,εH𝝂\displaystyle\mathbf{I}_{\Lambda_{1,\varepsilon}}\tilde{u}_{N}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{I}_{\Lambda_{1,\varepsilon}}H_{\boldsymbol{\nu}}

with the equality and unconditional in L2(U,X;γ)L^{2}(U,X;\gamma). Hence by Lemma 6.2 and (6.7) we have that

u~N𝐈Λ1,εu~NL2(U,X;γ)\displaystyle\|\tilde{u}_{N}-\mathbf{I}_{\Lambda_{1,\varepsilon}}\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)} =𝝂\Λ1,εu~N;𝝂(H𝝂𝐈Λ1,εH𝝂)L2(U,X;γ)\displaystyle=\left\|\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{1,\varepsilon}}\tilde{u}_{N;{\boldsymbol{\nu}}}(H_{\boldsymbol{\nu}}-\mathbf{I}_{\Lambda_{1,\varepsilon}}H_{\boldsymbol{\nu}})\right\|_{L^{2}(U,X;\gamma)}
𝝂\Λ1,εu~N;𝝂X(H𝝂L2(U;γ)+𝐈Λ1,εH𝝂L2(U;γ))\displaystyle\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{1,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}\big{(}\|H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}+\|\mathbf{I}_{\Lambda_{1,\varepsilon}}H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}\big{)}
𝝂1N\Λ1,εu~N;𝝂X(1+p𝝂(3))\displaystyle\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}\backslash\Lambda_{1,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}\left(1+p_{{\boldsymbol{\nu}}}(3)\right)
2𝝂1N\Λ1,εu~N;𝝂Xp𝝂(3).\displaystyle\leq 2\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}\backslash\Lambda_{1,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}p_{{\boldsymbol{\nu}}}(3).

Choosing r>4/p1r>4/p-1 (q:=p/(1p)q:=p/(1-p), τ=3\tau=3), according to Proposition 6.7, (6.14) and Theorem 4.8 (with (ϱj1)jp/(1p)()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{p/(1-p)}(\mathbb{N}) as in Theorem 4.8) the last sum is bounded by

C(𝝂1Nβ𝝂(r,ϱ)u~N,𝝂X2)ε12q4C(𝒃)δ2ε12q4=C(𝒃)δ2ε12p4(1p),C\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|\tilde{u}_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\right)\varepsilon^{\frac{1}{2}-\frac{q}{4}}\leq C({\boldsymbol{b}})\delta^{2}\varepsilon^{\frac{1}{2}-\frac{q}{4}}=C({\boldsymbol{b}})\delta^{2}\varepsilon^{\frac{1}{2}-\frac{p}{4(1-p)}},

and the constant C(𝒃)C({\boldsymbol{b}}) from Theorem 4.8 does not depend on NN and δ\delta. Hence, by (6.32) we obtain

u𝐈Λ1,εuL2(U,X;γ)C1ε12p4(1p).\|u-\mathbf{I}_{\Lambda_{1,\varepsilon}}u\|_{L^{2}(U,X;\gamma)}\leq C_{1}\varepsilon^{\frac{1}{2}-\frac{p}{4(1-p)}}. (6.33)

From (6.19) it follows that

|pts(Λ1,ε)|C2εq2=C2εp2(1p).|{\rm pts}(\Lambda_{1,\varepsilon})|\leq C_{2}\varepsilon^{-\frac{q}{2}}=C_{2}\varepsilon^{-\frac{p}{2(1-p)}}.

For every nn\in{\mathbb{N}}, we choose an εn>0\varepsilon_{n}>0 satisfying the condition

n/2C2εnp2(1p)n.n/2\leq C_{2}\varepsilon_{n}^{-\frac{p}{2(1-p)}}\leq n.

Then due to (6.33), the claim holds true for the chosen εn\varepsilon_{n}. ∎

Remark 6.14.

Comparing the best nn-term convergence result in Remark 4.10 with the interpolation result of Theorem 6.13, we observe that the convergence rate is reduced by 1/21/2, and moreover, rather than p(0,1)p\in(0,1) as in Remark 4.10, Theorem 6.13 requires p(0,2/3)p\in(0,2/3). This discrepancy can be explained as follows: Since (H𝝂)𝝂(H_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} forms an orthonormal basis of L2(U;γ)L^{2}(U;\gamma), for the best nn-term result we could resort to Parseval’s identity, which merely requires 2\ell^{2}-summability of the Hermite PC coefficients, i.e. (u𝝂X)𝝂2()(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{2}(\mathcal{F}). Due to (u𝝂X)𝝂2p2p(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{\frac{2p}{2-p}} by Theorem 4.9, this is ensured as long as p(0,1)p\in(0,1). On the other hand, for the interpolation result we had to use the triangle inequality, since the family (𝐈Λ1,εnH𝝂)𝝂(\mathbf{I}_{\Lambda_{1,\varepsilon_{n}}}H_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} of interpolated multivariate Hermite polynomials does not form an orthonormal family of L2(U;γ)L^{2}(U;\gamma). This argument requires the stronger condition (u𝝂X)𝝂1()(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{1}(\mathcal{F}), resulting in the stronger assumption p(0,2/3)p\in(0,2/3) of Theorem 6.13.

6.4 Quadrature convergence rate

We first prove a result on equality and unconditional convergence in the space XX for quadrature operators, which is similar to that in Lemma 6.11. It is needed to establish the quadrature convergence rate.

Lemma 6.15.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛(0,){\boldsymbol{b}}\in(0,\infty)^{\infty}. Let NN\in\mathbb{N}, and let u~N:UX\tilde{u}_{N}:U\to X be as in Definition 4.1 and u~N,𝛎\tilde{u}_{N,{\boldsymbol{\nu}}} as in Lemma 6.10. Let Λ1\Lambda\subset\mathcal{F}_{1} be a finite downward closed set.

Then we have

𝐐Λu~N=𝝂1Nu~N;𝝂𝐐ΛH𝝂\mathbf{Q}_{\Lambda}\tilde{u}_{N}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{Q}_{\Lambda}H_{\boldsymbol{\nu}} (6.34)

with the equality and unconditional convergence in the space XX.

Proof.

For a function v:UXv:U\to X by (6.8) and (6.24) we have

𝐐Λv=𝝂ΛσΛ;𝝂μ,𝝁𝝂v(χ𝝂,𝝁)UL𝝂,𝝁(𝒚)dγ(𝒚),\mathbf{Q}_{\Lambda}v=\sum_{{\boldsymbol{\nu}}\in\Lambda}\sigma_{\Lambda;{\boldsymbol{\nu}}}\sum_{\mu\in{\mathcal{F}},{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}v(\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}})\int_{U}L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}),

where χ𝝂,𝝁=(χνj,μj)j\chi_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}}=(\chi_{\nu_{j},\mu_{j}})_{j\in{\mathbb{N}}}, σΛ;𝝂\sigma_{\Lambda;{\boldsymbol{\nu}}}, L𝝂,𝝁L_{{\boldsymbol{\nu}},{\boldsymbol{\mu}}} are defined in (6.5) and (6.25), respectively. By using this representation, we can prove the lemma in a way similar to the proof of Lemma 6.11 with some appropriate modifications. ∎

Analogous to Theorem 6.13 we obtain the following result for the quadrature convergence with an improved convergence rate compared to interpolation.

Theorem 6.16.

Let uu be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) and some p(0,4/5)p\in(0,4/5). Let (c2,𝛎)𝛎(c_{2,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} be as in Lemma 6.5 with ϱ\varrho as in Theorem 4.8. Then there exist C>0C>0 and, for every NN\in\mathbb{N}, εn>0\varepsilon_{n}>0 such that |pts(Λ2,εn)|n|{\rm pts}(\Lambda_{2,\varepsilon_{n}})|\leq n (with Λ2,εn\Lambda_{2,\varepsilon_{n}} as in (6.17)) and

Uu(𝒚)dγ(𝒚)𝐐Λ2,εnuXCn2p+52.\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon_{n}}}u\right\|_{X}\ \leq\ Cn^{-\frac{2}{p}+\frac{5}{2}}.
Proof.

For ε>0\varepsilon>0 small enough and satisfying |Λ2,ε|>0|\Lambda_{2,\varepsilon}|>0, take NN\in\mathbb{N}, Nmax{jsupp(𝝂):𝝂Λ2,ε}N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,{\boldsymbol{\nu}}\in\Lambda_{2,\varepsilon}\} so large that

U[u(𝒚)u~N(𝒚)]dγ(𝒚)Xuu~NL2(U,X;γ)ε12p8(1p),\left\|\int_{U}\big{[}u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}})\big{]}\,\mathrm{d}\gamma({\boldsymbol{y}})\right\|_{X}\leq\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}\leq\varepsilon^{\frac{1}{2}-\frac{p}{8(1-p)}}, (6.35)

which is possible due to the (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphy of uu (cp. Definition 4.1 (iii)). An appropriate value of ε\varepsilon depending on nn will be chosen below. In the following for 𝝂{\boldsymbol{\nu}}\in\mathcal{F} we denote by u~N,𝝂\tilde{u}_{N,{\boldsymbol{\nu}}} the Wiener-Hermite PC expansion coefficient of u~N\tilde{u}_{N} and as earlier u𝝂u_{\boldsymbol{\nu}} is the Wiener-Hermite PC expansion coefficient of uu.

Because

Nmax{jsupp(𝝂):𝝂Λ2,ε}N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,{\boldsymbol{\nu}}\in\Lambda_{2,\varepsilon}\}

and χ0,0=0\chi_{0,0}=0, we have 𝐐Λ2,εu=𝐐Λ2,εu~N\mathbf{Q}_{\Lambda_{2,\varepsilon}}u=\mathbf{Q}_{\Lambda_{2,\varepsilon}}\tilde{u}_{N} (cp. Remark. 4.4). Hence by (6.35)

Uu(𝒚)dγ(𝒚)𝐐Λ2,εuX\displaystyle\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}u\right\|_{X} =Uu(𝒚)dγ(𝒚)𝐐Λ2,εu~NX\displaystyle=\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}\tilde{u}_{N}\right\|_{X}
ε12p8(1p)+Uu~N(𝒚)dγ(𝒚)𝐐Λ2,εu~NX.\displaystyle\leq\varepsilon^{\frac{1}{2}-\frac{p}{8(1-p)}}+\left\|\int_{U}\tilde{u}_{N}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}\tilde{u}_{N}\right\|_{X}. (6.36)

By Lemma 6.15 we have

𝐐Λ2,εu~N=𝝂1Nu~N;𝝂𝐐Λ2,εH𝝂=𝝂2Nu~N;𝝂𝐐Λ2,εH𝝂\mathbf{Q}_{\Lambda_{2,\varepsilon}}\tilde{u}_{N}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{Q}_{\Lambda_{2,\varepsilon}}H_{\boldsymbol{\nu}}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}^{N}}\tilde{u}_{N;{\boldsymbol{\nu}}}\mathbf{Q}_{\Lambda_{2,\varepsilon}}H_{\boldsymbol{\nu}}

with the equality and unconditional convergence in the space XX. Since Λ2,ε\Lambda_{2,\varepsilon} is nonempty and downward closed we have 𝟎Λ2,ε{\boldsymbol{0}}\in\Lambda_{2,\varepsilon}. Then, by Lemma 6.4, (6.10), and using

UH𝝂(𝒚)dγ(𝒚)=0\int_{U}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})=0

for all 𝟎𝝂\2{\boldsymbol{0}}\neq{\boldsymbol{\nu}}\in\mathcal{F}\backslash\mathcal{F}_{2}, we have that

Uu~N(𝒚)dγ(𝒚)𝐐Λ2,εu~NX\displaystyle\left\|\int_{U}\tilde{u}_{N}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}\tilde{u}_{N}\right\|_{X} =𝝂2\Λ2,εu~N;𝝂(UH𝝂(𝒚)dγ(𝒚)𝐐Λ2,εH𝝂)X\displaystyle=\left\|\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}\backslash\Lambda_{2,\varepsilon}}\tilde{u}_{N;{\boldsymbol{\nu}}}\left(\int_{U}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}H_{\boldsymbol{\nu}}\right)\right\|_{X}
𝝂2\Λ2,εu~N;𝝂X(H𝝂L2(U;γ)+|𝐐Λ2,εH𝝂|)\displaystyle\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}\backslash\Lambda_{2,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}(\|H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}+|\mathbf{Q}_{\Lambda_{2,\varepsilon}}H_{\boldsymbol{\nu}}|)
𝝂2\Λ2,εu~N;𝝂X(1+p𝝂(3))\displaystyle\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}\backslash\Lambda_{2,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}\left(1+p_{\boldsymbol{\nu}}(3)\right)
2𝝂2\Λ2,εu~N;𝝂Xp𝝂(3).\displaystyle\leq 2\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}\backslash\Lambda_{2,\varepsilon}}\|\tilde{u}_{N;{\boldsymbol{\nu}}}\|_{X}p_{\boldsymbol{\nu}}(3).

Choosing r>8/p5r>8/p-5 (q=p1pq=\frac{p}{1-p}, τ=3\tau=3), according to Proposition 6.7, (6.14) and Theorem 4.8 (with (ϱj1)jp/(1p)()(\varrho_{j}^{-1})_{j\in\mathbb{N}}\in\ell^{p/(1-p)}(\mathbb{N}) as in Theorem 4.8) the last sum is bounded by

C(𝝂β𝝂(r,ϱ)u~N,𝝂X2)ε12q8C(𝒃)δ2ε12q8=C(𝒃)ε12p8(1p),C\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|\tilde{u}_{N,{\boldsymbol{\nu}}}\|_{X}^{2}\right)\varepsilon^{\frac{1}{2}-\frac{q}{8}}\leq C({\boldsymbol{b}})\delta^{2}\varepsilon^{\frac{1}{2}-\frac{q}{8}}=C({\boldsymbol{b}})\varepsilon^{\frac{1}{2}-\frac{p}{8(1-p)}},

and the constant C(𝒃)C({\boldsymbol{b}}) from Theorem 4.8 does not depend on NN and δ\delta. Hence, by (6.35) and (6.4) we obtain that

Uu(𝒚)dγ(𝒚)𝐐Λ2,εuXC1ε12p8(1p).\displaystyle\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Lambda_{2,\varepsilon}}u\right\|_{X}\leq C_{1}\varepsilon^{\frac{1}{2}-\frac{p}{8(1-p)}}. (6.37)

From (6.19) it follows that

|pts(Λk,ϵ)|C2εq4=C2εp4(1p).|{\rm pts}(\Lambda_{k,\epsilon})|\leq C_{2}\varepsilon^{-\frac{q}{4}}=C_{2}\varepsilon^{-\frac{p}{4(1-p)}}.

For every nn\in{\mathbb{N}}, we choose an εn>0\varepsilon_{n}>0 satisfying the condition

n/2C2εnp4(1p)n.n/2\leq C_{2}\varepsilon_{n}^{-\frac{p}{4(1-p)}}\leq n.

Then due to (6.37) the claim holds true for the chosen εn\varepsilon_{n}. ∎

Remark 6.17.

Interpolation formulas based on index sets like

Λ(ξ):={𝝂:β𝝂(r,ϱ)ξ2/q},\Lambda(\xi):=\{{\boldsymbol{\nu}}\in\mathcal{F}:\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\leq\xi^{2/q}\},

(where ξ>0\xi>0 is a large parameter), have been proposed in [52, 43] for the parametric, elliptic divergence-form PDE (3.17) with log-Gaussian inputs (3.18) satisfying the assumptions of Theorem 3.38 with i=1i=1. There, dimension-independent convergence rates of sparse-grid interpolation were obtained. Based on the weighted 2\ell^{2}-summability of the Wiener-Hermite PC expansion coefficients of the form

𝝂β𝝂(r,ϱ)u𝝂X2<with(p𝝂(τ,λ)β𝝂(r,ϱ)1/2)𝝂q()(0<q<2),\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})\|u_{\boldsymbol{\nu}}\|_{X}^{2}<\infty\ \ \ \text{with}\ \ \ \big{(}p_{\boldsymbol{\nu}}(\tau,\lambda)\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{-1/2}\big{)}_{{\boldsymbol{\nu}}\in{\mathcal{F}}}\in\ell^{q}({\mathcal{F}})\ \ (0<q<2), (6.38)

the rate established in [52] is 12(1/q1/2)\frac{1}{2}(1/q-1/2) which lower than those obtained in the present analysis. The improved rate 1/q1/21/q-1/2 has been established in [44]. This rate coincides with the rate in Theorem 6.13 for the choice q=p/(1p)q=p/(1-p).

The existence of Smolyak type quadratures with a proof of dimension-independent convergence rates was shown first in [31] and then in [43]. In [31], the symmetry of the GM and corresponding cancellations were not exploited, and these quadrature formulas provide the convergence rate 12(1/q1/2)\frac{1}{2}(1/q-1/2) which is lower (albeit dimension-independent) convergence rates in terms of the number of function evaluations as in Theorems 6.13 and 6.16. By using this symmetry, for a given weighted 2\ell^{2}-summability of the Wiener-Hermite PC expansion coefficients (3.43) with σ𝝂=β𝝂(r,ϱ)1/2\sigma_{\boldsymbol{\nu}}=\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{1/2}, the rate established in [43] (see also [45]) is 2/q1/22/q-1/2 which coincides with the rate of convergence that was obtained in Theorem 6.16 for the choice q=p/(1p)q=p/(1-p).

7 Multilevel Smolyak sparse-grid interpolation and quadrature

In this section we introduce a multilevel interpolation and quadrature algorithm which are suitable for numerical implementation. The presentation and arguments follow mostly [112] and [111, Section 3.2], where multilevel algorithms for the uniform measure on the hypercube [1,1][-1,1]^{\infty} were analyzed (in contrast to the case of a product GM on UU, which we consider here). In Section 7.1, we introduce the setting for the multilevel algorithms, in particular a notation of “work-measure” related to the discretization of a Wiener-Hermite PC expansion coefficient u𝝂u_{\boldsymbol{\nu}} for 𝝂{\boldsymbol{\nu}} in the set of multi-indices that are active in a given (interpolation or quadrature) approximation. Section 7.2 describes the general structure of the algorithms, Section 7.3 addresses algorithms for the determination of sets Λ\Lambda\subset{\mathcal{F}} of active multi-indices and a corresponding allocation of discretization levels in linear in |Λ||\Lambda| work and memory. Section 7.4 addresses the error analysis of the Smolyak sparse-grid interpolation, and Section 7.5 contains the error analysis of the corresponding Smolyak sparse-grid quadrature algorithm. All algorithms are formulated and analyzed in terms of several abstract hypotheses. Section 7.6 verifies these abstract conditions for a concrete family of parametric, elliptic PDEs. Finally, Section 7.7 addresses convergence rates achieveable with the mentioned Smolyak sparse-grid interpolation and quadrature algorithms assuming at hand optimal multi-index sets. The major finding being that the corresponding rates differ only by logarithmic terms from the error bounds furnished by those realized by the algorithms in Sections 7.2-7.5.

7.1 Setting and notation

To approximate the solution uu to a parametric PDE as in the examples of the preceding sections, the interpolation operator 𝐈Λ\mathbf{I}_{\Lambda} introduced in Section 6.1.1 requires function values of uu at different interpolation points in the parameter space UU. For a parameter 𝒚U{\boldsymbol{y}}\in U, typically the PDE solution u(𝒚)u({\boldsymbol{y}}), which is a function belonging to a Sobolev space over a physical domain D{{D}}, is not given in closed form and has to be approximated. The idea of multilevel approximations is to combine interpolants of approximations to uu at different spatial accuracies, in order to reduce the overall computational complexity. This will now be formalized.

First, we assume given a sequence (𝔴l)l0(\mathfrak{w}_{l})_{l\in\mathbb{N}_{0}}\subset\mathbb{N}, exhibiting the properties of the following assumption. Throughout 𝔴l\mathfrak{w}_{l} will be interpreted as a measure for the computational complexity of evaluating an approximation ul:UXu^{l}:U\to X of u:UXu:U\to X at a parameter 𝒚U{\boldsymbol{y}}\in U. Here we use a superscript ll rather than a subscript for the approximation level, as the subscript is reserved for the dimension truncated version uNu_{N} of uu as in Definition 4.1.

Assumption 7.1.

The sequence (𝔴l)l00(\mathfrak{w}_{l})_{l\in\mathbb{N}_{0}}\subseteq\mathbb{N}_{0} is strictly monotonically increasing and 𝔴0=0\mathfrak{w}_{0}=0. There exists a constant K𝔚1K_{\mathfrak{W}}\geq 1 such that for all ll\in\mathbb{N}

  1. (i)

    j=0l𝔴jK𝔚𝔴l\sum_{j=0}^{l}\mathfrak{w}_{j}\leq K_{\mathfrak{W}}\mathfrak{w}_{l},

  2. (ii)

    lK𝔚(1+log(𝔴l))l\leq K_{\mathfrak{W}}(1+\log(\mathfrak{w}_{l})),

  3. (iii)

    𝔴lK𝔚(1+𝔴l1)\mathfrak{w}_{l}\leq K_{\mathfrak{W}}(1+\mathfrak{w}_{l-1}),

  4. (iv)

    for every r>0r>0 there exists C=C(r)>0C=C(r)>0 independent of ll such that

    j=l𝔴jrC(1+𝔴l)r.\sum_{j=l}^{\infty}\mathfrak{w}_{j}^{-r}\leq C(1+\mathfrak{w}_{l})^{-r}.

Assumption 7.1 is satisfied if (𝔴l)l(\mathfrak{w}_{l})_{l\in\mathbb{N}} is exponentially increasing, (for instance 𝔴l=2l\mathfrak{w}_{l}=2^{l}, ll\in\mathbb{N}). In the following we write 𝔚:={𝔴l:l0}\mathfrak{W}:=\{\mathfrak{w}_{l}\,:\,l\in\mathbb{N}_{0}\} and

x𝔚:=max{𝔴l:𝔴lx}.\lfloor x\rfloor_{\mathfrak{W}}:=\max\{\mathfrak{w}_{l}\,:\,\mathfrak{w}_{l}\leq x\}.

We work under the following hypothesis on the discretization errors in physical space: we quantify the convergence of the discretization scheme with respect to the discretization level ll\in\mathbb{N}. Specifically, we assume the approximation ulu^{l} to uu to behave asymptotically as

u(𝒚)ul(𝒚)XC(𝒚)𝔴lαl,\|u({\boldsymbol{y}})-u^{l}({\boldsymbol{y}})\|_{X}\leq C({\boldsymbol{y}})\mathfrak{w}_{l}^{-{\alpha}}\qquad\forall l\in\mathbb{N}, (7.1)

for some fixed convergence rate α>0{\alpha}>0 of the “physical space discretization” and with constant C(𝒚)>0C({\boldsymbol{y}})>0 depending on the parameter sequence 𝒚{\boldsymbol{y}}. We will make this assumption on ulu^{l} more precise shortly. If we think of ul(𝒚)H1(D)u^{l}({\boldsymbol{y}})\in H^{1}({{D}}) for the moment as a FEM approximation to the exact solution u(𝒚)H1(D)u({\boldsymbol{y}})\in H^{1}({{D}}) of some 𝒚{\boldsymbol{y}}-dependent elliptic PDE, then 𝔴l\mathfrak{w}_{l} could stand for the number of degrees of freedom of the finite element space. In this case α{\alpha} corresponds to the FEM convergence rate. Assumption (7.1) will for instance be satisfied if for each consecutive level the meshwidth is cut in half. Examples are provided by the FE spaces discussed in Section 2.6.2, Proposition 2.31. As long as the computational cost of computing the FEM solution is proportional to the dimension 𝔴l\mathfrak{w}_{l} of the FEM space, 𝔴lα\mathfrak{w}_{l}^{-{\alpha}} is the error in terms of the work 𝔴l\mathfrak{w}_{l}. Such an assumption usually holds in one spatial dimension, where the resulting stiffness matrix is tridiagonal. For higher spatial dimensions solving the corresponding linear system is often times not of linear complexity, in which case the convergence rate α>0{\alpha}>0 has to be adjusted accordingly.

We now state our assumptions on the sequence of functions (ul)l(u^{l})_{l\in\mathbb{N}} approximating uu. Equation (7.1) will hold in the L2L^{2} sense over all parameters 𝒚U{\boldsymbol{y}}\in U, cp. Assumption 7.2 (iii), and Definition 4.1 (ii).

Assumption 7.2.

Let XX be a separable Hilbert space and let (𝔴l)l0(\mathfrak{w}_{l})_{l\in\mathbb{N}_{0}} satisfy Assumption 7.1. Furthermore, 0<p1p2<0<p_{1}\leq p_{2}<\infty, 𝐛1p1(){\boldsymbol{b}}_{1}\in\ell^{p_{1}}(\mathbb{N}), 𝐛2p2(){\boldsymbol{b}}_{2}\in\ell^{p_{2}}(\mathbb{N}), ξ>0\xi>0, δ>0\delta>0 and there exist functions uL2(U,X;γ)u\in L^{2}(U,X;\gamma), (ul)lL2(U,X;γ)(u^{l})_{l\in\mathbb{N}}\subseteq L^{2}(U,X;\gamma) such that

  1. (i)

    uL2(U,X;γ)u\in L^{2}(U,X;\gamma) is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic,

  2. (ii)

    (uul)L2(U,X;γ)(u-u^{l})\in L^{2}(U,X;\gamma) is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic for every ll\in\mathbb{N},

  3. (iii)

    (uul)L2(U,X;γ)(u-u^{l})\in L^{2}(U,X;\gamma) is (𝒃2,ξ,δ𝔴lα,X)({\boldsymbol{b}}_{2},\xi,\delta\mathfrak{w}_{l}^{-{\alpha}},X)-holomorphic for every ll\in\mathbb{N}.

Remark 7.3.

Items (ii) and (iii) are two assumptions on the domain of holomorphic extension of the discretization error el:=uul:UXe_{l}:=u-u^{l}:U\to X. As pointed out in Remark 4.2, the faster the sequence 𝒃{\boldsymbol{b}} decays the larger the size of holomorphic extension, and the smaller δ\delta the smaller the upper bound of this extension.

Hence items (ii) and (iii) can be interpreted as follows: Item (ii) implies that ele_{l} has a large domain of holomorphic extension. Item (iii) is related to the assumption (7.1). It yields that by considering the extension of ele_{l} on a smaller domain, we can get a (ll-dependent) smaller upper bound of the extension of ele_{l} (in the sense of Definition 4.1 (ii)). Hence there is a tradeoff between choosing the size of the domain of the holomorphic extension and the upper bound of this extension.

7.2 Multilevel Smolyak sparse-grid algorithms

Let 𝐥=(l𝝂)𝝂0\mathbf{l}=(l_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subseteq\mathbb{N}_{0} be a family of natural numbers associating with each multiindex 𝝂{\boldsymbol{\nu}}\in\mathcal{F} of a PC expansion a discretization level l𝝂0l_{\boldsymbol{\nu}}\in\mathbb{N}_{0}. Typically, this is a family of discretization levels for some hierarchic, numerical approximation of the PDE in the physical domain D{D}, associating with each multiindex 𝝂{\boldsymbol{\nu}}\in\mathcal{F} of a PC expansion of the parametric solution in the parameter domain a possibly coefficient-dependent discretization level l𝝂0l_{\boldsymbol{\nu}}\in\mathbb{N}_{0}. With the sequence l𝝂0l_{\boldsymbol{\nu}}\in\mathbb{N}_{0}, we associate sets of multiindices via

Γj=Γj(𝐥):={𝝂:l𝝂j}j0.\Gamma_{j}=\Gamma_{j}(\mathbf{l}):=\{{\boldsymbol{\nu}}\in\mathcal{F}\,:\,l_{\boldsymbol{\nu}}\geq j\}\qquad\forall j\in\mathbb{N}_{0}. (7.2)

Throughout we will assume that

|𝐥|:=𝐥1():=𝝂l𝝂<|\mathbf{l}|:=\|\mathbf{l}\|_{\ell^{1}({\mathcal{F}})}:=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}l_{\boldsymbol{\nu}}<\infty

and that 𝐥\mathbf{l} is monotonically decreasing, meaning that 𝝂𝝁{\boldsymbol{\nu}}\leq{\boldsymbol{\mu}} implies l𝝂l𝝁l_{\boldsymbol{\nu}}\geq l_{\boldsymbol{\mu}}. In this case each Γj\Gamma_{j}\subseteq\mathcal{F}, jj\in\mathbb{N}, is finite and downward closed. Moreover Γ0=\Gamma_{0}=\mathcal{F}, and the sets (Γj)j0(\Gamma_{j})_{j\in{\mathbb{N}}_{0}} are nested according to

=Γ0Γ1Γ2.\mathcal{F}=\Gamma_{0}\supseteq\Gamma_{1}\supseteq\Gamma_{2}\dots.

With (ul)l(u^{l})_{l\in\mathbb{N}} as in Assumption 7.2, we now define the multilevel sparse-grid interpolation algorithm

𝐈ML𝐥u:=j(𝐈Γj𝐈Γj+1)uj.\mathbf{I}^{\rm ML}_{\mathbf{l}}u:=\sum_{j\in\mathbb{N}}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})u^{j}. (7.3)

A few remarks are in order. First, the index 𝐥\mathbf{l} indicates that the sets Γj=Γj(𝐥)\Gamma_{j}=\Gamma_{j}(\mathbf{l}) depend on the choice of 𝐥\mathbf{l}, although we usually simply write Γj\Gamma_{j} in order to keep the notation succinct. Secondly, due to |𝐥|<|\mathbf{l}|<\infty it holds

max𝝂l𝝂=:L<\max_{{\boldsymbol{\nu}}\in\mathcal{F}}l_{\boldsymbol{\nu}}=:L<\infty

and thus Γj=\Gamma_{j}=\emptyset for all j>Lj>L. Defining 𝐈\mathbf{I}_{\emptyset} as the constant 0 operator, the infinite series (7.3) can also be written as the finite sum

𝐈ML𝐥u=j=1L(𝐈Γj𝐈Γj+1)uj=𝐈Γ1u1+𝐈Γ2(u2u1)++𝐈ΓL(uLuL1),\mathbf{I}^{\rm ML}_{\mathbf{l}}u=\sum_{j=1}^{L}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})u^{j}=\mathbf{I}_{\Gamma_{1}}u^{1}+\mathbf{I}_{\Gamma_{2}}(u^{2}-u^{1})+\dots+\mathbf{I}_{\Gamma_{L}}(u^{L}-u^{L-1}),

where we used 𝐈ΓL+1=0\mathbf{I}_{\Gamma_{L+1}}=0. If we had Γ1==ΓL\Gamma_{1}=\dots=\Gamma_{L}, this sum would reduce to 𝐈ΓLuL\mathbf{I}_{\Gamma_{L}}u^{L}, which is the interpolant of the approximation uLu^{L} at the (highest) discretization level LL. The main observation of multilevel analyses is that it is beneficial not to choose all Γj\Gamma_{j} equal, but instead to balance out the accuracy of the interpolant 𝐈Γj\mathbf{I}_{\Gamma_{j}} (in the parameter) and the accuracy of the approximation uju^{j} of uu.

A multilevel sparse-grid quadrature algorithm is defined analogously via

𝐐ML𝐥u:=j(𝐐Γj𝐐Γj+1)uj,\mathbf{Q}^{\rm ML}_{\mathbf{l}}u:=\sum_{j\in\mathbb{N}}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})u^{j}, (7.4)

with Γj=Γj(𝐥)\Gamma_{j}=\Gamma_{j}(\mathbf{l}) as in (7.2). In the following we will prove algebraic convergence rates of multilevel interpolation and quadrature algorithms w.r.t. the L2(U,X;γ)L^{2}(U,X;\gamma)-norm and XX, respectively. The convergence rates will hold in terms of the work of computing 𝐈ML𝐥\mathbf{I}^{\rm ML}_{\mathbf{l}} and 𝐐ML𝐥\mathbf{Q}^{\rm ML}_{\mathbf{l}}.

As mentioned above, for a level ll\in\mathbb{N}, we interpret 𝔴l\mathfrak{w}_{l}\in\mathbb{N} as a measure of the computational complexity of evaluating ulu^{l} at an arbitrary parameter 𝒚U{\boldsymbol{y}}\in U. As discussed in Section 6.2.1, computing 𝐈Γju\mathbf{I}_{\Gamma_{j}}u or 𝐐Γju\mathbf{Q}_{\Gamma_{j}}u requires to evaluate the function uu at each parameter in the set pts(Γj)U{\rm pts}(\Gamma_{j})\subseteq U introduced in (6.11). We recall the bound

|pts(Γj)|𝝂Γjp𝝂(1),|{\rm pts}(\Gamma_{j})|\leq\sum_{{\boldsymbol{\nu}}\in\Gamma_{j}}p_{\boldsymbol{\nu}}(1),

on the cardinality of this set obtained in (6.12). As an upper bound of the work corresponding to the evaluation of all functions required for the multilevel interpolant in (7.3), we obtain

j𝔴j(𝝂Γj(𝐥)p𝝂(1)+𝝂Γj+1(𝐥)p𝝂(1)).\sum_{j\in\mathbb{N}}\mathfrak{w}_{j}\left(\sum_{{\boldsymbol{\nu}}\in\Gamma_{j}(\mathbf{l})}p_{\boldsymbol{\nu}}(1)+\sum_{{\boldsymbol{\nu}}\in\Gamma_{j+1}(\mathbf{l})}p_{\boldsymbol{\nu}}(1)\right). (7.5)

Since Γj+1Γj\Gamma_{j+1}\subseteq\Gamma_{j}, up the factor 22 the work of a sequence 𝐥\mathbf{l} is defined by

work(𝐥):=j=1L𝔴j𝝂Γj(𝐥)p𝝂(1)=𝝂(𝐥)p𝝂(1)j=1l𝝂𝔴j,\mathrm{work}(\mathbf{l}):=\sum_{j=1}^{L}\mathfrak{w}_{j}\sum_{{\boldsymbol{\nu}}\in\Gamma_{j}(\mathbf{l})}p_{\boldsymbol{\nu}}(1)=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}(\mathbf{l})}p_{\boldsymbol{\nu}}(1)\sum_{j=1}^{l_{\boldsymbol{\nu}}}\mathfrak{w}_{j}, (7.6)

where we used the definition of Γj(𝐥)\Gamma_{j}(\mathbf{l}) in (7.2), L:=max𝝂l𝝂<L:=\max_{{\boldsymbol{\nu}}\in\mathcal{F}}l_{\boldsymbol{\nu}}<\infty and the finiteness of the set

(𝐥):={𝝂:l𝝂>0}.\mathcal{F}(\mathbf{l}):=\{{\boldsymbol{\nu}}\in\mathcal{F}:l_{\boldsymbol{\nu}}>0\}.

The efficiency of the multilevel interpolant critically relies on a suitable choice of levels 𝐥=(l𝝂)𝝂\mathbf{l}=(l_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}. This will be achieved with the following algorithm, which constructs 𝐥\mathbf{l} based on two collections of positive real numbers, (c𝝂)𝝂q1()(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{1}}(\mathcal{F}) and (d𝝂)𝝂q2()(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{2}}(\mathcal{F}). The algorithm is justified due to Lemma 7.4 which was shown in Section 7.3. This technical lemma, which is a variant of [111, Lemma 3.2.7], constitutes the central part of the proofs of the convergence rate results presented in the rest of this section.

Algorithm 2 (l𝝂)𝝂=ConstructLevels((c𝝂)𝝂,(d𝝂)𝝂,q1,α,ε)(l_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}={\rm ConstructLevels}((c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}},(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}},q_{1},{\alpha},\varepsilon)
1:(l𝝂)𝝂(0)𝝂(l_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\leftarrow(0)_{{\boldsymbol{\nu}}\in\mathcal{F}}
2:Λε{𝝂:c𝝂1ε}\Lambda_{\varepsilon}\leftarrow\{{\boldsymbol{\nu}}\in\mathcal{F}\,:\,c_{{\boldsymbol{\nu}}}^{-1}\geq\varepsilon\}
3:for 𝝂Λε{\boldsymbol{\nu}}\in\Lambda_{\varepsilon} do
4:     δε1/2q1/4αd𝝂11+2α(𝝁Λεd𝝁11+2α)12α\delta\leftarrow\varepsilon^{-\frac{1/2-q_{1}/4}{{\alpha}}}d_{{\boldsymbol{\nu}}}^{\frac{-1}{1+2{\alpha}}}\left(\sum_{{\boldsymbol{\mu}}\in\Lambda_{\varepsilon}}d_{{\boldsymbol{\mu}}}^{\frac{-1}{1+2{\alpha}}}\right)^{\frac{1}{2{\alpha}}}
5:     l𝝂max{j0:𝔴jδ}l_{{\boldsymbol{\nu}}}\leftarrow\max\{j\in\mathbb{N}_{0}\,:\,\mathfrak{w}_{j}\leq\delta\}
6:return (l𝝂)𝝂(l_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}

Note that the determination of the sets Λε\Lambda_{\varepsilon} in line 2 of Algorithm 2 can be done with Algorithm 1.

7.3 Construction of an allocation of discretization levels

We detail the construction of an allocation of discretization levels along the coefficients of Wiener-Hermite PC expansion. It is valid for collections (u𝝂)𝝂(u_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} of Wiener-Hermite PC expansion coefficients taking values in a separable Hilbert space, say XX, with additional regularity, being XsXX^{s}\subset X, allowing for weaker (weighted) summability of the VsV^{s}-norms (u𝝂Xs)𝝂(\|u_{\boldsymbol{\nu}}\|_{X^{s}})_{{\boldsymbol{\nu}}\in\mathcal{F}}. In the setting of elliptic BVPs with log-Gaussian diffusion coefficient, X=V=H10(D)X=V=H^{1}_{0}({D}), and XsX^{s} is, for example, a weighted Kondrat’ev space in D{D} as introduced in Section 3.8.1. We phrase the result and the construction in abstract terms so that the allocation is applicable to more general settings, such as the parabolic IBVP in Section 4.3.2.

For a given, dense sequence (Xl)l0X({X}_{l})_{l\in{\mathbb{N}}_{0}}\subset X of nested, finite-dimensional subspaces and target accuracy 0<ε10<\varepsilon\leq 1, in the numerical approximation of Wiener-Hermite PC expansions of random fields uu taking values in XX, we consider approximating the Wiener-Hermite PC expansion coefficients u𝝂u_{\boldsymbol{\nu}} in XX from XlX_{l}. The assumed density of the sequence (Xl)l0X(X_{l})_{l\in{\mathbb{N}}_{0}}\subset X in XX ensures that for uL2(U,X;γ)u\in L^{2}(U,X;\gamma) the coefficients (u𝝂)𝝂X(u_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subset X are square summable, in the sense that (u𝝂X)𝝂2()(\|u_{\boldsymbol{\nu}}\|_{X})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell_{2}({\mathcal{F}})

The following lemma is a variation of [111, Lemma 3.2.7]. Its proof, is, with several minor modifications, taken from [111, Lemma 3.2.7]. We remark that the construction of the map 𝐥(ε,𝝂)\mathbf{l}(\varepsilon,{\boldsymbol{\nu}}), as described in the lemma, mimicks Algorithm 2. Again, a convergence rate is obtained that is not prone to the so-called “curse of dimensionality”, being limited only by the available sparsity in the coefficients of Wiener-Hermite PC expansion for the parametric solution manifold.

Lemma 7.4.

Let 𝔚={𝔴l:l0}\mathfrak{W}=\{\mathfrak{w}_{l}\,:\,l\in\mathbb{N}_{0}\} satisfy Assumption 7.1. Let q1[0,2)q_{1}\in[0,2), q2[q1,){q_{2}}\in[q_{1},\infty) and α>0{\alpha}>0. Let

  1. (i)

    (aj,𝝂)𝝂[0,)(a_{j,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subseteq[0,\infty) for every j0j\in\mathbb{N}_{0},

  2. (ii)

    (c𝝂)𝝂(0,)(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subseteq(0,\infty) and (d𝝂)𝝂(0,)(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}\subseteq(0,\infty) be such that

    (c𝝂1/2)𝝂q1()and(d𝝂1/2p𝝂(1/2+α))𝝂q2(),(c_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{1}}(\mathcal{F})\;\mbox{and}\;\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}p_{\boldsymbol{\nu}}(1/2+\alpha)\big{)}_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{{q_{2}}}(\mathcal{F}),
  3. (iii)
    supj0(𝝂aj,𝝂2c𝝂)1/2=:C1<,supj0(𝝂(𝔴jαaj,𝝂)2d𝝂)1/2=:C2<.\sup_{j\in\mathbb{N}_{0}}\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}a_{j,{\boldsymbol{\nu}}}^{2}c_{{\boldsymbol{\nu}}}\right)^{1/2}=:C_{1}<\infty,\qquad\sup_{j\in\mathbb{N}_{0}}\left(\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}(\mathfrak{w}_{j}^{{\alpha}}a_{j,{\boldsymbol{\nu}}})^{2}d_{{\boldsymbol{\nu}}}\right)^{1/2}=:C_{2}<\infty. (7.7)

For every ε>0\varepsilon>0 define Λε={𝛎:c𝛎1ε}\Lambda_{\varepsilon}=\{{\boldsymbol{\nu}}\in\mathcal{F}\,:\,c_{{\boldsymbol{\nu}}}^{-1}\geq\varepsilon\}, ωε,𝛎:=0\omega_{\varepsilon,{\boldsymbol{\nu}}}:=0 for all 𝛎\Λε{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}, and define

ωε,𝝂:=ε1/2q1/4αd𝝂11+2α(𝝁Λεd𝝁11+2α)12α𝔚𝔚𝝂Λε.\omega_{\varepsilon,{\boldsymbol{\nu}}}:=\left\lfloor\varepsilon^{-\frac{1/2-q_{1}/4}{{\alpha}}}d_{{\boldsymbol{\nu}}}^{\frac{-1}{1+2{\alpha}}}\Bigg{(}\sum_{{\boldsymbol{\mu}}\in\Lambda_{\varepsilon}}d_{{\boldsymbol{\mu}}}^{\frac{-1}{1+2{\alpha}}}\Bigg{)}^{\frac{1}{2{\alpha}}}\right\rfloor_{\mathfrak{W}}\in\mathfrak{W}\qquad\forall{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}.

Furthermore, for every ε>0\varepsilon>0 and 𝛎{\boldsymbol{\nu}}\in\mathcal{F} let lε,𝛎0l_{\varepsilon,{\boldsymbol{\nu}}}\in\mathbb{N}_{0} be the corresponding discretization level, i.e., ωε,𝛎=𝔴lε,𝛎\omega_{\varepsilon,{\boldsymbol{\nu}}}=\mathfrak{w}_{l_{\varepsilon,{\boldsymbol{\nu}}}}, and define the maximal discretization level

L(ε):=max{lε,𝝂:𝝂}.L(\varepsilon):=\max\{{l_{\varepsilon,{\boldsymbol{\nu}}}}\,:\,{\boldsymbol{\nu}}\in\mathcal{F}\}.

Denote 𝐥ε=(lε,𝛎)𝛎\mathbf{l}_{\varepsilon}=({l_{\varepsilon,{\boldsymbol{\nu}}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}.

Then there exists a constant C>0C>0 and tolerances εn(0,1]\varepsilon_{n}\in(0,1] such that for every nn\in{\mathbb{N}} holds work(𝐥εn)n\mathrm{work}(\mathbf{l}_{\varepsilon_{n}})\leq n and

𝝂j=lεn,𝝂L(εn)aj,𝝂C(1+logn)nR,\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon_{n},{\boldsymbol{\nu}}}}}^{L(\varepsilon_{n})}a_{j,{\boldsymbol{\nu}}}\leq C(1+\log n)n^{-R},

where the rate RR is given by

R=min{α,α(q111/2)α+q11q21}.R=\min\left\{{\alpha},\frac{{\alpha}(q_{1}^{-1}-1/2)}{{\alpha}+q_{1}^{-1}-{q_{2}}^{-1}}\right\}.
Proof.

Throughout this proof denote δ:=1/2q1/4>0\delta:=1/2-q_{1}/4>0. In the following

ω~ε,𝝂:=εδαd𝝂11+2α(𝝁Λεd𝝁11+2α)12α𝝂Λε,\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}:=\varepsilon^{-\frac{\delta}{{\alpha}}}d_{{\boldsymbol{\nu}}}^{\frac{-1}{1+2{\alpha}}}\Bigg{(}\sum_{{\boldsymbol{\mu}}\in\Lambda_{\varepsilon}}d_{{\boldsymbol{\mu}}}^{\frac{-1}{1+2{\alpha}}}\Bigg{)}^{\frac{1}{2{\alpha}}}\qquad\forall{\boldsymbol{\nu}}\in\Lambda_{\varepsilon},

i.e. ωε,𝝂=ω~ε,𝝂𝔚\omega_{\varepsilon,{\boldsymbol{\nu}}}=\lfloor\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}\rfloor_{\mathfrak{W}}. Note that 0<ω~ε,𝝂0<\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}} is well-defined for all 𝝂Λε{\boldsymbol{\nu}}\in\Lambda_{\varepsilon} since d𝝂>0d_{{\boldsymbol{\nu}}}>0 for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F} by assumption. Due to Assumption 7.1 (iii) it holds

ω~ε,𝝂K𝔚1+ωε,𝝂1+ω~ε,𝝂𝝂Λε.\frac{\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}}{K_{\mathfrak{W}}}\leq 1+\omega_{\varepsilon,{\boldsymbol{\nu}}}\leq 1+\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}\qquad\forall{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}. (7.8)

Since (c𝝂1/2)𝝂q1()(c_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{1}}(\mathcal{F}) and (7.7), we get

𝝂\Λεaj,𝝂(𝝂\Λεaj,𝝂2c𝝂)1/2(𝝂\Λεc𝝂1)1/2C1(c𝝂1εc𝝂q12c𝝂q121)1/2Cεδ\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}a_{j,{\boldsymbol{\nu}}}\leq\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}a_{j,{\boldsymbol{\nu}}}^{2}c_{{\boldsymbol{\nu}}}\Bigg{)}^{1/2}\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}c_{{\boldsymbol{\nu}}}^{-1}\Bigg{)}^{1/2}\leq C_{1}\Bigg{(}\sum_{c_{{\boldsymbol{\nu}}}^{-1}\leq\varepsilon}c_{{\boldsymbol{\nu}}}^{-\frac{q_{1}}{2}}c_{{\boldsymbol{\nu}}}^{\frac{q_{1}}{2}-1}\Bigg{)}^{1/2}\leq C\varepsilon^{\delta}

with the constant CC independent of jj and ε\varepsilon. Thus,

𝝂\Λεj=0L(ε)aj,𝝂=j=0L(ε)𝝂\Λεaj,𝝂C1(1+L(ε))εδ.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}\sum_{j=0}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}=\sum_{j=0}^{L(\varepsilon)}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}a_{j,{\boldsymbol{\nu}}}\leq C_{1}(1+L(\varepsilon))\varepsilon^{\delta}. (7.9)

Next with C2C_{2} as in (7.7),

𝝂Λεj=lε,𝝂L(ε)aj,𝝂\displaystyle\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}} =𝝂Λεj=lε,𝝂L(ε)aj,𝝂𝔴jα𝔴jαd𝝂1/2d𝝂1/2\displaystyle=\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}\mathfrak{w}_{j}^{\alpha}\mathfrak{w}_{j}^{-{\alpha}}d_{{\boldsymbol{\nu}}}^{1/2}d_{{\boldsymbol{\nu}}}^{-1/2}
(𝝂Λεj=0L(ε)(aj,𝝂𝔴jαd𝝂1/2)2)12(𝝂Λεjlε,𝝂(d𝝂1/2𝔴jα)2)12\displaystyle\leq\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j=0}^{L(\varepsilon)}\big{(}a_{j,{\boldsymbol{\nu}}}\mathfrak{w}_{j}^{\alpha}d_{{\boldsymbol{\nu}}}^{1/2}\big{)}^{2}\Bigg{)}^{\frac{1}{2}}\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j\geq{l_{\varepsilon,{\boldsymbol{\nu}}}}}\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}\mathfrak{w}_{j}^{-{\alpha}}\big{)}^{2}\Bigg{)}^{\frac{1}{2}}
C2(1+L(ε))(𝝂Λεjlε,𝝂(d𝝂1/2𝔴jα)2)12.\displaystyle\leq C_{2}(1+L(\varepsilon))\Bigg{(}\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j\geq{l_{\varepsilon,{\boldsymbol{\nu}}}}}\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}\mathfrak{w}_{j}^{-{\alpha}}\big{)}^{2}\Bigg{)}^{\frac{1}{2}}. (7.10)

Assumption 7.1 (iv) implies for some C3C_{3}

jlε,𝝂𝔴j2αC32(1+𝔴lε,𝝂)2α=C32(1+ωε,𝝂)2α,\sum_{j\geq{l_{\varepsilon,{\boldsymbol{\nu}}}}}\mathfrak{w}_{j}^{-2{\alpha}}\leq{C_{3}^{2}}(1+\mathfrak{w}_{{l_{\varepsilon,{\boldsymbol{\nu}}}}})^{-2{\alpha}}={C_{3}^{2}}(1+\omega_{\varepsilon,{\boldsymbol{\nu}}})^{-2{\alpha}},

so that by (7.8) and (7.3)

𝝂Λεj=lε,𝝂L(ε)aj,𝝂\displaystyle\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}} C3C2(1+L(ε))(𝝂Λε(d𝝂1/2(1+ωε,𝝂)α)2)12\displaystyle\leq C_{3}C_{2}(1+L(\varepsilon))\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}(1+\omega_{\varepsilon,{\boldsymbol{\nu}}})^{-{\alpha}}\big{)}^{2}\right)^{\frac{1}{2}}
C3C2K𝔚α(1+L(ε))(𝝂Λε(d𝝂1/2ω~ε,𝝂α)2)12.\displaystyle\leq C_{3}C_{2}{K_{\mathfrak{W}}^{\alpha}}(1+L(\varepsilon))\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}^{-{\alpha}}\big{)}^{2}\right)^{\frac{1}{2}}. (7.11)

Inserting the definition of ω~ε,𝝂\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}, we have

(𝝂Λε(d𝝂1/2ω~ε,𝝂α)2)12\displaystyle\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\big{(}d_{{\boldsymbol{\nu}}}^{-1/2}\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}^{-{\alpha}}\big{)}^{2}\right)^{\frac{1}{2}} =εδ(𝝁Λεd𝝁11+2α)α12α(𝝂Λεd𝝂1d𝝂2α1+2α)12=εδ,\displaystyle=\varepsilon^{\delta}\Bigg{(}\sum_{{\boldsymbol{\mu}}\in\Lambda_{\varepsilon}}d_{{\boldsymbol{\mu}}}^{\frac{-1}{1+2{\alpha}}}\Bigg{)}^{-{\alpha}\frac{1}{2{\alpha}}}\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}d_{{\boldsymbol{\nu}}}^{-1}d_{{\boldsymbol{\nu}}}^{\frac{2{\alpha}}{1+2{\alpha}}}\right)^{\frac{1}{2}}=\varepsilon^{\delta}, (7.12)

where we used

1+2α1+2α=(1+2α)+2α1+2α=11+2α.-1+\frac{2{\alpha}}{1+2{\alpha}}=\frac{-(1+2{\alpha})+2{\alpha}}{1+2{\alpha}}=\frac{-1}{1+2{\alpha}}.

Using Assumption 7.1 (ii) and the definition of work(𝐥ε)\mathrm{work}(\mathbf{l}_{\varepsilon}) in (7.6) we get

L(ε)log(1+max𝝂ωε,𝝂)log(1+work(𝐥ε)).L(\varepsilon)\leq\log(1+\max_{{\boldsymbol{\nu}}\in\mathcal{F}}\omega_{\varepsilon,{\boldsymbol{\nu}}})\leq\log(1+\mathrm{work}(\mathbf{l}_{\varepsilon})). (7.13)

Hence, (7.9), (7.3), (7.12) and (7.13) yield

𝝂j=lε,𝝂L(ε)aj,𝝂=𝝂Λεj=lε,𝝂L(ε)aj,𝝂+𝝂\Λεj=0L(ε)aj,𝝂C(1+log(work(𝐥ε)))εδ.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}=\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}+\sum_{{\boldsymbol{\nu}}\in\mathcal{F}\backslash\Lambda_{\varepsilon}}\sum_{j=0}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}\leq C\big{(}1+\log(\mathrm{work}(\mathbf{l}_{\varepsilon}))\big{)}\varepsilon^{\delta}. (7.14)

Next, we compute an upper bound for work(𝐥ε)\mathrm{work}(\mathbf{l}_{\varepsilon}). By definition of work(𝐥ε)\mathrm{work}(\mathbf{l}_{\varepsilon}) in (7.6), and using Assumption 7.1 (i) as well as ωε,𝝂=𝔴lε,𝝂\omega_{\varepsilon,{\boldsymbol{\nu}}}=\mathfrak{w}_{{l_{\varepsilon,{\boldsymbol{\nu}}}}},

work(𝐥ε)\displaystyle\mathrm{work}(\mathbf{l}_{\varepsilon}) =𝝂Λεp𝝂(1){j:jlε,𝝂}𝔴j𝝂Λεp𝝂(1)K𝔚ωε,𝝂\displaystyle=\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}p_{\boldsymbol{\nu}}(1)\sum_{\{j\in\mathbb{N}\,:\,j\leq{l_{\varepsilon,{\boldsymbol{\nu}}}}\}}\mathfrak{w}_{j}\leq\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}p_{\boldsymbol{\nu}}(1)K_{\mathfrak{W}}\omega_{\varepsilon,{\boldsymbol{\nu}}}
K𝔚𝝂Λεp𝝂(1)ω~ε,𝝂K𝔚εδα(𝝂Λεp𝝂(1)d𝝂11+2α)12α+1\displaystyle\leq K_{\mathfrak{W}}\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}p_{\boldsymbol{\nu}}(1)\tilde{\omega}_{\varepsilon,{\boldsymbol{\nu}}}\leq K_{\mathfrak{W}}\varepsilon^{-\frac{\delta}{{\alpha}}}\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}p_{\boldsymbol{\nu}}(1)d_{{\boldsymbol{\nu}}}^{\frac{-1}{1+2{\alpha}}}\right)^{\frac{1}{2{\alpha}}+1}
=K𝔚εδα(𝝂Λε(p𝝂(1/2+α)d𝝂1/2)21+2α)12α+1,\displaystyle=K_{\mathfrak{W}}\varepsilon^{-\frac{\delta}{{\alpha}}}\left(\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}\left(p_{\boldsymbol{\nu}}(1/2+{\alpha})d_{{\boldsymbol{\nu}}}^{-1/2}\right)^{\frac{2}{1+2{\alpha}}}\right)^{\frac{1}{2{\alpha}}+1}, (7.15)

where we used p𝝂(1)=p𝝂(1/2+α)2/(1+2α)p_{\boldsymbol{\nu}}(1)=p_{\boldsymbol{\nu}}(1/2+{\alpha})^{2/(1+2{\alpha})} and the fact that p𝝂(1)1p_{\boldsymbol{\nu}}(1)\geq 1 for all 𝝂{\boldsymbol{\nu}}.

We distinguish between the two cases

21+2αq2and21+2α<q2.\frac{2}{1+2{\alpha}}\geq{q_{2}}\qquad\text{and}\qquad\frac{2}{1+2{\alpha}}<{q_{2}}.

In the first case, since (p𝝁(1/2+α)d𝝁1/2)𝝁q2()(p_{\boldsymbol{\mu}}(1/2+{\alpha})d_{{\boldsymbol{\mu}}}^{-1/2})_{{\boldsymbol{\mu}}\in\mathcal{F}}\in\ell^{{q_{2}}}(\mathcal{F}), (7.3) implies

work(𝐥ε)Cεδα\mathrm{work}(\mathbf{l}_{\varepsilon})\leq C\varepsilon^{-\frac{\delta}{{\alpha}}} (7.16)

and hence, log(work(𝐥ε))log(Cεδα)\log(\mathrm{work}(\mathbf{l}_{\varepsilon}))\leq\log\big{(}C\varepsilon^{-\frac{\delta}{{\alpha}}}\big{)}. Then (7.14) together with (7.16) implies

𝝂j=lε,𝝂L(ε)aj,𝝂C(1+|log(ε1)|)εδ.\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}\leq C(1+{|\log(\varepsilon^{-1})|})\varepsilon^{\delta}.

For every nn\in{\mathbb{N}}, we can find εn>0\varepsilon_{n}>0 such that n2Cεnδαn\frac{n}{2}\leq C\varepsilon_{n}^{-\frac{\delta}{{\alpha}}}\leq n. Then the claim of the corollary in the case 21+2αq2\frac{2}{1+2{\alpha}}\geq q_{2} holds true for the chosen εn\varepsilon_{n}.

Finally, let us address the case 21+2α<q2\frac{2}{1+2{\alpha}}<{q_{2}}. Then, by (7.3) and using Hölder’s inequality with q21+2α2>1{q_{2}}\frac{1+2{\alpha}}{2}>1 we get

work(𝐥ε)K𝔚εδα(p𝝂(1/2+α)d𝝂1/2)𝝂q2()1α|Λε|(12q2(1+2α))1+2α2α.\mathrm{work}(\mathbf{l}_{\varepsilon})\leq K_{\mathfrak{W}}\varepsilon^{-\frac{\delta}{{\alpha}}}\|(p_{\boldsymbol{\nu}}(1/2+{\alpha})d_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\|_{\ell^{{q_{2}}}(\mathcal{F})}^{\frac{1}{{\alpha}}}|\Lambda_{\varepsilon}|^{\big{(}1-\frac{2}{q_{2}(1+2\alpha)}\big{)}\frac{1+2{\alpha}}{2{\alpha}}}.

Since

|Λε|=𝝂Λε1=c𝝂1εc𝝂q12c𝝂q12Cεq12,|\Lambda_{\varepsilon}|=\sum_{{\boldsymbol{\nu}}\in\Lambda_{\varepsilon}}1=\sum_{c_{{\boldsymbol{\nu}}}^{-1}\geq\varepsilon}c_{{\boldsymbol{\nu}}}^{-\frac{q_{1}}{2}}c_{{\boldsymbol{\nu}}}^{\frac{q_{1}}{2}}\leq C\varepsilon^{-\frac{q_{1}}{2}},

we obtain

work(𝐥ε)K𝔚εδαq12(12q2(1+2α))1+2α2αCεq12α(α1q2+1q1).\displaystyle\mathrm{work}(\mathbf{l}_{\varepsilon})\leq K_{\mathfrak{W}}\varepsilon^{-\frac{\delta}{{\alpha}}-\frac{q_{1}}{2}(1-\frac{2}{q_{2}(1+2\alpha)})\frac{1+2{\alpha}}{2{\alpha}}}\leq C\varepsilon^{-\frac{q_{1}}{2\alpha}\big{(}\alpha-\frac{1}{q_{2}}+\frac{1}{q_{1}}\big{)}}.

For every nn\in{\mathbb{N}}, we can find εn>0\varepsilon_{n}>0 such that

n2Cεnq12α(α1q2+1q1)n.\frac{n}{2}\leq C\varepsilon_{n}^{-\frac{q_{1}}{2\alpha}\big{(}\alpha-\frac{1}{q_{2}}+\frac{1}{q_{1}}\big{)}}\leq n.

Thus the claim also holds true in the case 21+2α<q2\frac{2}{1+2{\alpha}}<q_{2}. ∎

7.4 Multilevel Smolyak sparse-grid interpolation algorithm

We are now in position to formulate a multilevel Smolyak sparse-grid interpolation convergence theorem. To this end, we observe that our proofs of approximation rates have been constructive: rather than being based on a best NN-term selection from the infinite set of Wiener-Hermite PC expansion coefficients, a constructive selection process of “significant” Wiener-Hermite PC expansion coefficients, subject to a given prescribed approximation tolerance, has been provided. In the present section, we turn this into a concrete, numerical selection process with complexity bounds. In particular, we provide an a-priori allocation of discretization levels to Wiener-Hermite PC expansion coefficients. This results on the one hand in an explicit, algorithmic definition of a family of multilevel interpolants which is parametrized by an approximation threshold ε>0\varepsilon>0. On the other hand, it will result in mathematical convergence rate bounds in terms of computational work rather than in terms of, for example, number of active Wiener-Hermite PC expansion coefficients, which rate bounds are free from the curse of dimensionality.

The idea is as follows: let 𝒃1=(b1,j)jp1(){\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}}\in\ell^{p_{1}}({\mathbb{N}}), 𝒃2=(b2,j)jp2(){\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}}\in\ell^{p_{2}}({\mathbb{N}}), and ξ\xi be the two sequences and constant from Assumption 7.2. For two constants K>0K>0 and r>3r>3 (which are still at our disposal and which will be specified below), set for all jj\in\mathbb{N}

ϱ1,j:=b1,jp11ξ4𝒃1p1,ϱ2,j:=b2,jp21ξ4𝒃2p2.\varrho_{1,j}:=b_{1,j}^{p_{1}-1}\frac{\xi}{4\|{\boldsymbol{b}}_{1}\|_{\ell^{p_{1}}}},\qquad\varrho_{2,j}:=b_{2,j}^{p_{2}-1}\frac{\xi}{4\|{\boldsymbol{b}}_{2}\|_{\ell^{p_{2}}}}. (7.17)

We let for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F} (as in Lemma 6.5 for k=1k=1 and with τ=3\tau=3)

c𝝂:=jmax{1,Kϱ1,j}2νjr3,d𝝂:=jmax{1,Kϱ2,j}2νjr3.c_{{\boldsymbol{\nu}}}:=\prod_{j\in\mathbb{N}}\max\{1,K\varrho_{1,j}\}^{2}\nu_{j}^{r-3},\qquad d_{{\boldsymbol{\nu}}}:=\prod_{j\in\mathbb{N}}\max\{1,K\varrho_{2,j}\}^{2}\nu_{j}^{r-3}. (7.18)

Based on those two multi-index collections, Algorithm 2 provides a collection of discretization levels which sequence depends on ε>0\varepsilon>0 and is indexed over \mathcal{F}. We denote it by 𝐥ε=(lε,𝝂)𝝂\mathbf{l}_{\varepsilon}=(l_{\varepsilon,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}. We now state an upper bound for the error of the corresponding multilevel interpolants in terms of the work measure in (7.6) as ε0\varepsilon\to 0.

Theorem 7.5.

Let uL2(U,X;γ)u\in L^{2}(U,X;\gamma) and ulL2(U,X;γ)u^{l}\in L^{2}(U,X;\gamma), ll\in\mathbb{N}, satisfy Assumption 7.2 with some constants α>0{\alpha}>0 and 0<p1<2/30<p_{1}<2/3 and p1p2<1p_{1}\leq p_{2}<1. Set q1:=p1/(1p1)q_{1}:=p_{1}/(1-p_{1}). Assume that r>2(1+(α+1)q1)/q1+3r>2(1+({\alpha}+1)q_{1})/q_{1}+3 (for rr as defined in (7.18)). There exist constants K>0K>0 (in (7.18)) and C>0C>0 such that the following holds.

For every nn\in{\mathbb{N}}, there are positive constants εn(0,1]\varepsilon_{n}\in(0,1] such that work(𝐥εn)n\mathrm{work}(\mathbf{l}_{\varepsilon_{n}})\leq n and with 𝐥εn=(lεn,𝛎)𝛎\mathbf{l}_{\varepsilon_{n}}=({l_{\varepsilon_{n},{\boldsymbol{\nu}}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} as defined in Lemma 7.4 (where c𝛎c_{\boldsymbol{\nu}}, d𝛎d_{\boldsymbol{\nu}} as in (7.18)) it holds

u𝐈𝐥εnMLuL2(U,X;γ)C(1+logn)nR\|u-\mathbf{I}_{\mathbf{l}_{\varepsilon_{n}}}^{\rm ML}u\|_{L^{2}(U,X;\gamma)}\ \leq\ C(1+\log n)n^{-R}

with the convergence rate

R:=min{α,α(p113/2)α+p11p21}.R:=\min\left\{{\alpha},\frac{{\alpha}(p_{1}^{-1}-3/2)}{{\alpha}+p_{1}^{-1}-p_{2}^{-1}}\right\}. (7.19)
Proof.

Throughout this proof we write 𝒃1=(b1,j)j{\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}} and 𝒃2=(b2,j)j{\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}} for the two sequences in Assumption 7.2. We observe that Γj\Gamma_{j} defined in (7.2) is downward closed for all j0j\in\mathbb{N}_{0}. This can be easily deduced from the fact that the multi-index collections (c𝝂)𝝂(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} and (d𝝂)𝝂(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} are monotonically increasing (i.e., e.g., 𝝂𝝁{\boldsymbol{\nu}}\leq{\boldsymbol{\mu}} implies c𝝂c𝝁c_{{\boldsymbol{\nu}}}\leq c_{{\boldsymbol{\mu}}}) and the definition of Λε\Lambda_{\varepsilon} and lε,𝝂l_{\varepsilon,{\boldsymbol{\nu}}} in Algorithm 2. We will use this fact throughout the proof, without mentioning it at every instance.

Step 1. Given nn\in\mathbb{N}, we choose ε:=εn\varepsilon:=\varepsilon_{n} as in Lemma 7.4. Fix NN\in\mathbb{N} such that

N>max{j:jsupp(𝝂),lε,𝝂>0}N>\max\{j:\,j\in\operatorname{supp}({\boldsymbol{\nu}}),\,l_{\varepsilon,{\boldsymbol{\nu}}}>0\}

and so large that

uu~NL2(U,X;γ)nR,\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}\leq n^{-R}, (7.20)

where u~N:UX\tilde{u}_{N}:U\to X is as in Definition 4.1. This is possible due to

limNuu~NL2(U,X;γ)=0,\lim_{N\to\infty}\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}=0,

which holds by the (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphy of uu. By Assumption 7.2, for every jj\in\mathbb{N} the function ej:=uujL2(U,X;γ)e^{j}:=u-u^{j}\in L^{2}(U,X;\gamma) is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic and (𝒃2,ξ,δ𝔴jγ,X)({\boldsymbol{b}}_{2},\xi,\delta\mathfrak{w}_{j}^{\gamma},X)-holomorphic. For notational convenience we set e0:=u0=uL2(U,X;γ)e^{0}:=u-0=u\in L^{2}(U,X;\gamma), so that e0e^{0} is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic and (𝒃2,ξ,δ,X)({\boldsymbol{b}}_{2},\xi,\delta,X)-holomorphic. Hence, for every j0j\in\mathbb{N}_{0} there exists a function e~Nj=u~Nu~Nj\tilde{e}_{N}^{j}=\tilde{u}_{N}-\tilde{u}_{N}^{j} as in Definition 4.1 (iii).

In the rest of the proof we use the following facts:

  1. (i)

    By Lemma 6.10, for every j0j\in\mathbb{N}_{0}, with the Wiener-Hermite PC expansion coefficients

    e~N,𝝂j:=UH𝝂(𝒚)e~Nj(𝒚)dγ(𝒚),\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}:=\int_{U}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\tilde{e}_{N}^{j}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}),

    it holds

    e~Nj(𝒚)=𝝂e~N,𝝂jH𝝂(𝒚)𝒚U,\tilde{e}_{N}^{j}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\qquad\forall{\boldsymbol{y}}\in U,

    with pointwise absolute convergence.

  2. (ii)

    By Lemma 6.5, upon choosing K>0K>0 in (7.18) large enough, and because r>3r>3,

    C0c𝝂p𝝂(3)β𝝂(r,ϱ1),C0d𝝂p𝝂(3)β𝝂(r,ϱ2)𝝂1.C_{0}c_{{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(3)\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{1}),\qquad C_{0}d_{{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(3)\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{2})\qquad\forall{\boldsymbol{\nu}}\in\mathcal{F}_{1}.

    We observe that by definition of ϱi{\boldsymbol{\varrho}}_{i}, i{1,2}i\in\{1,2\}, in (7.17), it holds ϱi,jbi,j(1pi)\varrho_{i,j}\sim b_{i,j}^{-(1-p_{i})} and therefore (ϱi,j1)jqi()(\varrho_{i,j}^{-1})_{j\in\mathbb{N}}\in\ell^{q_{i}}(\mathbb{N}) with qi:=pi/(1pi)q_{i}:=p_{i}/(1-p_{i}), i{1,2}i\in\{1,2\}.

  3. (iii)

    Due to r>2(1+(α+1)q1)/q1+3r>2(1+({\alpha}+1)q_{1})/q_{1}+3, the condition of Lemma 6.6 is satisfied (with k=1k=1, τ=3\tau=3 and θ=(α+1)q1\theta=({\alpha}+1)q_{1}). Hence the lemma gives

    𝝂p𝝂((α+1)q1)c𝝂q1/2<(p𝝂(α+1)c𝝂1/2)𝝂q1()\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(({\alpha}+1)q_{1})c_{{\boldsymbol{\nu}}}^{-q_{1}/2}<\infty\qquad\Rightarrow\qquad(p_{\boldsymbol{\nu}}({\alpha}+1)c_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{1}}(\mathcal{F})

    and similarly

    𝝂p𝝂((α+1)q2)d𝝂q2/2<(p𝝂(α+1)d𝝂1/2)𝝂q2().\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(({\alpha}+1)q_{2})d_{{\boldsymbol{\nu}}}^{-q_{2}/2}<\infty\qquad\Rightarrow\qquad(p_{\boldsymbol{\nu}}({\alpha}+1)d_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{2}}(\mathcal{F}).
  4. (iv)

    By Theorem 4.8 and item (ii), for all j0j\in\mathbb{N}_{0}

    C0𝝂c𝝂e~N,𝝂jX2p𝝂(3)𝝂β𝝂(r,ϱ1)e~N,𝝂jX2Cδ2C_{0}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}c_{{\boldsymbol{\nu}}}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}p_{\boldsymbol{\nu}}(3)\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{1})\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}\leq C\delta^{2}

    and

    C0𝝂d𝝂e~N,𝝂jX2p𝝂(3)𝝂β𝝂(r,ϱ2)e~N,𝝂jX2Cδ2𝔴j2α,C_{0}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}d_{{\boldsymbol{\nu}}}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}p_{\boldsymbol{\nu}}(3)\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{2})\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}\leq C\frac{\delta^{2}}{\mathfrak{w}_{j}^{2{\alpha}}},

    with the constant CC independent of jj, 𝔴j\mathfrak{w}_{j} and NN.

  5. (v)

    Because Nmax{jsupp(𝝂):lε,𝝂0}N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,l_{\varepsilon,{\boldsymbol{\nu}}}\geq 0\} and χ0,0=0\chi_{0,0}=0 we have

    𝐈Γj(uuj)=𝐈Γjej=𝐈Γje~Nj\mathbf{I}_{\Gamma_{j}}(u-u^{j})=\mathbf{I}_{\Gamma_{j}}e^{j}=\mathbf{I}_{\Gamma_{j}}\tilde{e}_{N}^{j}

    for all jj\in\mathbb{N} (cp. Remark 4.4). Similarly 𝐈Γju=𝐈Γju~N\mathbf{I}_{\Gamma_{j}}u=\mathbf{I}_{\Gamma_{j}}\tilde{u}_{N} for all jj\in\mathbb{N}.

Step 2. Observe that Γj=\Gamma_{j}=\emptyset for all j>L(ε):=max𝝂lε,𝝂j>L(\varepsilon):=\max_{{\boldsymbol{\nu}}\in\mathcal{F}}{l_{\varepsilon,{\boldsymbol{\nu}}}} (cp. (7.2)), which is finite due to |𝐥ε|<|\mathbf{l}_{\varepsilon}|<\infty. With the conventions 𝐈Γ0=𝐈=Id\mathbf{I}_{\Gamma_{0}}=\mathbf{I}_{\mathcal{F}}={\rm Id} (i.e. 𝐈Γ0\mathbf{I}_{\Gamma_{0}} is the identity) and 𝐈0\mathbf{I}_{\emptyset}\equiv 0 this implies

u=𝐈Γ0u=j=0L(ε)(𝐈Γj𝐈Γj+1)u=(𝐈Γ0𝐈Γ1)u++(𝐈ΓL(ε)1𝐈ΓL(ε))u+𝐈ΓL(ε)u.u=\mathbf{I}_{\Gamma_{0}}u=\sum_{j=0}^{L(\varepsilon)}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})u=(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})u+\dots+(\mathbf{I}_{\Gamma_{L(\varepsilon)-1}}-\mathbf{I}_{\Gamma_{L(\varepsilon)}})u+\mathbf{I}_{\Gamma_{L(\varepsilon)}}u.

By definition of the multilevel interpolant in (7.3)

𝐈ML𝐥εu=j=1L(ε)(𝐈Γj𝐈Γj+1)uj=(𝐈Γ1𝐈Γ2)u1++(𝐈ΓL(ε)1𝐈ΓL(ε))uL(ε)1+𝐈ΓL(ε)uL(ε).\mathbf{I}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u=\sum_{j=1}^{L(\varepsilon)}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})u^{j}=(\mathbf{I}_{\Gamma_{1}}-\mathbf{I}_{\Gamma_{2}})u^{1}+\dots+(\mathbf{I}_{\Gamma_{L(\varepsilon)-1}}-\mathbf{I}_{\Gamma_{L(\varepsilon)}})u^{L(\varepsilon)-1}+\mathbf{I}_{\Gamma_{L(\varepsilon)}}u^{L(\varepsilon)}.

By item (v) of Step 1, we can write

(𝐈Γ0𝐈Γ1)u=u𝐈Γ1u=u𝐈Γ1u~N=(uu~N)+(𝐈Γ0𝐈Γ1)u~N=(uu~N)+(𝐈Γ0𝐈Γ1)e~N0,(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})u=u-\mathbf{I}_{\Gamma_{1}}u=u-\mathbf{I}_{\Gamma_{1}}\tilde{u}_{N}=(u-\tilde{u}_{N})+(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})\tilde{u}_{N}=(u-\tilde{u}_{N})+(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})\tilde{e}_{N}^{0},

where in the last equality we used eN0=uNe_{N}^{0}=u_{N}, by definition of e0=ue^{0}=u (and e~N0=u~NL2(U,X;γ)\tilde{e}_{N}^{0}=\tilde{u}_{N}\in L^{2}(U,X;\gamma) as in Definition 4.1). Hence, again by item (v),

u𝐈ML𝐥εu\displaystyle u-\mathbf{I}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u =(𝐈Γ0𝐈Γ1)u+j=1L(ε)(𝐈Γj𝐈Γj+1)(uuj)\displaystyle=(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})u+\sum_{j=1}^{L(\varepsilon)}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})(u-u^{j})
=(uu~N)+(𝐈Γ0𝐈Γ1)u~N+j=1L(ε)(𝐈Γj𝐈Γj+1)e~Nj\displaystyle=(u-\tilde{u}_{N})+(\mathbf{I}_{\Gamma_{0}}-\mathbf{I}_{\Gamma_{1}})\tilde{u}_{N}+\sum_{j=1}^{L(\varepsilon)}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})\tilde{e}_{N}^{j}
=(uu~N)+j=0L(ε)(𝐈Γj𝐈Γj+1)e~Nj.\displaystyle=(u-\tilde{u}_{N})+\sum_{j=0}^{L(\varepsilon)}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})\tilde{e}_{N}^{j}. (7.21)

We will use this representation to bound the norm u𝐈ML𝐥εuL2(U,X;γ)\|u-\mathbf{I}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u\|_{L^{2}(U,X;\gamma)}. From item (i) of Step 1 it follows that for every j0j\in\mathbb{N}_{0}

e~Nj(𝒚)=𝝂e~N,𝝂jH𝝂(𝒚),\tilde{e}_{N}^{j}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}H_{\boldsymbol{\nu}}({\boldsymbol{y}}),

with the equality and unconditional convergence in the space XX for all 𝒚U{\boldsymbol{y}}\in U. Therefore, by the same argument as in the proof of Lemma 6.11, we can prove that

(𝐈Γj𝐈Γj+1)e~Nj=𝝂e~N,𝝂j(𝐈Γj𝐈Γj+1)H𝝂(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})\tilde{e}_{N}^{j}=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}} (7.22)

with equality and unconditional convergence in the space L2(N,X;γN)L^{2}(\mathbb{R}^{N},X;\gamma_{N}).

Using (7.4) and

(𝐈Γj𝐈Γj+1)H𝝂=0(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}=0

for all 𝝂Γj+1Γj{\boldsymbol{\nu}}\in\Gamma_{j+1}\subseteq\Gamma_{j} by Lemma 6.2, we get

u𝐈ML𝐥εuL2(U,X;γ)uu~NL2(U,X;γ)+𝝂j=lε,𝝂L(ε)e~N,𝝂jX(𝐈Γj𝐈Γj+1)H𝝂L2(U;γ).\|u-\mathbf{I}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u\|_{L^{2}(U,X;\gamma)}\leq\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}+\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}\|(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}. (7.23)

Step 3. We wish to apply Lemma 7.4 to the bound (7.23). By (6.7), we have for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F}

(𝐈Γj𝐈Γj+1)H𝝂L2(U;γ)𝐈ΓjH𝝂L2(U;γ)+𝐈Γj+1H𝝂L2(U;γ)2p𝝂(3).\|(\mathbf{I}_{\Gamma_{j}}-\mathbf{I}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}\leq\|\mathbf{I}_{\Gamma_{j}}H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}+\|\mathbf{I}_{\Gamma_{j+1}}H_{\boldsymbol{\nu}}\|_{L^{2}(U;\gamma)}\leq 2p_{\boldsymbol{\nu}}(3).

Note these inequalities also hold when j=0j=0, that is when 𝐈Γ0=Id\mathbf{I}_{\Gamma_{0}}={\rm Id}. By items (iii) and (iv) of Step 1, the collections (aj,𝝂)𝝂(a_{j,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, j0j\in\mathbb{N}_{0}, and (c𝝂)𝝂(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, (d𝝂)𝝂(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, satisfy the assumptions of Lemma 7.4. Therefore, (7.23), (7.20) and Lemma 7.4 give

u𝐈ML𝐥εnuL2(U,X;γ)nR+𝝂j=lεn,𝝂L(εn)aj,𝝂C(1+logn)nR,\|u-\mathbf{I}^{\rm ML}_{\mathbf{l}_{\varepsilon_{n}}}u\|_{L^{2}(U,X;\gamma)}\leq n^{-R}+\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon_{n},{\boldsymbol{\nu}}}}}^{L(\varepsilon_{n})}a_{j,{\boldsymbol{\nu}}}\leq C(1+\log n)n^{-R},

with the convergence rate

R=min{α,α(q111/2)α+q11q21}=min{α,α(p113/2)α+p11p21},R=\min\left\{{\alpha},\frac{{\alpha}(q_{1}^{-1}-1/2)}{{\alpha}+q_{1}^{-1}-q_{2}^{-1}}\right\}=\min\left\{{\alpha},\frac{{\alpha}(p_{1}^{-1}-3/2)}{{\alpha}+p_{1}^{-1}-p_{2}^{-1}}\right\},

where we used q1=p1/(1p1)q_{1}=p_{1}/(1-p_{1}) and q2=p2/(1p2)q_{2}=p_{2}/(1-p_{2}) as stated in item (ii) of Step 1. ∎

7.5 Multilevel Smolyak sparse-grid quadrature algorithm

We next formulate an analog of Theorem 7.5 for a multilevel Smolyak sparse-grid quadrature algorithm. First, the definition of the multi-index sets in (7.18) (which are used to construct the quadrature via Algorithm 2) has to be slightly adjusted. Then, we state and prove a convergence rate result for the corresponding algorithm. Its proof is along the lines of the proof of Theorem 7.5.

Let 𝒃1=(b1,j)jp1(){\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}}\in\ell^{p_{1}}({\mathbb{N}}), 𝒃2=(b2,j)jp2(){\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}}\in\ell^{p_{2}}({\mathbb{N}}), and ξ\xi be the two sequences and the constant from Assumption 7.2. For two constants K>0K>0 and r>3r>3, which are still at our disposal and which will be defined below, we set for all jj\in\mathbb{N}

ϱ1,j:=b1,jp11ξ4𝒃1p1,ϱ2,j:=b2,jp21ξ4𝒃2p2.\varrho_{1,j}:=b_{1,j}^{p_{1}-1}\frac{\xi}{4\|{\boldsymbol{b}}_{1}\|_{\ell^{p_{1}}}},\qquad\varrho_{2,j}:=b_{2,j}^{p_{2}-1}\frac{\xi}{4\|{\boldsymbol{b}}_{2}\|_{\ell^{p_{2}}}}. (7.24)

Furthermore, we let for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F} (as in Lemma 6.5 for k=2k=2 and with τ=3\tau=3)

c𝝂:=jmax{1,Kϱ1,j}4νjr3,d𝝂:=jmax{1,Kϱ2,j}4νjr3.c_{{\boldsymbol{\nu}}}:=\prod_{j\in\mathbb{N}}\max\{1,K\varrho_{1,j}\}^{4}\nu_{j}^{r-3},\qquad d_{{\boldsymbol{\nu}}}:=\prod_{j\in\mathbb{N}}\max\{1,K\varrho_{2,j}\}^{4}\nu_{j}^{r-3}. (7.25)
Theorem 7.6.

Let uL2(U,X;γ)u\in L^{2}(U,X;\gamma) and ulL2(U,X;γ)u^{l}\in L^{2}(U,X;\gamma), ll\in\mathbb{N}, satisfy Assumption 7.2 with some constants α>0{\alpha}>0 and 0<p1<4/50<p_{1}<4/5 and p1p2<1p_{1}\leq p_{2}<1. Set q1:=p1/(1p1)q_{1}:=p_{1}/(1-p_{1}). Assume that r>2(1+(α+1)q1/2)/q1+3r>2(1+({\alpha}+1)q_{1}/2)/q_{1}+3 (for rr in (7.25)). There exist constants K>0K>0 (in (7.25)) and C>0C>0 such that the following holds.

There exist C>0C>0 and, for every nn\in{\mathbb{N}} there exists εn(0,1]\varepsilon_{n}\in(0,1] such that such that work(𝐥εn)n\mathrm{work}(\mathbf{l}_{\varepsilon_{n}})\leq n and with 𝐥εn=(lεn,𝛎)𝛎\mathbf{l}_{\varepsilon_{n}}=({l_{\varepsilon_{n},{\boldsymbol{\nu}}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} as in Corollary 7.4 (with c𝛎c_{\boldsymbol{\nu}}, d𝛎d_{\boldsymbol{\nu}} as in (7.25)) it holds

Uu(𝒚)dγ(𝒚)𝐐𝐥εnMLuXC(1+logn)nR,\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\mathbf{l}_{\varepsilon_{n}}}^{\rm ML}u\right\|_{X}\leq\ C(1+\log n)n^{-R},

with the convergence rate

R:=min{α,α(2p115/2)α+2p112p21}.R:=\min\left\{{\alpha},\frac{{\alpha}(2p_{1}^{-1}-5/2)}{{\alpha}+2p_{1}^{-1}-2p_{2}^{-1}}\right\}. (7.26)
Proof.

Throughout this proof we write 𝒃1=(b1,j)j{\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}} and 𝒃2=(b2,j)j{\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}} for the two sequences in Assumption 7.2. As in the proof of Theorem 7.5 we highlight that the multi-index set Γj\Gamma_{j} which was defined in (7.2) is downward closed for all j0j\in\mathbb{N}_{0}.

Step 1. Given nn\in\mathbb{N}, we choose ε:=εn\varepsilon:=\varepsilon_{n} as in Lemma 7.4. Fix NN\in\mathbb{N} such that N>max{j:jsupp(𝝂),lε,𝝂>0}N>\max\{j:\,j\in\operatorname{supp}({\boldsymbol{\nu}}),\,l_{\varepsilon,{\boldsymbol{\nu}}}>0\} and so large that

U(u(𝒚)u~N(𝒚))dγ(𝒚)XnR,\left\|\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})\right\|_{X}\leq n^{-R}, (7.27)

where u~N:UX\tilde{u}_{N}:U\to X is as in Definition 4.1 (this is possible due limNuu~NL2(U,X;γ)=0\lim_{N\to\infty}\|u-\tilde{u}_{N}\|_{L^{2}(U,X;\gamma)}=0 which holds by the (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphy of uu).

By Assumption 7.2, for every jj\in\mathbb{N} the function ej:=uujL2(U,X;γ)e^{j}:=u-u^{j}\in L^{2}(U,X;\gamma) is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic and (𝒃2,ξ,δ𝔴jα,X)({\boldsymbol{b}}_{2},\xi,\delta\mathfrak{w}_{j}^{\alpha},X)-holomorphic. For notational convenience we set e0:=u0=uL2(U,X;γ)e^{0}:=u-0=u\in L^{2}(U,X;\gamma), so that e0e^{0} is (𝒃1,ξ,δ,X)({\boldsymbol{b}}_{1},\xi,\delta,X)-holomorphic and (𝒃2,ξ,δ,X)({\boldsymbol{b}}_{2},\xi,\delta,X)-holomorphic. Hence for every j0j\in\mathbb{N}_{0} there exists a function e~Nj=u~Nu~Nj\tilde{e}_{N}^{j}=\tilde{u}_{N}-\tilde{u}_{N}^{j} as in Definition 4.1 (iii).

The following assertions are identical to the ones in the proof of Theorem 7.5, except that we now admit different summability exponents q1q_{1} and q2q_{2}.

  1. (i)

    By Lemma 6.10, for every j0j\in\mathbb{N}_{0}, with the Wiener-Hermite PC expansion coefficients

    e~N,𝝂j:=UH𝝂(𝒚)e~Nj(𝒚)dγ(𝒚),\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}:=\int_{U}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\tilde{e}_{N}^{j}({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}),

    it holds

    e~Nj(𝒚)=𝝂e~N,𝝂jH𝝂(𝒚)𝒚U,\tilde{e}_{N}^{j}({\boldsymbol{y}})=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}H_{\boldsymbol{\nu}}({\boldsymbol{y}})\qquad\forall{\boldsymbol{y}}\in U,

    with pointwise absolute convergence.

  2. (ii)

    By Lemma 6.5, upon choosing K>0K>0 in (7.18) large enough, and because r>3r>3,

    C0c𝝂p𝝂(3)β𝝂(r,ϱ1),C0d𝝂p𝝂(3)β𝝂(r,ϱ2)𝝂2.C_{0}c_{{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(3)\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{1}),\qquad C_{0}d_{{\boldsymbol{\nu}}}p_{\boldsymbol{\nu}}(3)\leq\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{2})\qquad\forall{\boldsymbol{\nu}}\in\mathcal{F}_{2}.

    Remark that by definition of ϱi{\boldsymbol{\varrho}}_{i}, i{1,2}i\in\{1,2\}, in (7.24), it holds ϱi,jbi,j(1pi)\varrho_{i,j}\sim b_{i,j}^{-(1-p_{i})} and therefore (ϱi,j1)jqi()(\varrho_{i,j}^{-1})_{j\in\mathbb{N}}\in\ell^{q_{i}}(\mathbb{N}) with qi:=pi/(1pi)q_{i}:=p_{i}/(1-p_{i}), i{1,2}i\in\{1,2\}.

  3. (iii)

    Due to r>2(1+2(α+1)q1)/q1+3r>2(1+2({\alpha}+1)q_{1})/q_{1}+3, the condition of Lemma 6.6 is satisfied (with k=2k=2, τ=3\tau=3 and θ=(α+1)q1/2\theta=({\alpha}+1)q_{1}/2). Hence the lemma gives

    𝝂p𝝂((α+1)q1/2)c𝝂q1/4<(p𝝂(α+1)c𝝂1/2)𝝂q1/2()\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(({\alpha}+1)q_{1}/2)c_{{\boldsymbol{\nu}}}^{-q_{1}/4}<\infty\qquad\Rightarrow\qquad(p_{\boldsymbol{\nu}}({\alpha}+1)c_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{1}/2}(\mathcal{F})

    and similarly

    𝝂p𝝂((α+1)q2/2)d𝝂q2/4<(p𝝂(α+1)d𝝂1/2)𝝂q2/2().\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}p_{\boldsymbol{\nu}}(({\alpha}+1)q_{2}/2)d_{{\boldsymbol{\nu}}}^{-q_{2}/4}<\infty\qquad\Rightarrow\qquad(p_{\boldsymbol{\nu}}({\alpha}+1)d_{{\boldsymbol{\nu}}}^{-1/2})_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{2}/2}(\mathcal{F}).
  4. (iv)

    By Theorem 4.8 and item (ii), for all j0j\in\mathbb{N}_{0}

    C0𝝂2c𝝂e~N,𝝂jX2p𝝂(3)𝝂2β𝝂(r,ϱ1)e~N,𝝂jX2Cδ2C_{0}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}}c_{{\boldsymbol{\nu}}}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}p_{\boldsymbol{\nu}}(3)\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{1})\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}\leq C\delta^{2}

    and

    C0𝝂2d𝝂e~N,𝝂jX2p𝝂(3)𝝂2β𝝂(r,ϱ2)e~N,𝝂jX2Cδ2𝔴j2α,C_{0}\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}}d_{{\boldsymbol{\nu}}}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}p_{\boldsymbol{\nu}}(3)\leq\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}}\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}}_{2})\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}^{2}\leq C\frac{\delta^{2}}{\mathfrak{w}_{j}^{2{\alpha}}},

    with the constant CC independent of jj, 𝔴j\mathfrak{w}_{j} and NN.

  5. (v)

    Because Nmax{jsupp(𝝂):lε,𝝂0}N\geq\max\{j\in\operatorname{supp}({\boldsymbol{\nu}})\,:\,l_{\varepsilon,{\boldsymbol{\nu}}}\geq 0\} and χ0,0=0\chi_{0,0}=0 we have 𝐐Γj(uuj)=𝐐Γjej=𝐐Γje~Nj\mathbf{Q}_{\Gamma_{j}}(u-u^{j})=\mathbf{Q}_{\Gamma_{j}}e^{j}=\mathbf{Q}_{\Gamma_{j}}\tilde{e}_{N}^{j} for all jj\in\mathbb{N} (cp. Remark 4.4). Similarly 𝐐Γju=𝐐Γju~N\mathbf{Q}_{\Gamma_{j}}u=\mathbf{Q}_{\Gamma_{j}}\tilde{u}_{N} for all jj\in\mathbb{N}.

Step 2. Observe that Γj=\Gamma_{j}=\emptyset for all

j>L(ε):=max𝝂lε,𝝂j>L(\varepsilon):=\max_{{\boldsymbol{\nu}}\in\mathcal{F}}{l_{\varepsilon,{\boldsymbol{\nu}}}}

(cp. (7.2)), which is finite due to |𝐥ε|<|\mathbf{l}_{\varepsilon}|<\infty. With the conventions

𝐐Γ0=𝐐=Udγ(𝒚)\mathbf{Q}_{\Gamma_{0}}=\mathbf{Q}_{\mathcal{F}}=\int_{U}\cdot\,\mathrm{d}\gamma({\boldsymbol{y}})

(i.e. 𝐐Γ0\mathbf{Q}_{\Gamma_{0}} is the exact integral operator) and 𝐐0\mathbf{Q}_{\emptyset}\equiv 0 this implies

Uu(𝒚)dγ(𝒚)\displaystyle\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}}) =𝐐Γ0u=j=0L(ε)(𝐐Γj𝐐Γj+1)u\displaystyle=\mathbf{Q}_{\Gamma_{0}}u=\sum_{j=0}^{L(\varepsilon)}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})u
=(𝐐Γ0𝐐Γ1)u++(𝐐ΓL(ε)1𝐐ΓL(ε))u+𝐐ΓL(ε)u.\displaystyle=(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})u+\ldots+(\mathbf{Q}_{\Gamma_{L(\varepsilon)-1}}-\mathbf{Q}_{\Gamma_{L(\varepsilon)}})u+\mathbf{Q}_{\Gamma_{L(\varepsilon)}}u.

By definition of the multilevel quadrature in (7.4)

𝐐ML𝐥εu\displaystyle\mathbf{Q}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u =j=1L(ε)(𝐐Γj𝐐Γj+1)uj\displaystyle=\sum_{j=1}^{L(\varepsilon)}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})u^{j}
=(𝐐Γ1𝐐Γ2)u1++(𝐐ΓL(ε)1𝐐ΓL(ε))uL(ε)+𝐐ΓL(ε)uL(ε).\displaystyle=(\mathbf{Q}_{\Gamma_{1}}-\mathbf{Q}_{\Gamma_{2}})u^{1}+\ldots+(\mathbf{Q}_{\Gamma_{L(\varepsilon)-1}}-\mathbf{Q}_{\Gamma_{L(\varepsilon)}})u^{L(\varepsilon)}+\mathbf{Q}_{\Gamma_{L(\varepsilon)}}u^{L(\varepsilon)}.

By item (v) of Step 1, we can write

(𝐐Γ0𝐐Γ1)u\displaystyle(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})u =Uu(𝒚)dγ(𝒚)𝐐Γ1u\displaystyle=\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Gamma_{1}}u
=Uu(𝒚)dγ(𝒚)𝐐Γ1u~N\displaystyle=\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\Gamma_{1}}\tilde{u}_{N}
=U(u(𝒚)u~N(𝒚))dγ(𝒚)+(𝐐Γ0𝐐Γ1)u~N\displaystyle=\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})+(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})\tilde{u}_{N}
=U(u(𝒚)u~N(𝒚))dγ(𝒚)+(𝐐Γ0𝐐Γ1)e~N0,\displaystyle=\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})+(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})\tilde{e}_{N}^{0},

where in the last equality we used eN0=uNe_{N}^{0}=u_{N}, by definition of e0=ue^{0}=u (and e~N0=u~NL2(U,X;γ)\tilde{e}_{N}^{0}=\tilde{u}_{N}\in L^{2}(U,X;\gamma) as in Definition 4.1). Hence, again by item (v),

Uu(𝒚)dγ(𝒚)𝐐ML𝐥εu\displaystyle\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u =(𝐐Γ0𝐐Γ1)u+j=1L(ε)(𝐐Γj𝐐Γj+1)(uuj)\displaystyle=(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})u+\sum_{j=1}^{L(\varepsilon)}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})(u-u^{j})
=U(u(𝒚)u~N(𝒚))dγ(𝒚)+(𝐐Γ0𝐐Γ1)u~N+j=1L(ε)(𝐐Γj𝐐Γj+1)e~Nj\displaystyle=\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})+(\mathbf{Q}_{\Gamma_{0}}-\mathbf{Q}_{\Gamma_{1}})\tilde{u}_{N}+\sum_{j=1}^{L(\varepsilon)}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})\tilde{e}_{N}^{j}
=U(u(𝒚)u~N(𝒚))dγ(𝒚)+j=0L(ε)(𝐐Γj𝐐Γj+1)e~Nj.\displaystyle=\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})+\sum_{j=0}^{L(\varepsilon)}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})\tilde{e}_{N}^{j}.

Let us bound the norm. From item (i) of Step 1 it follows that for every j0j\in\mathbb{N}_{0},

e~Nj(𝒚)=𝝂1Ne~N,𝝂jH𝝂(𝒚),\tilde{e}_{N}^{j}({\boldsymbol{y}})=\sum_{{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}H_{\boldsymbol{\nu}}({\boldsymbol{y}}),

with the equality and unconditional convergence in XX for all 𝒚1N{\boldsymbol{y}}\in\mathcal{F}_{1}^{N}. Hence similar to Lemma 6.15 we have

(𝐐Γj𝐐Γj+1)eNj=𝝂1Ne~N,𝝂j(𝐐Γj𝐐Γj+1)H𝝂(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})e_{N}^{j}=\sum_{{{\boldsymbol{\nu}}\in\mathcal{F}_{1}^{N}}}\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}

with the equality and unconditional convergence in XX. Since (𝐐Γj𝐐Γj+1)H𝝂=0X(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}=0\in X for all 𝝂Γj+1Γj{\boldsymbol{\nu}}\in\Gamma_{j+1}\subseteq\Gamma_{j} and all 𝝂\2{\boldsymbol{\nu}}\in\mathcal{F}\backslash\mathcal{F}_{2} by Lemma 6.2, we get

Uu(𝒚)dγ(𝒚)𝐐ML𝐥εuXU(u(𝒚)u~N(𝒚))dγ(𝒚)X+𝝂2j=lε,𝝂L(ε)e~N,𝝂jX|(𝐐Γj𝐐Γj+1)H𝝂|.\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u\right\|_{X}\leq\left\|\int_{U}(u({\boldsymbol{y}})-\tilde{u}_{N}({\boldsymbol{y}}))\,\mathrm{d}\gamma({\boldsymbol{y}})\right\|_{X}+\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{2}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}|(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}|. (7.28)

Step 3. We wish to apply Lemma 7.4 to the bound (7.28). By (6.10), for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F}

|(𝐐Γj𝐐Γj+1)H𝝂||𝐐ΓjH𝝂|+|𝐐Γj+1H𝝂|2p𝝂(3).|(\mathbf{Q}_{\Gamma_{j}}-\mathbf{Q}_{\Gamma_{j+1}})H_{\boldsymbol{\nu}}|\leq|\mathbf{Q}_{\Gamma_{j}}H_{\boldsymbol{\nu}}|+|\mathbf{Q}_{\Gamma_{j+1}}H_{\boldsymbol{\nu}}|\leq 2p_{\boldsymbol{\nu}}(3).

Define

aj,𝝂:=e~N,𝝂jXp𝝂(3)𝝂2,a_{j,{\boldsymbol{\nu}}}:=\|\tilde{e}_{N,{\boldsymbol{\nu}}}^{j}\|_{X}p_{\boldsymbol{\nu}}(3)\qquad\forall{\boldsymbol{\nu}}\in\mathcal{F}_{2},

and aj,𝝂:=0a_{j,{\boldsymbol{\nu}}}:=0 for 𝝂\2{\boldsymbol{\nu}}\in\mathcal{F}\backslash\mathcal{F}_{2}. By items (iii) and (iv) of Step 1, the collections (aj,𝝂)𝝂(a_{j,{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, j0j\in\mathbb{N}_{0}, and (c𝝂)𝝂(c_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, (d𝝂)𝝂(d_{{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, satisfy the assumptions of Lemma 7.4 (with q~1:=q1/2\tilde{q}_{1}:=q_{1}/2 and q~2:=q2/2\tilde{q}_{2}:=q_{2}/2). Therefore (7.28), (7.27) and Lemma 7.4 give

Uu(𝒚)dγ(𝒚)𝐐ML𝐥εuXnR+𝝂j=lε,𝝂L(ε)aj,𝝂C(1+logn)nR,\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}^{\rm ML}_{\mathbf{l}_{\varepsilon}}u\right\|_{X}\leq n^{-R}+\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}\sum_{j={l_{\varepsilon,{\boldsymbol{\nu}}}}}^{L(\varepsilon)}a_{j,{\boldsymbol{\nu}}}\leq C(1+\log n)n^{-R},

with

R=min{α,α(q~111/2)α+q~11q~21}=min{α,α(2p115/2)α+2p112p21},R=\min\left\{{\alpha},\frac{{\alpha}(\tilde{q}_{1}^{-1}-1/2)}{{\alpha}+\tilde{q}_{1}^{-1}-\tilde{q}_{2}^{-1}}\right\}=\min\left\{{\alpha},\frac{{\alpha}(2p_{1}^{-1}-5/2)}{{\alpha}+2p_{1}^{-1}-2p_{2}^{-1}}\right\},

where we used q~1=q1/2=p1/(22p1)\tilde{q}_{1}=q_{1}/2=p_{1}/(2-2p_{1}) and q~2=q2/2=p2/(22p2)\tilde{q}_{2}=q_{2}/2=p_{2}/(2-2p_{2}) as stated in item (ii) of Step 1. ∎

7.6 Examples for multilevel interpolation and quadrature

We revisit the examples in Sections 4 and 5, and demonstrate how to verify the assumptions required for the multilevel convergence rate results in Theorems 7.5 and 7.6.

7.6.1 Parametric diffusion coefficient in polygonal domain

Let D2{{D}}\subseteq\mathbb{R}^{2} be a bounded polygonal domain, and consider once more the elliptic equation

div(a𝒰(a))=fin D,𝒰(a)=0on D,-\operatorname{div}(a\nabla{\mathcal{U}}(a))=f\quad\text{in }{{D}},\qquad{\mathcal{U}}(a)=0\quad\text{on }\partial{{D}}, (7.29)

as in Section 4.3.1.

For s0s\in\mathbb{N}_{0} and ϰ\varkappa\in\mathbb{R}, recall the Kondrat’ev spaces 𝒲s(D){\mathcal{W}}^{s}_{\infty}({{D}}) and 𝒦sϰ(D){\mathcal{K}}^{s}_{\varkappa}({{D}}) with norms

u𝒦sϰ:=|𝜶|srD|𝜶|ϰD𝜶uL2andu𝒲s:=|𝜶|srD|𝜶|D𝜶uL\|u\|_{{\mathcal{K}}^{s}_{\varkappa}}:=\sum_{|{\boldsymbol{\alpha}}|\leq s}\|r_{D}^{|{\boldsymbol{\alpha}}|-\varkappa}D^{\boldsymbol{\alpha}}u\|_{L^{2}}\qquad\text{and}\qquad\|u\|_{{\mathcal{W}}^{s}_{\infty}}:=\sum_{|{\boldsymbol{\alpha}}|\leq s}\|r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}u\|_{L^{\infty}}

introduced in Section 3.8.1. Here, as earlier, rD:D[0,1]r_{{{D}}}:{{D}}\to[0,1] denotes a fixed smooth function that coincides with the distance to the nearest corner, in a neighbourhood of each corner. According to Theorem 3.29, assuming s2s\geq 2, f𝒦s2ϰ1(D)f\in{\mathcal{K}}^{s-2}_{\varkappa-1}({{D}}) and a𝒲s1(D)a\in{\mathcal{W}}^{s-1}_{\infty}({{D}}) the solution 𝒰(a){\mathcal{U}}(a) of (7.29) belongs to 𝒦ϰ+1s(D){\mathcal{K}}_{\varkappa+1}^{s}({{D}}) provided that with

ρ(a):=essinf𝒙D(a(𝒙))>0,\rho(a):=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,\Re(a({\boldsymbol{x}}))>0,
|ϰ|<ρ(a)τaL,|\varkappa|<\frac{\rho(a)}{\tau\|a\|_{L^{\infty}}}, (7.30)

where τ\tau is a constant depending on D{{D}} and ss. Our goal is to treat, in a unified manner, a family of diffusion coefficients a(𝒚)a({\boldsymbol{y}}), 𝒚U{\boldsymbol{y}}\in U, where for certain 𝒚U{\boldsymbol{y}}\in U the diffusion coefficient a(𝒚)a({\boldsymbol{y}}) is such that the right-hand side of (7.30) might be arbitrarily small. This only leaves us with the choice ϰ=0\varkappa=0, see Remark 3.31. On the other hand, the motivation of using Kondrat’ev spaces in the analysis of approximations to PDE solutions 𝒰(a(𝒚)){\mathcal{U}}(a({\boldsymbol{y}})), is that functions in 𝒦ϰ+1s(D){\mathcal{K}}_{\varkappa+1}^{s}({{D}}) on polygonal domains in 2\mathbb{R}^{2} can be approximated with the optimal convergence rate s12\frac{s-1}{2} w.r.t. the H1H^{1}-norm by suitable finite element spaces (on graded meshes; i.e. this analysis accounts for corner singularities which prevent optimal convergence rates on uniform meshes). Such results are well-known, see for example [25], however they require ϰ>0\varkappa>0. For this reason we need a stronger regularity result, giving uniform 𝒦ϰ+1s{\mathcal{K}}_{\varkappa+1}^{s}-regularity with ϰ>0\varkappa>0 independent of the parameter. This is the purpose of the next theorem. For its proof we shall need the following lemma, which is shown in a similar way as in [113, Lemma C.2]. We recall that

fWs:=|𝝂|sD𝝂fL.\|f\|_{W^{s}_{\infty}}:=\sum_{|{\boldsymbol{\nu}}|\leq s}\|D^{\boldsymbol{\nu}}f\|_{L^{\infty}}.
Lemma 7.7.

Let s0s\in\mathbb{N}_{0} and let D2{{D}}\subseteq\mathbb{R}^{2} be a bounded polygonal domain, dd\in\mathbb{N}.

Then there exist CsC_{s} and C~s\tilde{C}_{s} such that for any two functions ff, g𝒲s(D)g\in{\mathcal{W}}^{s}_{\infty}({{D}})

  1. (i)

    fg𝒲sCsf𝒲sg𝒲s\|fg\|_{{\mathcal{W}}^{s}_{\infty}}\leq C_{s}\|f\|_{{\mathcal{W}}^{s}_{\infty}}\|g\|_{{\mathcal{W}}^{s}_{\infty}},

  2. (ii)

    1f𝒲sC~sf𝒲ssessinf𝒙D|f(𝒙)|s+1\|\frac{1}{f}\|_{{\mathcal{W}}^{s}_{\infty}}\leq\tilde{C}_{s}\frac{\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{s}}{\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{{D}}}|f({\boldsymbol{x}})|^{s+1}} if essinf𝒙D|f(𝒙)|>0\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{{D}}}|f({\boldsymbol{x}})|>0.

These statements remain true if 𝒲s(D){\mathcal{W}}^{s}_{\infty}({{D}}) is replaced by Ws(D)W^{s}_{\infty}({{D}}). Furthermore, if ϰ\varkappa\in\mathbb{R}, then for f𝒦ϰs(D)f\in{\mathcal{K}}_{\varkappa}^{s}({{D}}) and a𝒲s(D)a\in{\mathcal{W}}^{s}_{\infty}({{D}})

  1. (iii)

    fa𝒦ϰsCsf𝒦ϰsa𝒲s\|fa\|_{{\mathcal{K}}_{\varkappa}^{s}}\leq C_{s}\|f\|_{{\mathcal{K}}_{\varkappa}^{s}}\|a\|_{{\mathcal{W}}^{s}_{\infty}},

  2. (iv)

    fa𝒦ϰ1s1Cs1f𝒦ϰ+1sa𝒲s\|\nabla f\cdot\nabla a\|_{{\mathcal{K}}_{\varkappa-1}^{s-1}}\leq C_{s-1}\|f\|_{{\mathcal{K}}_{\varkappa+1}^{s}}\|a\|_{{\mathcal{W}}^{s}_{\infty}} if s1s\geq 1.

Proof.

We will only prove (i) and (ii) for functions in 𝒲s(D){\mathcal{W}}^{s}_{\infty}({{D}}). The case of Ws(D)W^{s}_{\infty}({{D}}) is shown similarly (by omitting all occurring functions rDr_{{D}} in the following).

Step 1. We start with (i), and show a slightly more general bound: for τ\tau\in\mathbb{R} introduce

f𝒲sτ,:=|𝝂|srDτ+|𝝂|D𝝂fL,\|f\|_{{\mathcal{W}}^{s}_{\tau,\infty}}:=\sum_{|{\boldsymbol{\nu}}|\leq s}\|r_{{D}}^{\tau+|{\boldsymbol{\nu}}|}D^{\boldsymbol{\nu}}f\|_{L^{\infty}},

i.e. 𝒲s0,(D)=𝒲s(D){\mathcal{W}}^{s}_{0,\infty}({{D}})={\mathcal{W}}^{s}_{\infty}({{D}}). We will show that for τ1+τ2=τ\tau_{1}+\tau_{2}=\tau

fg𝒲sτ,Csf𝒲sτ1,g𝒲sτ2,.\|fg\|_{{\mathcal{W}}^{s}_{\tau,\infty}}\leq C_{s}\|f\|_{{\mathcal{W}}^{s}_{\tau_{1},\infty}}\|g\|_{{\mathcal{W}}^{s}_{\tau_{2},\infty}}. (7.31)

Item (i) then follows with τ=τ1=τ2=0\tau=\tau_{1}=\tau_{2}=0.

Using the multivariate Leibniz rule for Lipschitz functions, for any multiindex 𝝂0d{\boldsymbol{\nu}}\in\mathbb{N}_{0}^{d} with dd\in\mathbb{N} fixed,

D𝝂(fg)=𝝁𝝂(𝝂𝝁)D𝝂𝝁fD𝝁g.D^{\boldsymbol{\nu}}(fg)=\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}D^{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}}fD^{{\boldsymbol{\mu}}}g. (7.32)

Thus if |𝝂|s|{\boldsymbol{\nu}}|\leq s

rDτ+|𝝂|D𝝂(fg)L𝝁𝝂(𝝂𝝁)rDτ1+|𝝂𝝁|D𝝂𝝁fLrDτ2+|𝝁|D𝝁gL2|𝝂|f𝒲sτ1,g𝒲sτ2,,\|r_{{D}}^{\tau+|{\boldsymbol{\nu}}|}D^{\boldsymbol{\nu}}(fg)\|_{L^{\infty}}\leq\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}\|r_{{D}}^{\tau_{1}+|{\boldsymbol{\nu}}-{\boldsymbol{\mu}}|}D^{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}}f\|_{L^{\infty}}\|r_{{D}}^{\tau_{2}+|{\boldsymbol{\mu}}|}D^{{\boldsymbol{\mu}}}g\|_{L^{\infty}}\leq 2^{|{\boldsymbol{\nu}}|}\|f\|_{{\mathcal{W}}^{s}_{\tau_{1},\infty}}\|g\|_{{\mathcal{W}}^{s}_{\tau_{2},\infty}},

where we used (𝝂𝝁)=j=1d(νjμj)\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}=\prod_{j=1}^{d}\binom{\nu_{j}}{\mu_{j}} and i=0νj(νji)=2νj\sum_{i=0}^{\nu_{j}}\binom{\nu_{j}}{i}=2^{\nu_{j}}. We conclude

fg𝒲sτ,CsfWsτ,gWs\|fg\|_{{\mathcal{W}}^{s}_{\tau,\infty}}\leq C_{s}\|f\|_{W^{s}_{\tau,\infty}}\|g\|_{W^{s}_{\infty}}

with Cs=|𝝂|s2|𝝂|C_{s}=\sum_{|{\boldsymbol{\nu}}|\leq s}2^{|{\boldsymbol{\nu}}|}. Hence (i) holds.

Step 2. We show (ii), and claim that for all |𝝂|s|{\boldsymbol{\nu}}|\leq s it holds

D𝝂(1f)=p𝝂f|𝝂|+1,D^{\boldsymbol{\nu}}\left(\frac{1}{f}\right)=\frac{p_{\boldsymbol{\nu}}}{f^{|{\boldsymbol{\nu}}|+1}}\,, (7.33)

where p𝝂p_{\boldsymbol{\nu}} satisfies

p𝝂𝒲s|𝝂||𝝂|,C^|𝝂|f𝒲s|𝝂|\|p_{\boldsymbol{\nu}}\|_{{\mathcal{W}}^{s-|{\boldsymbol{\nu}}|}_{|{\boldsymbol{\nu}}|,\infty}}\leq\hat{C}_{|{\boldsymbol{\nu}}|}\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{|{\boldsymbol{\nu}}|} (7.34)

for some C^|𝝂|\hat{C}_{|{\boldsymbol{\nu}}|} solely depending on |𝝂||{\boldsymbol{\nu}}|. We proceed by induction over |𝝂||{\boldsymbol{\nu}}| and start with |𝝂|=1|{\boldsymbol{\nu}}|=1, i.e., 𝝂=𝒆j=(δij)i=1d{\boldsymbol{\nu}}={\boldsymbol{e}}_{j}=(\delta_{ij})_{i=1}^{d} for some j{1,,d}j\in\{1,\dots,d\}. Then D𝒆j1f=jff2D^{{\boldsymbol{e}}_{j}}\frac{1}{f}=\frac{-\partial_{j}f}{f^{2}} and p𝒆j=jfp_{{\boldsymbol{e}}_{j}}=-\partial_{j}f satisfies

p𝒆j𝒲s11,=|𝝁|s1rD1+|𝝁|D𝝁p𝒆jL=|𝝁|s1rD|𝝁+𝒆j|D𝝁+𝒆jfLf𝒲s,\|p_{{\boldsymbol{e}}_{j}}\|_{{\mathcal{W}}^{s-1}_{1,\infty}}=\sum_{|{\boldsymbol{\mu}}|\leq s-1}\|r_{{D}}^{1+|{\boldsymbol{\mu}}|}D^{{\boldsymbol{\mu}}}p_{{\boldsymbol{e}}_{j}}\|_{L^{\infty}}=\sum_{|{\boldsymbol{\mu}}|\leq s-1}\|r_{{D}}^{|{\boldsymbol{\mu}}+{\boldsymbol{e}}_{j}|}D^{{\boldsymbol{\mu}}+{\boldsymbol{e}}_{j}}f\|_{L^{\infty}}\leq\|f\|_{{\mathcal{W}}^{s}_{\infty}},

i.e. C^1=1\hat{C}_{1}=1. For the induction step fix 𝝂{\boldsymbol{\nu}} with 1<|𝝂|<s1<|{\boldsymbol{\nu}}|<s and j{1,,d}j\in\{1,\dots,d\}. Then by the induction hypothesis D𝝂1f=p𝝂f|𝝂|+1D^{\boldsymbol{\nu}}\frac{1}{f}=\frac{p_{\boldsymbol{\nu}}}{f^{|{\boldsymbol{\nu}}|+1}} and

D𝝂+𝒆j1f=j(p𝝂f|𝝂|+1)=f|𝝂|+1jp𝝂(|𝝂|+1)f|𝝂|p𝝂jff2|𝝂|+2=fjp𝝂(|𝝂|+1)p𝝂jff|𝝂|+2,D^{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{j}}\frac{1}{f}=\partial_{j}\left(\frac{p_{{\boldsymbol{\nu}}}}{f^{|{\boldsymbol{\nu}}|+1}}\right)=\frac{f^{|{\boldsymbol{\nu}}|+1}\partial_{j}p_{{\boldsymbol{\nu}}}-(|{\boldsymbol{\nu}}|+1)f^{|{\boldsymbol{\nu}}|}p_{{\boldsymbol{\nu}}}\partial_{j}f}{f^{2|{\boldsymbol{\nu}}|+2}}=\frac{f\partial_{j}p_{{\boldsymbol{\nu}}}-(|{\boldsymbol{\nu}}|+1)p_{{\boldsymbol{\nu}}}\partial_{j}f}{f^{|{\boldsymbol{\nu}}|+2}},

and thus

p𝝂+𝒆j:=fjp𝝂(|𝝂|+1)p𝝂jf.p_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{j}}:=f\partial_{j}p_{{\boldsymbol{\nu}}}-(|{\boldsymbol{\nu}}|+1)p_{{\boldsymbol{\nu}}}\partial_{j}f.

Observe that

jg𝒲sτ,=|𝝁|srDτ+|𝝁|D𝝁+𝒆jgL|𝝁|s+1rDτ+|𝝁|1D𝝁gL=g𝒲s+1τ1,.\|\partial_{j}g\|_{{\mathcal{W}}^{s}_{\tau,\infty}}=\sum_{|{\boldsymbol{\mu}}|\leq s}\|r_{{D}}^{\tau+|{\boldsymbol{\mu}}|}D^{{\boldsymbol{\mu}}+{\boldsymbol{e}}_{j}}g\|_{L^{\infty}}\leq\sum_{|{\boldsymbol{\mu}}|\leq s+1}\|r_{{D}}^{\tau+|{\boldsymbol{\mu}}|-1}D^{\boldsymbol{\mu}}g\|_{L^{\infty}}=\|g\|_{{\mathcal{W}}^{s+1}_{\tau-1,\infty}}. (7.35)

Using (7.31) and (7.35), we get with τ:=|𝝂|+1\tau:=|{\boldsymbol{\nu}}|+1

p𝝂+𝒆j𝒲sττ,\displaystyle\|p_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{j}}\|_{{\mathcal{W}}^{s-\tau}_{\tau,\infty}} fjp𝝂𝒲sττ,+(|𝝂|+1)p𝝂jf𝒲sττ,\displaystyle\leq\|f\partial_{j}p_{{\boldsymbol{\nu}}}\|_{{\mathcal{W}}^{s-\tau}_{\tau,\infty}}+(|{\boldsymbol{\nu}}|+1)\|p_{{\boldsymbol{\nu}}}\partial_{j}f\|_{{\mathcal{W}}^{s-\tau}_{\tau,\infty}}
Csτf𝒲sτ0,jp𝝂𝒲sττ,+(|𝝂|+1)Csτp𝝂𝒲sττ1,jf𝒲sτ1,\displaystyle\leq C_{s-\tau}\|f\|_{{\mathcal{W}}^{s-\tau}_{0,\infty}}\|\partial_{j}p_{{\boldsymbol{\nu}}}\|_{{\mathcal{W}}^{s-\tau}_{\tau,\infty}}+(|{\boldsymbol{\nu}}|+1)C_{s-\tau}\|p_{{\boldsymbol{\nu}}}\|_{{\mathcal{W}}^{s-\tau}_{\tau-1,\infty}}\|\partial_{j}f\|_{{\mathcal{W}}^{s-\tau}_{1,\infty}}
Csτf𝒲sτ0,p𝝂𝒲sτ+1τ1,+(|𝝂|+1)Csτp𝝂𝒲sτ+1τ1,f𝒲sτ+10,.\displaystyle\leq C_{s-\tau}\|f\|_{{\mathcal{W}}^{s-\tau}_{0,\infty}}\|p_{{\boldsymbol{\nu}}}\|_{{\mathcal{W}}^{s-\tau+1}_{\tau-1,\infty}}+(|{\boldsymbol{\nu}}|+1)C_{s-\tau}\|p_{{\boldsymbol{\nu}}}\|_{{\mathcal{W}}^{s-\tau+1}_{\tau-1,\infty}}\|f\|_{{\mathcal{W}}^{s-\tau+1}_{0,\infty}}.

Due to τ1=|𝝂|\tau-1=|{\boldsymbol{\nu}}| and the induction hypothesis (7.34) for p𝝂p_{\boldsymbol{\nu}},

p𝝂+𝒆j𝒲s(|𝝂|+1)|𝝂|+1,\displaystyle\|p_{{\boldsymbol{\nu}}+{\boldsymbol{e}}_{j}}\|_{{\mathcal{W}}^{s-(|{\boldsymbol{\nu}}|+1)}_{|{\boldsymbol{\nu}}|+1,\infty}} Cs(|𝝂|+1)(C^|𝝂|f𝒲s(|𝝂|+1)f𝒲s|𝝂|+(|𝝂|+1)C^|𝝂|f𝒲s|𝝂|f𝒲s|𝝂|)\displaystyle\leq C_{s-(|{\boldsymbol{\nu}}|+1)}\left(\hat{C}_{|{\boldsymbol{\nu}}|}\|f\|_{{\mathcal{W}}^{s-(|{\boldsymbol{\nu}}|+1)}_{\infty}}\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{|{\boldsymbol{\nu}}|}+(|{\boldsymbol{\nu}}|+1)\hat{C}_{|{\boldsymbol{\nu}}|}\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{|{\boldsymbol{\nu}}|}\|f\|_{{\mathcal{W}}^{s-|{\boldsymbol{\nu}}|}_{\infty}}\right)
Cs(|𝝂|+1)C^|𝝂|(|𝝂|+2)f𝒲s|𝝂|+1.\displaystyle\leq C_{s-(|{\boldsymbol{\nu}}|+1)}\hat{C}_{|{\boldsymbol{\nu}}|}(|{\boldsymbol{\nu}}|+2)\|f\|_{{\mathcal{W}}_{s}^{\infty}}^{|{\boldsymbol{\nu}}|+1}.

In all this shows the claim with C^1:=1\hat{C}_{1}:=1 and inductively for 1<ks1<k\leq s,

C^k:=CskC^k1(k+1).\hat{C}_{k}:=C_{s-k}\hat{C}_{k-1}(k+1).

By (7.33) and (7.34), for every |𝝂|s|{\boldsymbol{\nu}}|\leq s

rD|𝝂|D𝝂(1f)LC^|𝝂|f𝒲s|𝝂|essinf𝒙D|f(𝒙)||𝝂|+1.\left\|r_{{D}}^{|{\boldsymbol{\nu}}|}D^{\boldsymbol{\nu}}\left(\frac{1}{f}\right)\right\|_{L^{\infty}}\leq\hat{C}_{|{\boldsymbol{\nu}}|}\frac{\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{|{\boldsymbol{\nu}}|}}{\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{{D}}}|f({\boldsymbol{x}})|^{|{\boldsymbol{\nu}}|+1}}\,.

Due to

f𝒲sfLessinf𝒙D|f(𝒙)|,\|f\|_{{\mathcal{W}}^{s}_{\infty}}\geq\|f\|_{L^{\infty}}\geq\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{{D}}}|f({\boldsymbol{x}})|,

this implies

1f𝒲s=|𝝂|srD|𝝂|D𝝂(1f)LC~sf𝒲ssessinf𝒙D|f(𝒙)|s+1\left\|\frac{1}{f}\right\|_{{\mathcal{W}}^{s}_{\infty}}=\sum_{|{\boldsymbol{\nu}}|\leq s}\left\|r_{{D}}^{|{\boldsymbol{\nu}}|}D^{\boldsymbol{\nu}}\left(\frac{1}{f}\right)\right\|_{L^{\infty}}\leq\tilde{C}_{s}\frac{\|f\|_{{\mathcal{W}}^{s}_{\infty}}^{s}}{\operatorname{ess\,inf}_{{\boldsymbol{x}}\in{{D}}}|f({\boldsymbol{x}})|^{s+1}}

with C~s:=|𝝂|sC^|𝝂|\tilde{C}_{s}:=\sum_{|{\boldsymbol{\nu}}|\leq s}\hat{C}_{|{\boldsymbol{\nu}}|}.

Step 3. We show (iii) and (iv). If f𝒦ϰs(D)f\in{\mathcal{K}}_{\varkappa}^{s}({{D}}) and a𝒲s(D)a\in{\mathcal{W}}^{s}_{\infty}({{D}}), then by (7.32) for Sobolev functions,

rD𝝂ϰD𝝂(fa)=𝝁𝝂(𝝂𝝁)(rD|𝝂𝝁|ϰD𝝂𝝁f)(rD|𝝁|D𝝁a)r_{{D}}^{{\boldsymbol{\nu}}-\varkappa}D^{\boldsymbol{\nu}}(fa)=\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}(r_{{D}}^{|{\boldsymbol{\nu}}-{\boldsymbol{\mu}}|-\varkappa}D^{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}}f)(r_{{D}}^{|{\boldsymbol{\mu}}|}D^{\boldsymbol{\mu}}a)

and hence

fa𝒦ϰs\displaystyle\|fa\|_{{\mathcal{K}}_{\varkappa}^{s}} =|𝝂|srD|𝝂|ϰD𝝂(fa)L2\displaystyle=\sum_{|{\boldsymbol{\nu}}|\leq s}\|r_{{D}}^{|{\boldsymbol{\nu}}|-\varkappa}D^{\boldsymbol{\nu}}(fa)\|_{L^{2}}
|𝝂|s𝝁𝝂(𝝂𝝁)rD|𝝂𝝁|ϰD𝝂𝝁fL2rD|𝝁|D𝝁aL\displaystyle\leq\sum_{|{\boldsymbol{\nu}}|\leq s}\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}\|r_{{D}}^{|{\boldsymbol{\nu}}-{\boldsymbol{\mu}}|-\varkappa}D^{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}}f\|_{L^{2}}\|r_{{D}}^{|{\boldsymbol{\mu}}|}D^{\boldsymbol{\mu}}a\|_{L^{\infty}}
Cs|𝝂|srD|𝝂|ϰD𝝂fL2|𝝁|srD|𝝁|D𝝁aL\displaystyle\leq C_{s}\sum_{|{\boldsymbol{\nu}}|\leq s}\|r_{{D}}^{|{\boldsymbol{\nu}}|-\varkappa}D^{{\boldsymbol{\nu}}}f\|_{L^{2}}\sum_{|{\boldsymbol{\mu}}|\leq s}\|r_{{D}}^{|{\boldsymbol{\mu}}|}D^{\boldsymbol{\mu}}a\|_{L^{\infty}}
=Csf𝒦ϰsa𝒲s.\displaystyle=C_{s}\|f\|_{{\mathcal{K}}_{\varkappa}^{s}}\|a\|_{{\mathcal{W}}^{s}_{\infty}}.

Finally if s1s\geq 1,

fa𝒦ϰ1s1\displaystyle\|\nabla f\cdot\nabla a\|_{{\mathcal{K}}_{\varkappa-1}^{s-1}} =|𝝂|s1rD|𝝂|ϰ+1D𝝂(j=1djfja)L2\displaystyle=\sum_{|{\boldsymbol{\nu}}|\leq s-1}\left\|r_{{D}}^{|{\boldsymbol{\nu}}|-\varkappa+1}D^{\boldsymbol{\nu}}\left(\sum_{j=1}^{d}\partial_{j}f\partial_{j}a\right)\right\|_{L^{2}}
|𝝂|s1𝝁𝝂(𝝂𝝁)j=1drD|𝝂𝝁|ϰD𝝂𝝁+𝒆jfL2rD|𝝁|+1D𝝁+𝒆jaL\displaystyle\leq\sum_{|{\boldsymbol{\nu}}|\leq s-1}\sum_{{\boldsymbol{\mu}}\leq{\boldsymbol{\nu}}}\binom{{\boldsymbol{\nu}}}{{\boldsymbol{\mu}}}\sum_{j=1}^{d}\|r_{{D}}^{|{\boldsymbol{\nu}}-{\boldsymbol{\mu}}|-\varkappa}D^{{\boldsymbol{\nu}}-{\boldsymbol{\mu}}+{\boldsymbol{e}}_{j}}f\|_{L^{2}}\|r_{{D}}^{|{\boldsymbol{\mu}}|+1}D^{{\boldsymbol{\mu}}+{\boldsymbol{e}}_{j}}a\|_{L^{\infty}}
Cs1d|𝝂|srD|𝝂|ϰ1D𝝂fL2|𝝁|srD|𝝁|D𝝁aL\displaystyle\leq C_{s-1}d\sum_{|{\boldsymbol{\nu}}|\leq s}\|r_{{D}}^{|{\boldsymbol{\nu}}|-\varkappa-1}D^{{\boldsymbol{\nu}}}f\|_{L^{2}}\sum_{|{\boldsymbol{\mu}}|\leq s}\|r_{{D}}^{|{\boldsymbol{\mu}}|}D^{\boldsymbol{\mu}}a\|_{L^{\infty}}
=Cs1df𝒦ϰ+1sa𝒲s.\displaystyle=C_{s-1}d\|f\|_{{\mathcal{K}}_{\varkappa+1}^{s}}\|a\|_{{\mathcal{W}}^{s}_{\infty}}.\qed

The proof of the next theorem is based on Theorem 3.29. In order to get regularity in 𝒦ϰ+1s(D){\mathcal{K}}_{\varkappa+1}^{s}({{D}}) with ϰ>0\varkappa>0 independent of the diffusion coefficient aa, we now assume aW1(D)𝒲s1(D)a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}_{\infty}^{s-1}({{D}}) in lieu of the weaker assumption a𝒲s1a\in{\mathcal{W}}^{s-1}_{\infty} that was required in Theorem 3.29.

Theorem 7.8.

Let D2{{D}}\subseteq\mathbb{R}^{2} be a bounded polygonal domain and ss\in\mathbb{N}, s2s\geq 2. Then there exist ϰ>0\varkappa>0 and Cs>0C_{s}>0 depending on D{{D}} and ss (but independent of aa) such that for all aW1(D)𝒲s1(D)a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}_{\infty}^{s-1}({{D}}) and all f𝒦ϰ1s2(D)f\in{\mathcal{K}}_{\varkappa-1}^{s-2}({{D}}) the weak solution 𝒰H01(D){\mathcal{U}}\in H_{0}^{1}({{D}}) of (7.29) satisfies with Ns:=s(s1)2N_{s}:=\frac{s(s-1)}{2}

𝒰𝒦ϰ+1sCs1ρ(a)(a𝒲s1+aW1ρ(a))Nsf𝒦ϰ1s2.\|{\mathcal{U}}\|_{{\mathcal{K}}_{\varkappa+1}^{s}}\leq C_{s}\frac{1}{\rho(a)}\left(\frac{\|a\|_{{\mathcal{W}}^{s-1}_{\infty}}+\|a\|_{W^{1}_{\infty}}}{\rho(a)}\right)^{N_{s}}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}. (7.36)
Proof.

Throughout this proof let ϰ(0,1)\varkappa\in(0,1) be a constant such that

Δ:𝒦ϰ+1j(D)H01(D)𝒦ϰ1j2(D)-\Delta:{\mathcal{K}}_{\varkappa+1}^{j}({{D}})\cap H_{0}^{1}({{D}})\to{\mathcal{K}}_{\varkappa-1}^{j-2}({{D}}) (7.37)

is a boundedly invertible operator for all j{2,,s}j\in\{2,\dots,s\}; such ϰ\varkappa exists by Theorem 3.29, and ϰ\varkappa merely depends on D{{D}} and ss.

Step 1. We prove the theorem for s=2s=2, in which case aW1𝒲1=W1a\in W^{1}_{\infty}\cap{\mathcal{W}}^{1}_{\infty}=W^{1}_{\infty}.

Applying Theorem 3.29 directly to (7.29) yields the existence of some ϰ~(0,ϰ)\tilde{\varkappa}\in(0,\varkappa) (depending on aa) such that 𝒰𝒦ϰ~+12{\mathcal{U}}\in{\mathcal{K}}_{\tilde{\varkappa}+1}^{2}. Here we use

f𝒦ϰ10(D)𝒦ϰ~10(D)f\in{\mathcal{K}}_{\varkappa-1}^{0}({{D}})\hookrightarrow{\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}})

due to ϰ~(0,ϰ)\tilde{\varkappa}\in(0,\varkappa). By the Leibniz rule for Sobolev functions we can write

div(a𝒰)=aΔ𝒰a𝒰-\operatorname{div}(a\nabla{\mathcal{U}})=-a\Delta{\mathcal{U}}-\nabla a\cdot\nabla{\mathcal{U}}

in the sense of 𝒦ϰ~10(D){\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}}): (i) it holds Δ𝒰𝒦ϰ~10(D)\Delta{\mathcal{U}}\in{\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}}) and

aW1(D)L(D)a\in W^{1}_{\infty}({{D}})\hookrightarrow L^{\infty}({{D}})

which implies aΔ𝒰𝒦ϰ~10(D)a\Delta{\mathcal{U}}\in{\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}}) (ii) it holds

𝒰𝒦ϰ~1(D)𝒦ϰ~0(D),\nabla{\mathcal{U}}\in{\mathcal{K}}_{\tilde{\varkappa}}^{1}({{D}})\hookrightarrow{\mathcal{K}}_{\tilde{\varkappa}}^{0}({{D}}),

and aL(D)\nabla a\in L^{\infty}({{D}}) which implies a𝒰𝒦ϰ~10(D)\nabla a\cdot\nabla{\mathcal{U}}\in{\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}}). Hence,

div(a𝒰)=aΔ𝒰a𝒰=f,-\operatorname{div}(a\nabla{\mathcal{U}})=-a\Delta{\mathcal{U}}-\nabla a\cdot\nabla{\mathcal{U}}=f,

and further

Δ𝒰=1a(f+a𝒰)=:f~𝒦ϰ~10(D)-\Delta{\mathcal{U}}=\frac{1}{a}\Big{(}f+\nabla a\cdot\nabla{\mathcal{U}}\Big{)}=:\tilde{f}\in{\mathcal{K}}_{\tilde{\varkappa}-1}^{0}({{D}})

since 1aL(D)\frac{1}{a}\in L^{\infty}({{D}}) due to ρ(a)>0\rho(a)>0. Our goal is to show that in fact f~𝒦ϰ10(D)\tilde{f}\in{\mathcal{K}}_{\varkappa-1}^{0}({{D}}). Because of Δ𝒰=f~-\Delta{\mathcal{U}}=\tilde{f} and 𝒰|D0{\mathcal{U}}|_{\partial{{D}}}\equiv 0, Theorem 3.29 then implies

𝒰𝒦ϰ+12Cf~𝒦ϰ10\|{\mathcal{U}}\|_{{\mathcal{K}}_{\varkappa+1}^{2}}\leq C\|\tilde{f}\|_{{\mathcal{K}}_{\varkappa-1}^{0}} (7.38)

for a constant CC solely depending on D{{D}}.

Denote by CHC_{H} a constant (solely depending on D{{D}}) such that

rD1vL2CHvL2vH01(D).\|r_{{D}}^{-1}v\|_{L^{2}}\leq C_{H}\|\nabla v\|_{L^{2}}\qquad\forall v\in H_{0}^{1}({{D}}).

This constant exists as a consequence of Hardy’s inequality, see e.g. [64] and [82, 93] for the statement and proof of the inequality on bounded Lipschitz domains. Then due to

ρ(a)𝒰L22\displaystyle\rho(a)\|\nabla{\mathcal{U}}\|_{L^{2}}^{2} (Da𝒰𝒰¯d𝒙)=(Df𝒰¯d𝒙)\displaystyle\leq\Re\left(\int_{{D}}a\nabla{\mathcal{U}}\cdot\overline{\nabla{\mathcal{U}}}\,\mathrm{d}{\boldsymbol{x}}\right)=\Re\left(\int_{{D}}f\overline{{\mathcal{U}}}\,\mathrm{d}{\boldsymbol{x}}\right)
rD1ϰfL2rDϰ1𝒰L2f𝒦ϰ10rD1𝒰L2\displaystyle\leq\|r_{{D}}^{1-\varkappa}f\|_{L^{2}}\|r_{{D}}^{\varkappa-1}{\mathcal{U}}\|_{L^{2}}\leq\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}\|r_{{D}}^{-1}{\mathcal{U}}\|_{L^{2}}
CHf𝒦ϰ10𝒰L2\displaystyle\leq C_{H}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}\|\nabla{\mathcal{U}}\|_{L^{2}}

it holds

𝒰L2CHf𝒦ϰ10ρ(a).\|\nabla{\mathcal{U}}\|_{L^{2}}\leq\frac{C_{H}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}}{\rho(a)}.

Hence, using rD1ϰ1r_{{D}}^{1-\varkappa}\leq 1, we have that

f~𝒦ϰ10\displaystyle\|\tilde{f}\|_{{\mathcal{K}}_{\varkappa-1}^{0}} =rD1ϰa(f+a𝒰)L2\displaystyle=\left\|\frac{r_{{D}}^{1-\varkappa}}{a}\Big{(}f+\nabla a\cdot\nabla{\mathcal{U}}\Big{)}\right\|_{L^{2}}
1aL(rD1ϰfL2+aL𝒰L2)\displaystyle\leq\left\|\frac{1}{a}\right\|_{L^{\infty}}\left(\|r_{{D}}^{1-\varkappa}f\|_{L^{2}}+\|\nabla a\|_{L^{\infty}}\|\nabla{\mathcal{U}}\|_{L^{2}}\right)
1ρ(a)(f𝒦ϰ10+aW1CHf𝒦ϰ10ρ(a))\displaystyle\leq\frac{1}{\rho(a)}\left(\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}+\|a\|_{W^{1}_{\infty}}\frac{C_{H}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}}{\rho(a)}\right)
=f𝒦ϰ10ρ(a)(1+CHaW1ρ(a))\displaystyle=\frac{\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}}{\rho(a)}\left(1+\frac{C_{H}\|a\|_{W^{1}_{\infty}}}{\rho(a)}\right)
(1+CH)1ρ(a)aW1ρ(a)f𝒦ϰ10.\displaystyle\leq(1+C_{H})\frac{1}{\rho(a)}\frac{\|a\|_{W^{1}_{\infty}}}{\rho(a)}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{0}}.

The statement follows by (7.38).

Step 2. For general ss\in\mathbb{N}, s2s\geq 2, we proceed by induction. Assume the theorem holds for s12s-1\geq 2. Then for

f𝒦ϰ1s2(D)𝒦ϰ1s3(D)f\in{\mathcal{K}}_{\varkappa-1}^{s-2}({{D}})\hookrightarrow{\mathcal{K}}_{\varkappa-1}^{s-3}({{D}})

and

aW1(D)𝒲s1(D)W1(D)𝒲s2(D),a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}})\hookrightarrow W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-2}_{\infty}({{D}}),

we get

𝒰𝒦ϰ+1s1Cs1ρ(a)(aW1+a𝒲s2ρ(a))Ns1f𝒦ϰ1s3.\|{\mathcal{U}}\|_{{\mathcal{K}}_{\varkappa+1}^{s-1}}\leq\frac{C_{s-1}}{\rho(a)}\left(\frac{\|a\|_{W^{1}_{\infty}}+\|a\|_{{\mathcal{W}}^{s-2}_{\infty}}}{\rho(a)}\right)^{N_{s-1}}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-3}}. (7.39)

As in Step 1, it holds

Δ𝒰=1a(f+a𝒰)=:f~.-\Delta{\mathcal{U}}=\frac{1}{a}\Big{(}f+\nabla a\cdot\nabla{\mathcal{U}}\Big{)}=:\tilde{f}.

By Lemma 7.7 and (7.39), for some constant CC (which can change in each line, but solely depends on D{{D}} and ss) we have that

f~𝒦ϰ1s2\displaystyle\|\tilde{f}\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}} C1a𝒲s2,f+a𝒰𝒦ϰ1s2\displaystyle\leq C\left\|\frac{1}{a}\right\|_{{\mathcal{W}}^{s-2,\infty}}\|f+\nabla a\cdot\nabla{\mathcal{U}}\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}
Ca𝒲s2s2ρ(a)s1(f𝒦ϰ1s2+a𝒲s1𝒰𝒦ϰ+1s1)\displaystyle\leq C\frac{\|a\|_{{\mathcal{W}}^{s-2}}^{s-2}}{\rho(a)^{s-1}}\left(\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}+\|a\|_{{\mathcal{W}}^{s-1}_{\infty}}\|{\mathcal{U}}\|_{{\mathcal{K}}_{\varkappa+1}^{s-1}}\right)
Ca𝒲s2s2ρ(a)s1(f𝒦ϰ1s2+Cs1a𝒲s1ρ(a)(aW1+a𝒲s2ρ(a))Ns1f𝒦ϰ1s3)\displaystyle\leq C\frac{\|a\|_{{\mathcal{W}}^{s-2}}^{s-2}}{\rho(a)^{s-1}}\left(\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}+C_{s-1}\frac{\|a\|_{{\mathcal{W}}^{s-1}_{\infty}}}{\rho(a)}\left(\frac{\|a\|_{W^{1}_{\infty}}+\|a\|_{{\mathcal{W}}_{\infty}^{s-2}}}{\rho(a)}\right)^{N_{s-1}}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-3}}\right)
C1ρ(a)(aW1+a𝒲s1ρ(a))Ns1+1+(s2)f𝒦ϰ1s2.\displaystyle\leq C\frac{1}{\rho(a)}\left(\frac{\|a\|_{W^{1}_{\infty}}+\|a\|_{{\mathcal{W}}_{\infty}^{s-1}}}{\rho(a)}\right)^{N_{s-1}+1+(s-2)}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}.

Note that

Ns1+(s1)=(s1)(s2)2+(s1)=s(s1)2=Ns.N_{s-1}+(s-1)=\frac{(s-1)(s-2)}{2}+(s-1)=\frac{s(s-1)}{2}=N_{s}.

We now use (7.39) and the fact that (7.37) is a boundedly invertible isomorphism to conclude that there exist CsC_{s} such that (7.36) holds. ∎

Throughout the rest of this section D{{D}} is assumed a bounded polygonal domain and ϰ>0\varkappa>0 the constant from Theorem 7.8.

Assumption 7.9.

For some fixed ss\in\mathbb{N}, s2s\geq 2, there exist constants C>0C>0 and α>0\alpha>0, and a sequence (Xl)l(X_{l})_{l\in\mathbb{N}} of subspaces of X=H01(D;)=:H01X=H_{0}^{1}({{D}};\mathbb{C})=:H_{0}^{1}, such that

  1. (i)

    𝔴l:=dim(Xl)\mathfrak{w}_{l}:={\rm dim}(X_{l}), ll\in\mathbb{N}, satisfies Assumption 7.1 (for some K𝔚>0K_{\mathfrak{W}}>0),

  2. (ii)

    for all ll\in\mathbb{N}

    sup0u𝒦ϰ+1sinfvXluvH01u𝒦ϰ+1sC𝔴lα.\sup_{0\neq u\in{\mathcal{K}}_{\varkappa+1}^{s}}\frac{\inf_{v\in X_{l}}\|u-v\|_{H_{0}^{1}}}{\|u\|_{{\mathcal{K}}_{\varkappa+1}^{s}}}\leq C\mathfrak{w}_{l}^{-\alpha}. (7.40)

The constant α\alpha in Assumption 7.9 can be interpreted as the convergence rate of the finite element method. For the Kondrat’ev space 𝒦ϰ+1s(D){\mathcal{K}}_{\varkappa+1}^{s}({{D}}), finite element spaces XlX_{l} of piecewise polynomials of degree s1s-1 have been constructed in [25, Theorem 4.4], which achieve the optimal (in space dimension 22) convergence rate

α=s12\alpha=\frac{s-1}{2} (7.41)

in (7.40). For these spaces, Assumption 7.9 holds with this α\alpha, which consequently allows us to retain optimal convergence rates. Nonetheless we keep the discussion general in the following, and assume arbitrary positive α>0\alpha>0.

We next introduce the finite element solutions of (7.29) in the spaces XlX_{l}, and provide the basic error estimate.

Lemma 7.10.

Let Assumption 7.9 be satisfied for some s2s\geq 2. Let f𝒦ϰ1s2(D)f\in{\mathcal{K}}_{\varkappa-1}^{s-2}({{D}}) and

aW1(D)𝒲s1(D)L(D)a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}})\subseteq L^{\infty}({{D}})

with ρ(a)>0\rho(a)>0 and denote for ll\in\mathbb{N} by 𝒰l(a)Xl{\mathcal{U}}^{l}(a)\in X_{l} the unique solution of

Da(𝒰l)v¯d𝒙=f,vvXl,\int_{{D}}a(\nabla{\mathcal{U}}^{l})^{\top}\overline{\nabla v}\,\mathrm{d}{\boldsymbol{x}}=\left\langle f,v\right\rangle\qquad\forall v\in X_{l},

where the right hand side denotes the (sesquilinear) dual pairing between H1(D)H^{-1}({{D}}) and H01(D)H_{0}^{1}({{D}}). Then for the solution 𝒰(a)H01(D){\mathcal{U}}(a)\in H_{0}^{1}({{D}}) it holds with the constants NsN_{s}, CsC_{s} from Theorem. 7.8,

𝒰(a)𝒰l(a)H01𝔴lαCaLρ(a)𝒰(a)𝒦ϰ+1s𝔴lαCCs(aW1+a𝒲s1)Ns+1ρ(a)Ns+2f𝒦ϰ1s2.\|{\mathcal{U}}(a)-{\mathcal{U}}^{l}(a)\|_{H_{0}^{1}}\leq\mathfrak{w}_{l}^{-\alpha}C\frac{\|a\|_{L^{\infty}}}{\rho(a)}\|{\mathcal{U}}(a)\|_{{\mathcal{K}}_{\varkappa+1}^{s}}\leq\mathfrak{w}_{l}^{-\alpha}CC_{s}\frac{(\|a\|_{W^{1}_{\infty}}+\|a\|_{{\mathcal{W}}^{s-1}_{\infty}})^{N_{s}+1}}{\rho(a)^{N_{s}+2}}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}\;.

Here C>0C>0 is the constant from Assumption 7.9.

Proof.

By Céa’s lemma in complex form we derive that

𝒰(a)𝒰l(a)H01aLρ(a)infvXl𝒰(a)vH01.\displaystyle\|{\mathcal{U}}(a)-{\mathcal{U}}^{l}(a)\|_{H_{0}^{1}}\leq\frac{\|a\|_{L^{\infty}}}{\rho(a)}\inf_{v\in X_{l}}\|{\mathcal{U}}(a)-v\|_{H_{0}^{1}}.

Hence the assertion follows by Assumption 7.9 and (7.36). ∎

Throughout the rest of this section, as earlier we expand the logarithm of the diffusion coefficient

a(𝒚)=exp(jyjψj)a({\boldsymbol{y}})=\exp\Bigg{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\Bigg{)}

in terms of a sequence ψjW1(D)𝒲s1(D)\psi_{j}\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}_{\infty}^{s-1}({{D}}), jj\in\mathbb{N}. Denote

b1,j:=ψjL,b2,j:=max{ψjW1,ψj𝒲s1}b_{1,j}:=\|\psi_{j}\|_{L^{\infty}},\quad b_{2,j}:=\max\big{\{}\|\psi_{j}\|_{W^{1}_{\infty}},\|\psi_{j}\|_{{\mathcal{W}}_{\infty}^{s-1}}\big{\}} (7.42)

and 𝒃1:=(b1,j)j{\boldsymbol{b}}_{1}:=(b_{1,j})_{j\in\mathbb{N}}, 𝒃2:=(b2,j)j{\boldsymbol{b}}_{2}:=(b_{2,j})_{j\in\mathbb{N}}.

Example 7.11.

Let D=[0,1]{{D}}=[0,1] and ψj(x)=sin(jx)jr\psi_{j}(x)=\sin(jx)j^{-r} for some r>2r>2. Then 𝐛1p1(){\boldsymbol{b}}_{1}\in\ell^{p_{1}}({\mathbb{N}}) for every p1>1rp_{1}>\frac{1}{r} and 𝐛2p2(){\boldsymbol{b}}_{2}\in\ell^{p_{2}}({\mathbb{N}}) for every p2>1r(s1)p_{2}>\frac{1}{r-(s-1)}.

In the next proposition we verify Assumption 7.2. This will yield validity of the multilevel convergence rates proved in Theorems 7.5 and 7.6 in the present setting as we discuss subsequently.

Proposition 7.12.

Let Assumption 7.9 be satisfied for some s2s\geq 2 and α>0\alpha>0. Let 𝐛1p1(){\boldsymbol{b}}_{1}\in\ell^{p_{1}}({\mathbb{N}}), 𝐛2p2(){\boldsymbol{b}}_{2}\in\ell^{p_{2}}({\mathbb{N}}) with p1p_{1}, p2(0,1)p_{2}\in(0,1).

Then there exist ξ>0\xi>0 and δ>0\delta>0 such that

u(𝒚):=𝒰(exp(jyjψj))u({\boldsymbol{y}}):={\mathcal{U}}\Bigg{(}\exp\Bigg{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\Bigg{)}\Bigg{)} (7.43)

is (𝐛1,ξ,δ,H01)({\boldsymbol{b}}_{1},\xi,\delta,H_{0}^{1})-holomorphic, and for every ll\in\mathbb{N}

  1. (i)

    ul(𝒚):=𝒰l(exp(jyjψj))u^{l}({\boldsymbol{y}}):={\mathcal{U}}^{l}(\exp(\sum_{j\in\mathbb{N}}y_{j}\psi_{j})) is (𝒃1,ξ,δ,H01)({\boldsymbol{b}}_{1},\xi,\delta,H_{0}^{1})-holomorphic,

  2. (ii)

    uulu-u^{l} is (𝒃1,ξ,δ,H01)({\boldsymbol{b}}_{1},\xi,\delta,H_{0}^{1})-holomorphic,

  3. (iii)

    uulu-u^{l} is (𝒃2,ξ,δ𝔴lα,H01)({\boldsymbol{b}}_{2},\xi,\delta\mathfrak{w}_{l}^{-\alpha},H_{0}^{1})-holomorphic.

Proof.

Step 1. We show (i) and (ii). The argument to show that ulu^{l} is (𝒃1,ξ,δ,H01)({\boldsymbol{b}}_{1},\xi,\delta,H_{0}^{1})-holomorphic (for some constants ξ>0\xi>0, δ>0\delta>0 independent of ll) is essentially the same as in Section 4.3.1.

We wish to apply Theorem 4.11 with E=L(D)E=L^{\infty}({{D}}) and X=H10X=H^{1}_{0}. To this end let

O1={aL(D;):ρ(a)>0}L(D;).O_{1}=\{a\in L^{\infty}({{D}};\mathbb{C})\,:\,\rho(a)>0\}\subset L^{\infty}({{D}};\mathbb{C}).

By assumption, b1,j=ψjLb_{1,j}=\|\psi_{j}\|_{L^{\infty}} satisfies 𝒃1=(b1,j)jp1()1(){\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}}\in\ell^{p_{1}}({\mathbb{N}})\subseteq\ell^{1}({\mathbb{N}}), which corresponds to assumption (iv) of Theorem 4.11. It remains to verify assumptions (i), (ii) and (iii) of Theorem 4.11:

  1. (i)

    𝒰l:O1H01{\mathcal{U}}^{l}:O_{1}\to H_{0}^{1} is holomorphic: This is satisfied because the operation of inversion of linear operators is holomorphic on the set of boundedly invertible linear operators. Denote by Al:XlXlA_{l}:X_{l}\to X_{l}^{\prime} the differential operator

    Alu=div(au)XlA_{l}u=-\operatorname{div}(a\nabla u)\in X_{l}^{\prime}

    via

    Alu,v=Dauv¯d𝒙vXl.\left\langle A_{l}u,v\right\rangle=\int_{{{D}}}a\nabla u^{\top}\overline{\nabla v}\,\mathrm{d}{\boldsymbol{x}}\quad\forall v\in X_{l}.

    Observe that AlA_{l} depends boundedly and linearly (thus holomorphically) on aa, and therefore, the map aAl(a)1f=𝒰l(a)a\mapsto A_{l}(a)^{-1}f={\mathcal{U}}^{l}(a) is a composition of holomorphic functions. We refer once more to [111, Example 1.2.38] for more details.

  2. (ii)

    It holds for all aOa\in O

    𝒰l(a)H01fXlρ(a)fH1ρ(a).\|{\mathcal{U}}^{l}(a)\|_{H_{0}^{1}}\leq\frac{\|f\|_{X_{l}^{\prime}}}{\rho(a)}\leq\frac{\|f\|_{H^{-1}}}{\rho(a)}.

    The first inequality follows by the same calculation as (4.20) (but with XX replaced by XlX_{l}), and the second inequality follows by the definition of the dual norm, viz

    fXl=sup0vXl|f,v|vH01sup0vH01|f,v|vH01=fH1.\|f\|_{X_{l}^{\prime}}=\sup_{0\neq v\in X_{l}}\frac{|\left\langle f,v\right\rangle|}{\|v\|_{H_{0}^{1}}}\leq\sup_{0\neq v\in H_{0}^{1}}\frac{|\left\langle f,v\right\rangle|}{\|v\|_{H_{0}^{1}}}=\|f\|_{H^{-1}}.
  3. (iii)

    For all aa, bOb\in O we have

    𝒰l(a)𝒰l(b)H01fH11min{ρ(a),ρ(b)}2abL,\|{\mathcal{U}}^{l}(a)-{\mathcal{U}}^{l}(b)\|_{H_{0}^{1}}\leq\|f\|_{H^{-1}}\frac{1}{\min\{\rho(a),\rho(b)\}^{2}}\|a-b\|_{L^{\infty}},

    which follows again by the same calculation as in in the proof of (4.21).

According to Theorem 4.11 the map

𝒰lL2(U,Xl;γ)L2(U,H01;γ){\mathcal{U}}^{l}\in L^{2}(U,X_{l};\gamma)\subseteq L^{2}(U,H_{0}^{1};\gamma)

is (𝒃1,ξ1,C~1,H01)({\boldsymbol{b}}_{1},\xi_{1},\tilde{C}_{1},H_{0}^{1})-holomorphic, for some fixed constants ξ1>0\xi_{1}>0 and C~1>0\tilde{C}_{1}>0 depending on O1O_{1} but independent of ll. In fact the argument also works with H01H_{0}^{1} instead of XlX_{l}, i.e. also uu is (𝒃1,ξ1,C~1,H01)({\boldsymbol{b}}_{1},\xi_{1},\tilde{C}_{1},H_{0}^{1})-holomorphic (with the same constants ξ1\xi_{1} and C~1\tilde{C}_{1}).

Finally, it follows directly from the definition that the difference uulu-u^{l} is (𝒃1,ξ,2δ,H01)({\boldsymbol{b}}_{1},\xi,2\delta,H_{0}^{1})-holomorphic.

Step 2. To show (iii), we set

O2={aW1(D)𝒲s1(D):ρ(a)>0},O_{2}=\{a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}})\,:\,\rho(a)>0\},

and verify again assumptions (i), (ii) and (iii) of Theorem 4.11, but now with “EE” in this lemma being W1(D)𝒲s1(D)W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}}). First, observe that with

b2,j:=max{ψj𝒲s1,ψjW1},b_{2,j}:=\max\big{\{}\|\psi_{j}\|_{{\mathcal{W}}^{s-1}_{\infty}},\|\psi_{j}\|_{W^{1}_{\infty}}\big{\}},

by assumption

𝒃2=(b2,j)jp2()1(){\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}}\in\ell^{p_{2}}({\mathbb{N}})\hookrightarrow\ell^{1}({\mathbb{N}})

which corresponds to the assumption (iv) of Theorem 4.11.

For every ll\in\mathbb{N}:

  1. (i)

    𝒰𝒰l:O2H01(D){\mathcal{U}}-{\mathcal{U}}^{l}:O_{2}\to H_{0}^{1}({{D}}) is holomorphic: Since O2O_{2} can be considered a subset of O1O_{1} (and O2O_{2} is equipped with a stronger topology than O1O_{1}), Fréchet differentiability follows by Fréchet differentiability of

    𝒰𝒰l:O1H01(D),{\mathcal{U}}-{\mathcal{U}}^{l}:O_{1}\to H_{0}^{1}({{D}}),

    which holds by Step 1.

  2. (ii)

    For every aO2a\in O_{2}

    (𝒰𝒰l)(a)H01𝔴lαCCsf𝒦ϰ1s2=:δl(aW1+a𝒲s1)Ns+1ρ(a)Ns+2\|({\mathcal{U}}-{\mathcal{U}}^{l})(a)\|_{H_{0}^{1}}\leq\underbrace{\mathfrak{w}_{l}^{-\alpha}CC_{s}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}}_{=:\delta_{l}}\frac{(\|a\|_{W^{1}_{\infty}}+\|a\|_{{\mathcal{W}}^{s-1}_{\infty}})^{N_{s}+1}}{\rho(a)^{N_{s}+2}}

    by Lemma 7.10.

  3. (iii)

    For every aa, bO2O1b\in O_{2}\subseteq O_{1}, by Step 1 and (4.21),

    (𝒰𝒰l)(a)(𝒰𝒰l)(b)H01\displaystyle\|({\mathcal{U}}-{\mathcal{U}}^{l})(a)-({\mathcal{U}}-{\mathcal{U}}^{l})(b)\|_{H_{0}^{1}} 𝒰(a)𝒰(b)H01+𝒰l(a)𝒰l(b)H01\displaystyle\leq\|{\mathcal{U}}(a)-{\mathcal{U}}(b)\|_{H_{0}^{1}}+\|{\mathcal{U}}^{l}(a)-{\mathcal{U}}^{l}(b)\|_{H_{0}^{1}}
    fH12min{ρ(a),ρ(b)}2abL.\displaystyle\leq\|f\|_{H^{-1}}\frac{2}{\min\{\rho(a),\rho(b)\}^{2}}\|a-b\|_{L^{\infty}}.

    We conclude with Theorem 4.11 that there exist ξ2\xi_{2} and C~2\tilde{C}_{2} depending on O2O_{2}, D{{D}} but independent of ll such that uulu-u^{l} is (𝒃2,ξ2,C~2δl,H01)({\boldsymbol{b}}_{2},\xi_{2},\tilde{C}_{2}\delta_{l},H_{0}^{1})-holomorphic.

In all, the proposition holds with

ξ:=min{ξ1,ξ2}andδ:=max{C~1,C~2CCsf𝒦ϰ1s2}.\xi:=\min\{\xi_{1},\xi_{2}\}\qquad\text{and}\qquad\delta:=\max\{\tilde{C}_{1},\tilde{C}_{2}CC_{s}\|f\|_{{\mathcal{K}}_{\varkappa-1}^{s-2}}\}.\qed

Items (ii) and (iii) of Proposition 7.12 show that Assumption 7.9 implies validity of Assumption 7.2. This in turn allows us to apply Theorems 7.5 and 7.6. Specifically, assuming the optimal convergence rate α=s12\alpha=\frac{s-1}{2} in (7.41), we obtain that for uu in (7.43) and every nn\in\mathbb{N} there is ε:=εn>0\varepsilon:=\varepsilon_{n}>0 such that work(𝐥ε)n\mathrm{work}(\mathbf{l}_{\varepsilon})\leq n and the multilevel interpolant 𝐈ML𝐥\mathbf{I}^{\rm ML}_{\mathbf{l}} defined in (7.3) satisfies

u𝐈𝐥εMLuL2(U,H10;γ)C(1+logn)nRI,RI=min{s12,s12(1p132)s12+1p11p2},\|u-\mathbf{I}_{\mathbf{l}_{\varepsilon}}^{\rm ML}u\|_{L^{2}(U,{H^{1}_{0}};\gamma)}\leq C(1+\log n)n^{-R_{I}},\quad R_{I}=\min\left\{\frac{s-1}{2},\frac{\frac{s-1}{2}(\frac{1}{p_{1}}-\frac{3}{2})}{\frac{s-1}{2}+\frac{1}{p_{1}}-\frac{1}{p_{2}}}\right\},

and the multilevel quadrature operator 𝐐ML𝐥\mathbf{Q}^{\rm ML}_{\mathbf{l}} defined in (7.4) satisfies

Uu(𝒚)dγ(𝒚)𝐐𝐥εMLuH10C(1+logn)nRQ,RQ=min{s12,s12(2p152)s12+2p12p2}.\left\|\int_{U}u({\boldsymbol{y}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\mathbf{l}_{\varepsilon}}^{\rm ML}u\right\|_{H^{1}_{0}}\leq C(1+\log n)n^{-R_{Q}},\quad R_{Q}=\min\left\{\frac{s-1}{2},\frac{\frac{s-1}{2}(\frac{2}{p_{1}}-\frac{5}{2})}{\frac{s-1}{2}+\frac{2}{p_{1}}-\frac{2}{p_{2}}}\right\}.

Let us consider these convergence rates in the case where the ψj\psi_{j} are algebraically decreasing, with this decrease encoded by some r>1r>1: if for fixed but arbitrarily small ε>0\varepsilon>0 holds ψjLjrε\|\psi_{j}\|_{L^{\infty}}\sim j^{-r-\varepsilon}, and we assume (cp. Ex. 7.11)

max{ψjW1,ψj𝒲s1}jr+(s1)ε,\max\big{\{}\|\psi_{j}\|_{W^{1}_{\infty}},\|\psi_{j}\|_{{\mathcal{W}}^{s-1}_{\infty}}\big{\}}\sim j^{-r+(s-1)-\varepsilon},

then setting s:=rs:=r we can choose p1=1rp_{1}=\frac{1}{r} and p2=1p_{2}=1. Inserting those numbers, the convergence rates become

RI=min{r12,r12(r32)r12+r1}=r312andRQ=min{r12,r12(2r52)r12+2r2}=2r512.R_{I}=\min\left\{\frac{r-1}{2},\frac{\frac{r-1}{2}(r-\frac{3}{2})}{\frac{r-1}{2}+r-1}\right\}=\frac{r}{3}-\frac{1}{2}\quad\text{and}\quad R_{Q}=\min\left\{\frac{r-1}{2},\frac{\frac{r-1}{2}(2r-\frac{5}{2})}{\frac{r-1}{2}+2r-2}\right\}=\frac{2r}{5}-\frac{1}{2}.

7.6.2 Parametric holomorphy of the posterior density in Bayesian PDE inversion

Throughout this section we assume that D2{{D}}\subseteq\mathbb{R}^{2} is a polygonal Lipschitz domain and that f𝒦ϰ1s2(D)f\in{\mathcal{K}}_{\varkappa-1}^{s-2}({{D}}) with ϰ\varkappa as in Theorem 7.8.

As in Section 5, to treat the approximation of the (unnormalized) posterior density or its integral, we need an upper bound on u(𝒚)H01\|u({\boldsymbol{y}})\|_{H_{0}^{1}} for all 𝒚{\boldsymbol{y}}. This is achieved by considering (7.29) with diffusion coefficient a0+aa_{0}+a where

ρ(a0):=essinf𝒙D(a0)>0.\rho(a_{0}):=\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\Re(a_{0})>0.

The shift of the diffusion coefficient by a0a_{0} ensures uniform ellipticity for all

a{aL(D,):ρ(a)0}.a\in\{a\in L^{\infty}({{D}},{\mathbb{C}})\,:\,\rho(a)\geq 0\}.

As a consequence, solutions 𝒰(a0+a)X=H01(D;)=:H10{\mathcal{U}}(a_{0}+a)\in X=H_{0}^{1}({{D}};\mathbb{C})=:H^{1}_{0} of (7.29) satisfy the apriori bound (cp. (4.20))

𝒰(a0+a)H10fH1ρ(a0).\|{\mathcal{U}}(a_{0}+a)\|_{H^{1}_{0}}\leq\frac{\|f\|_{H^{-1}}}{\rho(a_{0})}. (7.44)
As before, for a sequence of subspaces (Xl)l(X_{l})_{l\in\mathbb{N}} of H01(D,)H_{0}^{1}({{D}},{\mathbb{C}}), for aOa\in O we denote by 𝒰l(a)Xl{\mathcal{U}}^{l}(a)\in X_{l} the finite element approximation to 𝒰(a){\mathcal{U}}(a). By the same calculation as for 𝒰{\mathcal{U}} it also holds
𝒰l(a0+a)H10fH1ρ(a0)\|{\mathcal{U}}^{l}(a_{0}+a)\|_{H^{1}_{0}}\leq\frac{\|f\|_{H^{-1}}}{\rho(a_{0})}
independent of ll.

Assuming that bj=ψjLb_{j}=\|\psi_{j}\|_{L^{\infty}} satisfies (bj)j1()(b_{j})_{j\in\mathbb{N}}\in\ell^{1}({\mathbb{N}}), the function u(𝒚)=𝒰(a0+a(𝒚))u({\boldsymbol{y}})={\mathcal{U}}(a_{0}+a({\boldsymbol{y}})) with

a(𝒚)=exp(jyjψj),a({\boldsymbol{y}})=\exp\bigg{(}\sum_{j\in\mathbb{N}}y_{j}\psi_{j}\bigg{)},

is well-defined. For a fixed observation 𝖉m{\boldsymbol{\mathfrak{d}}}\in\mathbb{R}^{m} consider again the (unnormalized) posterior density given in (5.4),

π~(𝒚|𝖉):=exp((𝖉𝓞(u(𝒚)))𝚪1(𝖉𝓞(u(𝒚)))).\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}):=\exp\left(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u({\boldsymbol{y}})))\right).

Recall that 𝓞:Xm{\boldsymbol{{\mathcal{O}}}}:X\to\mathbb{C}^{m} (the observation operator) is assumed to be a bounded linear map, and 𝚪m×m{\boldsymbol{\Gamma}}\in\mathbb{R}^{m\times m} (the noise covariance matrix) is symmetric positive definite. For ll\in\mathbb{N} (tagging discretization level of the PDE), and with ul(𝒚)=𝒰l(a0+a(𝒚))u^{l}({\boldsymbol{y}})={\mathcal{U}}^{l}(a_{0}+a({\boldsymbol{y}})), we introduce approximations

π~l(𝒚|𝖉):=exp((𝖉𝓞(ul(𝒚)))𝚪1(𝖉𝓞(ul(𝒚))))\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}):=\exp\left(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u^{l}({\boldsymbol{y}})))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u^{l}({\boldsymbol{y}})))\right)

to π~(𝒚|𝖉)\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}). In the following we show the analog of Proposition 7.12, that is we show validity of the assumptions required for the multilevel convergence results.

Lemma 7.13.

Let 𝓞:H01(D;)m{\boldsymbol{{\mathcal{O}}}}:H_{0}^{1}({{D}};\mathbb{C})\to\mathbb{C}^{m} be a bounded linear operator, 𝖉m{\boldsymbol{\mathfrak{d}}}\in\mathbb{C}^{m} and 𝚪m×m{\boldsymbol{\Gamma}}\in\mathbb{R}^{m\times m} symmetric positive definite. Set

Φ:={H01(D;)uexp((𝖉𝓞(u))𝚪1(𝖉𝓞(u))).\Phi:=\begin{cases}H_{0}^{1}({{D}};\mathbb{C})\to\mathbb{C}\\ u\mapsto\exp(-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u))).\end{cases}

Then the function Φ\Phi is continuously differentiable and for every r>0r>0 has a Lipschitz constant KK solely depending on 𝚪1\|{\boldsymbol{\Gamma}}^{-1}\|, 𝓞L(H01;m)\|{\boldsymbol{{\mathcal{O}}}}\|_{L(H_{0}^{1};\mathbb{C}^{m})}, 𝖉\|{\boldsymbol{\mathfrak{d}}}\| and rr, on the set

{uH01(D;):uH01<r}.\{u\in H_{0}^{1}({{D}};\mathbb{C})\,:\,\|u\|_{H_{0}^{1}}<r\}.
Proof.

The function Φ\Phi is continuously differentiable as a composition of continuously differentiable functions. Hence for uu, vv with w:=uvw:=u-v and with the derivative DΦ:H01L(H01;)D\Phi:H_{0}^{1}\to L(H_{0}^{1};\mathbb{C}) of Φ\Phi,

Φ(u)Φ(v)=01DΦ(v+tw)wdt.\Phi(u)-\Phi(v)=\int_{0}^{1}D\Phi(v+tw)w\,\mathrm{d}t. (7.45)

Due to the symmetry of 𝚪{\boldsymbol{\Gamma}} it holds

DΦ(u+tw)w=2𝓞(w)𝚪1(𝖉𝓞(u+tw))exp((𝖉𝓞(u+tw))𝚪1(𝖉𝓞(u+tw))).D\Phi(u+tw)w=2{\boldsymbol{{\mathcal{O}}}}(w)^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u+tw))\exp\Big{(}-({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u+tw))^{\top}{\boldsymbol{\Gamma}}^{-1}({\boldsymbol{\mathfrak{d}}}-{\boldsymbol{{\mathcal{O}}}}(u+tw))\Big{)}.

If uH01\|u\|_{H_{0}^{1}}, vH01<r\|v\|_{H_{0}^{1}}<r then also u+twH01<r\|u+tw\|_{H_{0}^{1}}<r for all t[0,1]t\in[0,1] and we can bound

|DΦ(u+tw)w|KwH01,|D\Phi(u+tw)w|\leq K\|w\|_{H_{0}^{1}},

where

K:=2𝓞L(H01;m)𝚪1(𝖉+r𝓞L(H01;m))exp(𝚪1(𝖉+𝓞L(H01;m)r)2).K:=2\|{\boldsymbol{{\mathcal{O}}}}\|_{L(H_{0}^{1};\mathbb{C}^{m})}\|{\boldsymbol{\Gamma}}^{-1}\|(\|{\boldsymbol{\mathfrak{d}}}\|+r\|{\boldsymbol{{\mathcal{O}}}}\|_{L(H_{0}^{1};\mathbb{C}^{m})})\exp(\|{\boldsymbol{\Gamma}}^{-1}\|(\|{\boldsymbol{\mathfrak{d}}}\|+\|{\boldsymbol{{\mathcal{O}}}}\|_{L(H_{0}^{1};\mathbb{C}^{m})}r)^{2}). (7.46)

The statement follows by (7.45). ∎

Remark 7.14.

The reason why we require the additional positive a0a_{0} term in (7.44), is to guarantee boundedness of the solution 𝒰(a){\mathcal{U}}(a) and Lipschitz continuity of Φ\Phi.

Proposition 7.15.

Let Assumption 7.9 be satisfied for some s2s\geq 2 and α>0\alpha>0. Let a0a_{0}, (ψj)jW1(D)𝒲s1(D)(\psi_{j})_{j\in\mathbb{N}}\subseteq W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}}) and 𝐛1p1(){\boldsymbol{b}}_{1}\in\ell^{p_{1}}({\mathbb{N}}), 𝐛2p2(){\boldsymbol{b}}_{2}\in\ell^{p_{2}}({\mathbb{N}}) with p1p_{1}, p2(0,1)p_{2}\in(0,1) (see (7.42) for the definition of 𝐛1{\boldsymbol{b}}_{1}, 𝐛2{\boldsymbol{b}}_{2}). Fix 𝖉m{\boldsymbol{\mathfrak{d}}}\in\mathbb{C}^{m}.

Then there exist ξ>0\xi>0 and δ>0\delta>0 such that π~(𝐲|𝖉)\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) is (𝐛1,ξ,δ,)({\boldsymbol{b}}_{1},\xi,\delta,\mathbb{C})-holomorphic, and for every ll\in\mathbb{N}

  1. (i)

    π~l(𝒚|𝖉)\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) is (𝒃1,ξ,δ,)({\boldsymbol{b}}_{1},\xi,\delta,\mathbb{C})-holomorphic,

  2. (ii)

    π~(𝒚|𝖉)π~l(𝒚|𝖉)\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})-\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) is (𝒃1,ξ,δ,H01)({\boldsymbol{b}}_{1},\xi,\delta,H_{0}^{1})-holomorphic,

  3. (iii)

    π~(𝒚|𝖉)π~l(𝒚|𝖉)\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})-\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) is (𝒃2,ξ,δ𝔴lα,H01)({\boldsymbol{b}}_{2},\xi,\delta\mathfrak{w}_{l}^{-\alpha},H_{0}^{1})-holomorphic.

Proof.

Step 1. We show (i) and (ii). Set

O1:={aL(D;):ρ(a)>0}.O_{1}:=\{a\in L^{\infty}({{D}};\mathbb{C})\,:\,\rho(a)>0\}.

By (7.44) for all aO1a\in O_{1} and all ll\in\mathbb{N} with r:=fH1ρ(a0)r:=\frac{\|f\|_{H^{-1}}}{\rho(a_{0})}

𝒰l(a0+a)H01rand𝒰(a0+a)H01r.\|{\mathcal{U}}^{l}(a_{0}+a)\|_{H_{0}^{1}}\leq r\qquad\text{and}\qquad\|{\mathcal{U}}(a_{0}+a)\|_{H_{0}^{1}}\leq r. (7.47)

As in Step 1 of the proof of Proposition 7.12, one can show that u(𝒚)=𝒰(a0+a(𝒚))u({\boldsymbol{y}})={\mathcal{U}}(a_{0}+a({\boldsymbol{y}})) and ul(𝒚)=𝒰(a0+a(𝒚))u^{l}({\boldsymbol{y}})={\mathcal{U}}(a_{0}+a({\boldsymbol{y}})) where a(𝒚)=exp(jyjψj)a({\boldsymbol{y}})=\exp(\sum_{j\in\mathbb{N}}y_{j}\psi_{j}) are (𝒃1,ξ1,C~1,H01)({\boldsymbol{b}}_{1},\xi_{1},\tilde{C}_{1},H_{0}^{1})-holomorphic for certain ξ1>0\xi_{1}>0 and C~1>0\tilde{C}_{1}>0 (the only difference to Proposition 7.12 is the affine offset a0a_{0} in (7.29), which ensures a positive lower bound for a+a0a+a_{0}). In the following Φ\Phi is as in Lemma 7.13 and Ta0(a):=a0+aT_{a_{0}}(a):=a_{0}+a so that

π~(𝒚|𝖉)=Φ(𝒰l(Ta0(a(𝒚)))).\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})=\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a({\boldsymbol{y}})))). (7.48)

With b1,j=ψjLb_{1,j}=\|\psi_{j}\|_{L^{\infty}}, by the assumption

𝒃1=(b1,j)jp1()1(){\boldsymbol{b}}_{1}=(b_{1,j})_{j\in\mathbb{N}}\in\ell^{p_{1}}({\mathbb{N}})\hookrightarrow\ell^{1}({\mathbb{N}})

which corresponds to assumption (iv) of Theorem 4.11. We now verify assumptions (i), (ii) and (iii) of Theorem 4.11 for (7.48).

For every ll\in\mathbb{N}:

  1. (i)

    The map

    Φ𝒰lTa0:{O1aΦ(𝒰(Ta0(a)))\Phi\circ{\mathcal{U}}^{l}\circ T_{a_{0}}:\begin{cases}O_{1}\to\mathbb{C}\\ a\mapsto\Phi({\mathcal{U}}(T_{a_{0}}(a)))\end{cases}

    is holomorphic as a composition of holomorphic functions.

  2. (ii)

    for all aO1a\in O_{1}, since 𝒰l(Ta0(a))H01r\|{\mathcal{U}}^{l}(T_{a_{0}}(a))\|_{H_{0}^{1}}\leq r

    |Φ(𝒰l(Ta0(a)))|exp((𝖉+𝓞L(H01(D;);m)r)2𝚪1)|\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a)))|\leq\exp((\|{\boldsymbol{\mathfrak{d}}}\|+\|{\boldsymbol{{\mathcal{O}}}}\|_{L(H_{0}^{1}({{D}};\mathbb{C});\mathbb{C}^{m})}r)^{2}\|{\boldsymbol{\Gamma}}^{-1}\|)

    and thus assumption (ii) of Theorem 4.11 is trivially satisfied for some δ>0\delta>0 independent of ll,

  3. (iii)

    for all aa, bO1b\in O_{1} by Lemma 7.13 and the same calculation as in (4.21)

    |Φ(𝒰l(Ta0(a)))Φ(𝒰l(Ta0(b)))|\displaystyle|\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a)))-\Phi({\mathcal{U}}^{l}(T_{a_{0}}(b)))| K𝒰l(Ta0(a))𝒰l(Ta0(b))H01\displaystyle\leq K\|{\mathcal{U}}^{l}(T_{a_{0}}(a))-{\mathcal{U}}^{l}(T_{a_{0}}(b))\|_{H_{0}^{1}} (7.49)
    KfH1ρ(a0)abL,\displaystyle\leq K\frac{\|f\|_{H^{-1}}}{\rho(a_{0})}\|a-b\|_{L^{\infty}},

    where KK is the constant given as in (7.46).

Now we can apply Theorem 4.11 to conclude that there exist ξ1\xi_{1}, δ1\delta_{1} (independent of ll) such that π~l(|𝖉)\tilde{\pi}^{l}(\cdot|{\boldsymbol{\mathfrak{d}}}) is (𝒃1,ξ1,δ1,H01)({\boldsymbol{b}}_{1},\xi_{1},\delta_{1},H_{0}^{1})-holomorphic for every ll\in\mathbb{N}. Similarly one shows that π~(|𝖉)\tilde{\pi}(\cdot|{\boldsymbol{\mathfrak{d}}}) is (𝒃1,ξ1,δ1,H01)({\boldsymbol{b}}_{1},\xi_{1},\delta_{1},H_{0}^{1})-holomorphic, and in particular π~(|𝖉)π~l(|𝖉)\tilde{\pi}(\cdot|{\boldsymbol{\mathfrak{d}}})-\tilde{\pi}^{l}(\cdot|{\boldsymbol{\mathfrak{d}}}) is (𝒃1,ξ1,2δ1,H01)({\boldsymbol{b}}_{1},\xi_{1},2\delta_{1},H_{0}^{1})-holomorphic.

Step 2. Set

O2={aW1(D)𝒲s1(D):ρ(a)>0}.O_{2}=\{a\in W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}})\,:\,\rho(a)>0\}.

We verify once more assumptions (i), (ii) and (iii) of Theorem 4.11 with “EE” in this lemma being W1(D)𝒲s1(D)W^{1}_{\infty}({{D}})\cap{\mathcal{W}}^{s-1}_{\infty}({{D}}). With b2,j=max{ψjW1,ψj𝒲s1}b_{2,j}=\max\{\|\psi_{j}\|_{W^{1}_{\infty}},\|\psi_{j}\|_{{\mathcal{W}}^{s-1}_{\infty}}\}, by the assumption

4𝒃2=(b2,j)jp2()1(),4{\boldsymbol{b}}_{2}=(b_{2,j})_{j\in\mathbb{N}}\in\ell^{p_{2}}({\mathbb{N}})\hookrightarrow\ell^{1}({\mathbb{N}}),

which corresponds to assumption (iv) of Theorem 4.11.

We will apply Theorem 4.11 with the function

π~(𝒚|𝖉)π~l(𝒚|𝖉)=Φ(𝒰(Ta0(a(𝒚))))Φ(𝒰l(Ta0(a(𝒚)))).\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})-\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})=\Phi({\mathcal{U}}(T_{a_{0}}(a({\boldsymbol{y}}))))-\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a({\boldsymbol{y}})))). (7.50)

For every ll\in\mathbb{N}:

  1. (i)

    By item (i) in Step 1 (and because O2O1O_{2}\subseteq O_{1})

    Φ𝒰Ta0Φ𝒰lTa0:{O2aΦ(𝒰(Ta0(a)))Φ(𝒰l(Ta0(a)))\Phi\circ{\mathcal{U}}\circ T_{a_{0}}-\Phi\circ{\mathcal{U}}^{l}\circ T_{a_{0}}:\begin{cases}O_{2}\to\mathbb{C}\\ a\mapsto\Phi({\mathcal{U}}(T_{a_{0}}(a)))-\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a)))\end{cases}

    is holomorphic,

  2. (ii)

    for every aO2a\in O_{2}, by Lemma 7.10

    𝒰(Ta0(a))𝒰l(Ta0(a))H01𝔴lαCCs(a0+aW1+a0+a𝒲s1)Ns+1ρ(a0+a)Ns+2f𝒦s2ϰ1.\|{\mathcal{U}}(T_{a_{0}}(a))-{\mathcal{U}}^{l}(T_{a_{0}}(a))\|_{H_{0}^{1}}\leq\mathfrak{w}_{l}^{-\alpha}CC_{s}\frac{(\|a_{0}+a\|_{W^{1}_{\infty}}+\|a_{0}+a\|_{{\mathcal{W}}^{s-1}_{\infty}})^{N_{s}+1}}{\rho(a_{0}+a)^{N_{s}+2}}\|f\|_{{\mathcal{K}}^{s-2}_{\varkappa-1}}.

    Thus by (7.47) and Lemma 7.13

    |Φ(𝒰(Ta0(a)))Φ(𝒰l(Ta0(a)))|𝔴lαKCCs(a0+aW1+a0+a𝒲s1)Ns+1ρ(a0+a)Ns+2f𝒦s2ϰ1,|\Phi({\mathcal{U}}(T_{a_{0}}(a)))-\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a)))|\leq\mathfrak{w}_{l}^{-\alpha}KCC_{s}\frac{(\|a_{0}+a\|_{W^{1}_{\infty}}+\|a_{0}+a\|_{{\mathcal{W}}^{s-1}_{\infty}})^{N_{s}+1}}{\rho(a_{0}+a)^{N_{s}+2}}\|f\|_{{\mathcal{K}}^{s-2}_{\varkappa-1}},
  3. (iii)

    for all aa, bO2O1b\in O_{2}\subseteq O_{1} by (7.49) (which also holds for 𝒰l{\mathcal{U}}^{l} replaced by 𝒰{\mathcal{U}}):

    |Φ(𝒰(Ta0(a)))Φ(𝒰l(Ta0(a)))(Φ(𝒰(b))Φ(𝒰l(b)))|2KfH1ρ(a0)abL.|\Phi({\mathcal{U}}(T_{a_{0}}(a)))-\Phi({\mathcal{U}}^{l}(T_{a_{0}}(a)))-(\Phi({\mathcal{U}}(b))-\Phi({\mathcal{U}}^{l}(b)))|\leq 2K\frac{\|f\|_{H^{-1}}}{\rho(a_{0})}\|a-b\|_{L^{\infty}}.

By Theorem 4.11 and (7.50) we conclude that there exists δ>0\delta>0 and ξ2\xi_{2} independent of ll such that π~(𝒚|𝖉)π~l(𝒚|𝖉)\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})-\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) is (𝒃2,δ𝔴lα,ξ2,H01)({\boldsymbol{b}}_{2},\delta\mathfrak{w}_{l}^{-\alpha},\xi_{2},H_{0}^{1})-holomorphic. ∎

Items (ii) and (iii) of Proposition 7.12 show that Assumption 7.9 implies validity of Assumption 7.2. This in turn allows us to apply Theorems 7.5 and 7.6. Specifically, assuming the optimal convergence rate (7.41), we obtain that for every nn\in{\mathbb{N}} there is ε:=εn>0\varepsilon:=\varepsilon_{n}>0 such that work(𝐥ε)n\mathrm{work}(\mathbf{l}_{\varepsilon})\leq n and the multilevel interpolant 𝐈ML𝐥\mathbf{I}^{\rm ML}_{\mathbf{l}} defined in (7.3) satisfies

π~(|𝖉)𝐈𝐥εMLπ~(|𝖉)L2(U,H10;γ)C(1+logn)nRI,RI=min{s12,s12(1p132)s12+1p11p2}.\|\tilde{\pi}(\cdot|{\boldsymbol{\mathfrak{d}}})-\mathbf{I}_{\mathbf{l}_{\varepsilon}}^{\rm ML}\tilde{\pi}(\cdot|{\boldsymbol{\mathfrak{d}}})\|_{L^{2}(U,{H^{1}_{0}};\gamma)}\leq C(1+\log n)n^{-R_{I}},\quad R_{I}=\min\left\{\frac{s-1}{2},\frac{\frac{s-1}{2}(\frac{1}{p_{1}}-\frac{3}{2})}{\frac{s-1}{2}+\frac{1}{p_{1}}-\frac{1}{p_{2}}}\right\}.

Of higher practical interest is the application of the multilevel quadrature operator 𝐐ML\mathbf{Q}^{\rm ML} defined in (7.4). In case the prior is chosen as γ\gamma, then

Uπ~(𝒚|𝖉)dγ(𝒚)\int_{U}\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}})

equals the normalization constant in (5.2). It can be approximated with the error converging like

|Uπ~(𝒚|𝖉)dγ(𝒚)𝐐𝐥εMLu|C(1+logn)nRQ,RQ:=min{s12,s12(2p152)s12+2p12p2}.\left|\int_{U}\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathbf{Q}_{\mathbf{l}_{\varepsilon}}^{\rm ML}u\right|\leq C(1+\log n)n^{-R_{Q}},\quad R_{Q}:=\min\left\{\frac{s-1}{2},\frac{\frac{s-1}{2}(\frac{2}{p_{1}}-\frac{5}{2})}{\frac{s-1}{2}+\frac{2}{p_{1}}-\frac{2}{p_{2}}}\right\}. (7.51)

Typically, one is not merely interested in the normalization constant

Z=Uπ~(𝒚|𝖉)dγ(𝒚),Z=\int_{U}\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}}),

but for example also in an estimate of the jjth parameter yjy_{j} given as the conditional expectation, which up to multiplying with the normalization constant 1Z\frac{1}{Z}, corresponds to

Uyjπ~(𝒚|𝖉)dγ(𝒚).\int_{U}y_{j}\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}}).

Since 𝒚yj{\boldsymbol{y}}\mapsto y_{j} is analytic, one can show the same convergence rate as in (7.51) for the multilevel quadrature applied with the approximations 𝒚yjπ~l(𝒚|𝖉){\boldsymbol{y}}\mapsto y_{j}\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}}) for ll\in\mathbb{N}. Moreover, for example if ϕ:H01(D;)\phi:H_{0}^{1}({{D}};\mathbb{C})\to\mathbb{C} is a bounded linear functional representing some quantity of interest, then we can show the same error convergence for the approximation of

Uϕ(u(𝒚))π~(𝒚|𝖉)dγ(𝒚)\int_{U}\phi(u({\boldsymbol{y}}))\tilde{\pi}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}})

with the multilevel quadrature applied with the approximations ϕ(ul(𝒚))π~l(𝒚|𝖉)dγ(𝒚)\phi(u^{l}({\boldsymbol{y}}))\tilde{\pi}^{l}({\boldsymbol{y}}|{\boldsymbol{\mathfrak{d}}})\,\mathrm{d}\gamma({\boldsymbol{y}}) to the integrand for ll\in\mathbb{N}.

7.7 Linear multilevel interpolation and quadrature approximation

In this section, we briefly recall some results from [43] (see also [45] for some corrections). The difference with Sections 7.1 - 7.6 is, that the interpolation and quadrature operators presented in this section are linear operators; in contrast, the operators 𝐈ML\mathbf{I}^{\rm ML}, 𝐐ML\mathbf{Q}^{\rm ML} in (7.3), (7.4) are in general nonlinear, since they build on the approximations unu^{n} of uu from Assumption 7.2. These approximations are not assumed to be linear (and, in general, are not linear) in uu.

In this section we proceed similarly, but with un:=Pnuu^{n}:=P_{n}u for a linear operator PnP_{n}; if uu denotes the solution of an elliptic PDE in H1(D)H^{1}({{D}}), PnP_{n} could for instance be the orthogonal projection from H1(D)H^{1}({{D}}) into some fixed finite dimensional subspace. We emphasize, that such operators are not available in practice, and many widely used implementable algorithms (such as the finite element method, boundary element method, finite differences) realize projections that are not of this type. We will discuss this in more detail in Remark 7.31. Therefore the present results are mainly of theoretical rather than of practical importance. On a positive note, the convergence rates for both, Smolyak sparse-grid interpolation and quadrature obtained in this section via thresholding (see (7.54) ahead) improve the rates shown in the previous sections for the discretization levels allocated via Algorithm 2 by a logarithmic factor, cp. Theorems 7.5 and 7.6. Yet we emphasize that the latter are computable (in linear complexity, see Sec. 6.2.4).

7.7.1 Multilevel Smolyak sparse-grid interpolation

In this section, we recall some results in [43] (see also, [45] for some corrections) on linear multilevel polynomial interpolation approximation in Bochner spaces.

In order to have a correct definition of interpolation operator let us impose some necessary restrictions on vL2(U,X;γ)v\in L^{2}(U,X;\gamma). Let {\mathcal{E}} be a γ\gamma-measurable subset in UU such that γ()=1\gamma({\mathcal{E}})=1 and {\mathcal{E}} contains all 𝒚U{\boldsymbol{y}}\in U with |𝒚|0<|{\boldsymbol{y}}|_{0}<\infty, where |𝒚|0|{\boldsymbol{y}}|_{0} denotes the number of nonzero components yjy_{j} of 𝒚{\boldsymbol{y}}. For a given {\mathcal{E}} and separable Hilbert space XX, let C(U)C_{\mathcal{E}}(U) the set of all functions vv on UU taking values in XX such that vv are continuous on {\mathcal{E}} w.r. to the local convex topology of U:=U:={\mathbb{R}}^{\infty} (see Example 2.5). We define L2(U,X,γ):=L2(U,X;γ)C(U)L^{2}_{\mathcal{E}}(U,X,\gamma):=L^{2}(U,X;\gamma)\cap C_{\mathcal{E}}(U). We will treat all elements vL2(U,X,γ)v\in L^{2}_{\mathcal{E}}(U,X,\gamma) as their representative belonging to C(U)C_{\mathcal{E}}(U). Throughout this and next sections, we fix a set {\mathcal{E}}.

We define the univariate operator ΔIm\Delta^{{\rm I}}_{m} for m0m\in{\mathbb{N}}_{0} by

ΔIm:=ImIm1,\Delta^{{\rm I}}_{m}:=\ I_{m}-I_{m-1},

with the convention I1=0I_{-1}=0, where ImI_{m} is defined in Section 6.1.1.

For vL2(U,X;γ)v\in L^{2}_{\mathcal{E}}(U,X;\gamma), we introduce the tensor product operator ΔI𝝂\Delta^{{\rm I}}_{\boldsymbol{\nu}} for 𝝂{\boldsymbol{\nu}}\in\mathcal{F} by

ΔI𝝂(v):=jΔIνj(v),\Delta^{{\rm I}}_{\boldsymbol{\nu}}(v):=\ \bigotimes_{j\in{\mathbb{N}}}\Delta^{{\rm I}}_{\nu_{j}}(v),

where the univariate operator ΔIνj\Delta^{{\rm I}}_{\nu_{j}} is applied to the univariate function j=1j1ΔIνj(v)\bigotimes_{j^{\prime}=1}^{j-1}\Delta^{{\rm I}}_{\nu_{j^{\prime}}}(v) by considering this function as a function of variable yjy_{j} with all remaining variables held fixed. From the definition of L2(U,X;γ)L^{2}_{\mathcal{E}}(U,X;\gamma) one infers that the operators ΔI𝝂\Delta^{{\rm I}}_{\boldsymbol{\nu}} are well-defined for all 𝝂{\boldsymbol{\nu}}\in\mathcal{F}.

Let us recall a setting from [43] of linear fully discrete polynomial interpolation of functions in the Bochner space L2(U,X2;γ)L^{2}(U,X^{2};\gamma), with the approximation error measured by the norm of the Bochner space L2(U,X1;γ)L^{2}(U,X^{1};\gamma) for separable Hilbert spaces X1X^{1} and X2X^{2}. To construct linear fully discrete methods of polynomial interpolation, besides weighted 2\ell^{2}-summabilities with respect to X1X^{1} and X2X^{2} we need an approximation property on the spaces X1X^{1} and X2X^{2} combined in the following assumption.

Assumption 7.16.

For the Hilbert spaces X1X^{1} and X2X^{2} and vL2(U,X2;γ)v\in L^{2}_{\mathcal{E}}(U,X^{2};\gamma) represented by the series

v=𝝂v𝝂H𝝂,v𝝂X2,v=\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}v_{\boldsymbol{\nu}}H_{\boldsymbol{\nu}},\quad v_{\boldsymbol{\nu}}\in X^{2}, (7.52)

there holds the following.

  • (i)

    X2X^{2} is a linear subspace of X1X^{1} and X1CX2\|\cdot\|_{X^{1}}\leq C\,\|\cdot\|_{X^{2}}.

  • (ii)

    For i=1,2i=1,2, there exist numbers qiq_{i} with 0<q1q2<0<q_{1}\leq q_{2}<\infty and q1<2q_{1}<2, and families (σi;𝝂)𝝂(\sigma_{i;{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}} of numbers strictly larger than 11 such that σi;𝒆jσi;𝒆j\sigma_{i;{\boldsymbol{e}}_{j^{\prime}}}\leq\sigma_{i;{\boldsymbol{e}}_{j}} if j<jj^{\prime}<j, and

    𝝂(σi;𝝂v𝝂Xi)2Mi<and(p𝝂(τ,λ)σi;𝝂1)𝝂qi()\sum_{{\boldsymbol{\nu}}\in\mathcal{F}}(\sigma_{i;{\boldsymbol{\nu}}}\|v_{\boldsymbol{\nu}}\|_{X^{i}})^{2}\leq M_{i}<\infty\quad\text{and}\quad\left(p_{{\boldsymbol{\nu}}}(\tau,\lambda)\sigma_{i;{\boldsymbol{\nu}}}^{-1}\right)_{{\boldsymbol{\nu}}\in\mathcal{F}}\in\ell^{q_{i}}(\mathcal{F})

    for every τ>176\tau>\frac{17}{6} and λ0\lambda\geq 0, where we recall that (𝒆j)j({\boldsymbol{e}}_{j})_{j\in{\mathbb{N}}} is the standard basis of 2()\ell^{2}({\mathbb{N}}).

  • (iii)

    There are a sequence (Vn)n0(V_{n})_{n\in{\mathbb{N}}_{0}} of subspaces VnX1V_{n}\subset X^{1} of dimension n\leq n, and a sequence (Pn)n0(P_{n})_{n\in{\mathbb{N}}_{0}} of linear operators from X1X^{1} into VnV_{n}, and a number α>0\alpha>0 such that

    Pn(v)X1CvX1,vPn(v)X1CnαvX2,n0,vX2.\|P_{n}(v)\|_{X^{1}}\leq C\|v\|_{X^{1}},\quad\|v-P_{n}(v)\|_{X^{1}}\leq Cn^{-\alpha}\|v\|_{X^{2}},\quad\forall n\in{\mathbb{N}}_{0},\quad\forall v\in X^{2}. (7.53)

Let Assumption 7.16 hold for Hilbert spaces X1X^{1} and X2X^{2} and vL2(U,X2;γ)v\in L^{2}_{\mathcal{E}}(U,X^{2};\gamma). Then we are able to construct a linear fully discrete polynomial interpolation approximation. We introduce the interpolation operator

G:L2(U,X2;γ)𝒱(G){\mathcal{I}}_{G}:L^{2}_{\mathcal{E}}(U,X^{2};\gamma)\to{\mathcal{V}}(G)

for a given finite set G0×G\subset{\mathbb{N}}_{0}\times\mathcal{F} by

Gv:=(k,𝝂)GδkΔI𝝂(v),{\mathcal{I}}_{G}v:=\ \sum_{(k,{\boldsymbol{\nu}})\in G}\delta_{k}\Delta^{{\rm I}}_{\boldsymbol{\nu}}(v),

where 𝒱(G){\mathcal{V}}(G) denotes the subspace in L2(U,X1;γ)L^{2}(U,X^{1};\gamma) of all functions vv of the form

v=(k,𝝂)GvkH𝝂,vkV2k.v\ =\ \sum_{(k,{\boldsymbol{\nu}})\in G}v_{k}H_{\boldsymbol{\nu}},\quad v_{k}\in V_{2^{k}}.

Notice that interpolation vGvv\mapsto{\mathcal{I}}_{G}v is a linear method of fully discrete polynomial interpolation approximation, which is the sum taken over the (finite) index set GG, of mixed tensor products of dyadic scale successive differences of “spatial” approximations to vv, and of successive differences of their parametric Lagrange interpolation polynomials.

Define for ξ>0\xi>0

G(ξ):={{(k,𝝂)0×: 2kσ2;𝝂q2ξ}ifα1/q21/2;{(k,𝝂)0×:σ1;𝝂q1ξ, 2(α+1/2)kσ2;𝝂ξϑ}ifα>1/q21/2,G(\xi):=\ \begin{cases}\big{\{}(k,{\boldsymbol{\nu}})\in{\mathbb{N}}_{0}\times\mathcal{F}:\,2^{k}\sigma_{2;{\boldsymbol{\nu}}}^{q_{2}}\leq\xi\big{\}}\quad&{\rm if}\ \alpha\leq 1/q_{2}-1/2;\\ \big{\{}(k,{\boldsymbol{\nu}})\in{\mathbb{N}}_{0}\times\mathcal{F}:\,\sigma_{1;{\boldsymbol{\nu}}}^{q_{1}}\leq\xi,\ 2^{(\alpha+1/2)k}\sigma_{2;{\boldsymbol{\nu}}}\leq\xi^{\vartheta}\big{\}}\quad&{\rm if}\ \alpha>1/q_{2}-1/2,\end{cases} (7.54)

where

ϑ:=1q1+12α(1q11q2).\vartheta:=\frac{1}{q_{1}}+\frac{1}{2\alpha}\bigg{(}\frac{1}{q_{1}}-\frac{1}{q_{2}}\bigg{)}. (7.55)

For any ξ>1\xi>1 we have that G(ξ)F(ξ)G(\xi)\subset F(\xi) where

F(ξ):={(k,𝝂)0×:klogξ,𝝂Λ(ξ)}F(\xi):=\{(k,{\boldsymbol{\nu}})\in{\mathbb{N}}_{0}\times{\mathcal{F}}:k\leq\log\xi,\ {\boldsymbol{\nu}}\in\Lambda(\xi)\}

and

Λ(ξ):={{𝝂:σ2;𝝂q2ξ}ifα1/q21/2;{𝝂:σ1;𝝂q1ξ}ifα>1/q21/2.\Lambda(\xi):=\ \begin{cases}\big{\{}{\boldsymbol{\nu}}\in\mathcal{F}:\,\sigma_{2;{\boldsymbol{\nu}}}^{q_{2}}\leq\xi\big{\}}\quad&{\rm if}\ \alpha\leq 1/q_{2}-1/2;\\ \big{\{}{\boldsymbol{\nu}}\in\mathcal{F}:\,\sigma_{1;{\boldsymbol{\nu}}}^{q_{1}}\leq\xi\big{\}}\ \quad&{\rm if}\ \alpha>1/q_{2}-1/2.\end{cases}

From [46, Lemma 3.3] it follows that

𝝂Λ(ξ)supp(𝝂){1,,Cξ}\bigcup_{{\boldsymbol{\nu}}\in\Lambda(\xi)}\operatorname{supp}({\boldsymbol{\nu}})\subset\{1,...,\lfloor C\xi\rfloor\} (7.56)

for some positive constant CC that is independent of ξ>1\xi>1. Denote by Γ𝝂\Gamma_{\boldsymbol{\nu}} and Γ(Λ)\Gamma(\Lambda), the set of interpolation points in the operators ΔI𝝂\Delta^{{\rm I}}_{\boldsymbol{\nu}} and 𝐈Λ\mathbf{I}_{\Lambda}, respectively. We have that

Γ𝝂={𝒚𝝂𝒆;𝒎:𝒆𝔼𝝂;mj=0,,sjej,j},\Gamma_{\boldsymbol{\nu}}=\{{\boldsymbol{y}}_{{\boldsymbol{\nu}}-{\boldsymbol{e}};{\boldsymbol{m}}}:{\boldsymbol{e}}\in\mathbb{E}_{\boldsymbol{\nu}};\ m_{j}=0,\ldots,s_{j}-e_{j},\ j\in{\mathbb{N}}\},

and

Γ(Λ)=𝝂ΛΓ𝝂,\Gamma(\Lambda)=\bigcup_{{\boldsymbol{\nu}}\in\Lambda}\Gamma_{\boldsymbol{\nu}},

where 𝔼𝝂\mathbb{E}_{\boldsymbol{\nu}} is the subset in \mathcal{F} of all 𝒆{\boldsymbol{e}} such that eje_{j} is 11 or 0 if νj>0\nu_{j}>0, and eje_{j} is 0 if νj=0\nu_{j}=0, and 𝒚𝝂;𝒎:=(yνj;mj)j{\boldsymbol{y}}_{{\boldsymbol{\nu}};{\boldsymbol{m}}}:=(y_{\nu_{j};m_{j}})_{j\in{\mathbb{N}}}. Hence, by (7.56)

Γ(Λ(ξ))CξU,\Gamma(\Lambda(\xi))\subset{\mathbb{R}}^{\lfloor C\xi\rfloor}\subset U,

and therefore, the operator G(ξ){\mathcal{I}}_{G(\xi)} is well-defined for any vL2(U,X2;γ)v\in L^{2}_{\mathcal{E}}(U,X^{2};\gamma) since vv is continuous on Cξ{\mathbb{R}}^{\lfloor C\xi\rfloor}.

Theorem 7.17.

Let Assumption 7.16 hold for Hilbert spaces X1X^{1} and X2X^{2} and vL2(U,X2;γ)v\in L^{2}_{\mathcal{E}}(U,X^{2};\gamma). Then for each nn\in{\mathbb{N}}, there exists a number ξn\xi_{n} such that for the interpolation operator

G(ξn):L2(U,X2;γ)𝒱(G(ξn)),{\mathcal{I}}_{G(\xi_{n})}:L^{2}_{\mathcal{E}}(U,X^{2};\gamma)\to{\mathcal{V}}(G(\xi_{n})),

we have dim𝒱(G(ξn))n\dim{\mathcal{V}}(G(\xi_{n}))\leq n and

vG(ξn)vL2(U,X1;γ)Cnmin(α,β).\|v-{\mathcal{I}}_{G(\xi_{n})}v\|_{L^{2}(U,X^{1};\gamma)}\leq Cn^{-\min(\alpha,\beta)}. (7.57)

The rate α\alpha is as in (7.53) and the rate β\beta is given by

β:=(1q112)αα+δ,δ:=1q11q2.\beta:=\left(\frac{1}{q_{1}}-\frac{1}{2}\right)\frac{\alpha}{\alpha+\delta},\quad\delta:=\frac{1}{q_{1}}-\frac{1}{q_{2}}. (7.58)

The constant CC in (7.57) is independent of vv and nn.

Remark 7.18.

Observe that the operator G(ξn){\mathcal{I}}_{G(\xi_{n})} can be represented in the form of a multilevel Smolyak sparse-grid interpolation with knk_{n} levels:

G(ξn)=k=0knδk𝐈Λk(ξn),{\mathcal{I}}_{G(\xi_{n})}\ =\ \sum_{k=0}^{k_{n}}\delta_{k}\mathbf{I}_{\Lambda_{k}(\xi_{n})},

where kn:=log2ξnk_{n}:=\lfloor\log_{2}\xi_{n}\rfloor, the operator 𝐈Λ\mathbf{I}_{\Lambda} is defined as in (6.5), and for k0k\in{\mathbb{N}}_{0} and ξ>1\xi>1,

Λk(ξ):={{𝒔:σ2;𝒔q22kξ}ifα1/q21/2;{𝒔:σ1;𝒔q1ξ,σ2;𝒔2(α+1/2)kξϑ}ifα>1/q21/2.\Lambda_{k}(\xi):=\ \begin{cases}\big{\{}{\boldsymbol{s}}\in{\mathcal{F}}:\,\sigma_{2;{\boldsymbol{s}}}^{q_{2}}\leq 2^{-k}\xi\big{\}}\quad&{\rm if}\ \alpha\leq 1/q_{2}-1/2;\\ \big{\{}{\boldsymbol{s}}\in{\mathcal{F}}:\,\sigma_{1;{\boldsymbol{s}}}^{q_{1}}\leq\xi,\ \sigma_{2;{\boldsymbol{s}}}\leq 2^{-(\alpha+1/2)k}\xi^{\vartheta}\big{\}}\quad&{\rm if}\ \alpha>1/q_{2}-1/2.\end{cases}

In Theorem 7.17, the multilevel polynomial interpolation of vL2(U,X2;γ)v\in L^{2}_{\mathcal{E}}(U,X^{2};\gamma) by operators G(ξn){\mathcal{I}}_{G(\xi_{n})} is a collocation method. It is based on the finite point-wise information in 𝒚{\boldsymbol{y}}, more precisely, on |Γ(Λ0(ξn))|=𝒪(n)|\Gamma(\Lambda_{0}(\xi_{n}))|={\mathcal{O}}(n) of particular values of vv at the interpolation points 𝒚Γ(Λ0(ξn)){\boldsymbol{y}}\in\Gamma(\Lambda_{0}(\xi_{n})) and the approximations of v(𝒚)v({\boldsymbol{y}}), 𝒚Γ(Λ0(ξn)){\boldsymbol{y}}\in\Gamma(\Lambda_{0}(\xi_{n})), by P2kv(𝒚)P_{2^{k}}v({\boldsymbol{y}}) for k=0,,log2ξnk=0,\ldots,\lfloor\log_{2}\xi_{n}\rfloor with log2ξn=𝒪(log2n)\lfloor\log_{2}\xi_{n}\rfloor={\mathcal{O}}(\log_{2}n).

7.7.2 Multilevel Smolyak sparse-grid quadrature

In this section, we recall results of [43] (see also [45]) on linear methods for numerical integration of functions from Bochner spaces as well as their linear functionals. We define the univariate operator ΔQm\Delta^{{\rm Q}}_{m} for even m0m\in{\mathbb{N}}_{0} by

ΔQm:=QmQm2,\Delta^{{\rm Q}}_{m}:=\ Q_{m}-Q_{m-2},

with the convention Q2:=0Q_{-2}:=0. We make use of the notation:

ev:={𝝂:νjeven,j}.\mathcal{F}_{\operatorname{ev}}:=\{{\boldsymbol{\nu}}\in\mathcal{F}:\nu_{j}\ {\rm even},\ j\in{\mathbb{N}}\}.

For a function vL2(U,X;γ)v\in L^{2}_{\mathcal{E}}(U,X;\gamma), we introduce the operator ΔQ𝝂\Delta^{{\rm Q}}_{\boldsymbol{\nu}} defined for 𝝂ev{\boldsymbol{\nu}}\in\mathcal{F}_{\operatorname{ev}} by

ΔQ𝝂(v):=jΔQνj(v),\Delta^{{\rm Q}}_{\boldsymbol{\nu}}(v):=\ \bigotimes_{j\in{\mathbb{N}}}\Delta^{{\rm Q}}_{\nu_{j}}(v),

where the univariate operator ΔQνj\Delta^{{\rm Q}}_{\nu_{j}} is applied to the univariate function j=1j1ΔQνj(v)\bigotimes_{j^{\prime}=1}^{j-1}\Delta^{{\rm Q}}_{\nu_{j^{\prime}}}(v) by considering this function as a univariate function of yjy_{j}, with all other variables held fixed. As ΔI𝝂\Delta^{{\rm I}}_{\boldsymbol{\nu}}, the operators ΔQ𝝂\Delta^{{\rm Q}}_{\boldsymbol{\nu}} are well-defined for all 𝝂ev{\boldsymbol{\nu}}\in\mathcal{F}_{\operatorname{ev}}.

Letting Assumption 7.16 hold for Hilbert spaces X1X^{1} and X2X^{2}, we can construct linear fully discrete quadrature operators. For a finite set G0×evG\subset{\mathbb{N}}_{0}\times\mathcal{F}_{\operatorname{ev}}, we introduce the quadrature operator 𝒬G\mathcal{Q}_{G} which is defined for vv by

𝒬Gv:=(k,𝝂)GδkΔQ𝝂(v).\mathcal{Q}_{G}v:=\ \sum_{(k,{\boldsymbol{\nu}})\in G}\delta_{k}\Delta^{{\rm Q}}_{\boldsymbol{\nu}}(v). (7.59)

If ϕ(X1)\phi\in(X^{1})^{\prime} is a bounded linear functional on X1X^{1}, for a finite set G0×evG\subset{\mathbb{N}}_{0}\times\mathcal{F}_{\operatorname{ev}}, the quadrature formula 𝒬Gv\mathcal{Q}_{G}v generates the quadrature formula 𝒬Gϕ,v\mathcal{Q}_{G}\langle\phi,v\rangle for integration of ϕ,v\langle\phi,v\rangle by

𝒬Gϕ,v:=ϕ,𝒬Gv.\mathcal{Q}_{G}\langle\phi,v\rangle:=\ \langle\phi,\mathcal{Q}_{G}v\rangle.

Define for ξ>0\xi>0,

Gev(ξ):={{(k,𝝂)0×ev: 2kσ2;𝝂q2ξ}ifα1/q21/2;{(k,𝝂)0×ev:σ1;𝝂q1ξ, 2(α+1/2)kσ2;𝝂ξϑ}ifα>1/q21/2,G_{\operatorname{ev}}(\xi):=\ \begin{cases}\big{\{}(k,{\boldsymbol{\nu}})\in{\mathbb{N}}_{0}\times\mathcal{F}_{\operatorname{ev}}:\,2^{k}\sigma_{2;{\boldsymbol{\nu}}}^{q_{2}}\leq\xi\big{\}}\ &{\rm if}\ \alpha\leq 1/q_{2}-1/2;\\ \big{\{}(k,{\boldsymbol{\nu}})\in{\mathbb{N}}_{0}\times\mathcal{F}_{\operatorname{ev}}:\,\sigma_{1;{\boldsymbol{\nu}}}^{q_{1}}\leq\xi,\ 2^{(\alpha+1/2)k}\sigma_{2;{\boldsymbol{\nu}}}\leq\xi^{\vartheta}\big{\}}\ &{\rm if}\ \alpha>1/q_{2}-1/2,\end{cases} (7.60)

where ϑ\vartheta is as in (7.55).

Theorem 7.19.

Let the hypothesis of Theorem 7.17 hold. Then we have the following.

  • (i)

    For each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    Uv(𝒚)dγ(𝒚)𝒬Gev(ξn)vX1Cnmin(α,β).\left\|\int_{U}v({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}v\right\|_{X^{1}}\leq Cn^{-\min(\alpha,\beta)}. (7.61)
  • (ii)

    Let ϕ(X1)\phi\in(X^{1})^{\prime} be a bounded linear functional on X1X^{1}. Then for each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    |Uϕ,v(𝒚)dγ(𝒚)𝒬Gev(ξn)ϕ,v|Cϕ(X1)nmin(α,β).\left|\int_{U}\langle\phi,v({\boldsymbol{y}})\rangle\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}\langle\phi,v\rangle\right|\leq C\|\phi\|_{(X^{1})^{\prime}}n^{-\min(\alpha,\beta)}. (7.62)

The rate α\alpha is as in (7.53) and the rate β\beta is given by (7.58). The constants CC in (7.61) and (7.62) are independent of vv and nn.

The proof Theorem 7.19 are related to approximations in the norm of L1(U,X;γ)L^{1}(U,X;\gamma) by special polynomial interpolation operators which generate the corresponding quadrature operators. Let us briefly describe this connection, for details see [43, 45].

Remark 7.20.

We define the univariate interpolation operator ΔIm\Delta^{{\rm I}*}_{m} for even m0m\in{\mathbb{N}}_{0} by

ΔIm:=ImIm2,\Delta^{{\rm I}*}_{m}:=\ I_{m}-I_{m-2},

with the convention I2=0I_{-2}=0. The interpolation operators ΔI𝝂\Delta^{{\rm I}*}_{\boldsymbol{\nu}} for 𝝂ev{\boldsymbol{\nu}}\in\mathcal{F}_{\operatorname{ev}}, IΛI^{*}_{\Lambda} for a finite set Λev\Lambda\subset\mathcal{F}_{\operatorname{ev}}, and G{\mathcal{I}}^{*}_{G} for a finite set G0×evG\subset{\mathbb{N}}_{0}\times{\mathcal{F}}_{\operatorname{ev}}, are defined in a similar way as the corresponding quadrature operators ΔQ𝝂\Delta^{{\rm Q}}_{\boldsymbol{\nu}}, QΛQ_{\Lambda} and 𝒬G\mathcal{Q}_{G} by replacing ΔQνj\Delta^{{\rm Q}}_{\nu_{j}} with ΔIνj\Delta^{{\rm I}*}_{\nu_{j}}, jj\in{\mathbb{N}}.

From the definitions it follows the equalities expressing the relationship between the interpolation and quadrature operators

QΛv=UIΛv(𝒚)dγ(𝒚),QΛϕ,v=Uϕ,IΛv(𝒚)dγ(𝒚),Q_{\Lambda}v\ =\ \int_{U}I^{*}_{\Lambda}v({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}}),\quad Q_{\Lambda}\langle\phi,v\rangle\ =\ \int_{U}\langle\phi,I^{*}_{\Lambda}v({\boldsymbol{y}})\rangle\,\,\mathrm{d}\gamma({\boldsymbol{y}}),

and

𝒬Gv=UGv(𝒚)dγ(𝒚),𝒬Gϕ,v=Uϕ,Gv(𝒚)dγ(𝒚).\mathcal{Q}_{G}v\ =\ \int_{U}{\mathcal{I}}^{*}_{G}v({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}}),\quad\mathcal{Q}_{G}\langle\phi,v\rangle\ =\ \int_{U}\langle\phi,{\mathcal{I}}^{*}_{G}v({\boldsymbol{y}})\rangle\,\,\mathrm{d}\gamma({\boldsymbol{y}}).
Remark 7.21.

Similarly to G(ξn){\mathcal{I}}_{G(\xi_{n})}, the operator 𝒬Gev(ξn)\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})} can be represented in the form of a multilevel Smolyak sparse-grid quadrature with knk_{n} levels:

𝒬Gev(ξn)=k=0knδkQΛev,k(ξn),\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}\ =\ \sum_{k=0}^{k_{n}}\delta_{k}Q_{\Lambda_{\operatorname{ev},k}(\xi_{n})},

where kn:=log2ξnk_{n}:=\lfloor\log_{2}\xi_{n}\rfloor,

QΛ:=𝝂ΛΔI𝝂,Λev,Q_{\Lambda}:=\ \sum_{{\boldsymbol{\nu}}\in\Lambda}\Delta^{{\rm I}}_{\boldsymbol{\nu}},\ \ \Lambda\subset{\mathcal{F}}_{\operatorname{ev}}, (7.63)

and for k0k\in{\mathbb{N}}_{0} and ξ>0\xi>0,

Λev,k(ξ):={{𝒔ev:σ2;𝒔q22kξ}ifα1/q21/2;{𝒔ev:σ1;𝒔q1ξ,σ2;𝒔2(α+1/2)kξϑ}ifα>1/q21/2.\Lambda_{\operatorname{ev},k}(\xi):=\ \begin{cases}\big{\{}{\boldsymbol{s}}\in{\mathcal{F}}_{\operatorname{ev}}:\,\sigma_{2;{\boldsymbol{s}}}^{q_{2}}\leq 2^{-k}\xi\big{\}}\quad&{\rm if}\ \alpha\leq 1/q_{2}-1/2;\\ \big{\{}{\boldsymbol{s}}\in{\mathcal{F}}_{\operatorname{ev}}:\,\sigma_{1;{\boldsymbol{s}}}^{q_{1}}\leq\xi,\ \sigma_{2;{\boldsymbol{s}}}\leq 2^{-(\alpha+1/2)k}\xi^{\vartheta}\big{\}}\quad&{\rm if}\ \alpha>1/q_{2}-1/2.\end{cases}
Remark 7.22.

The convergence rates established in Theorems 7.17 and 7.19 and in Theorems 7.5 and 7.6 are proven with respect to different parameters nn as the dimension of the approximation space and the work (7.5), respectively. However, we could define the work of the operators G(ξn){\mathcal{I}}_{G(\xi_{n})} and 𝒬Gev(ξn)\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})} similarly as

k=0kn2k|Γ(Λk(ξn))|,\sum_{k=0}^{k_{n}}2^{k}|\Gamma(\Lambda_{k}(\xi_{n}))|,

and

k=0kn2k|Γ(Λev,k(ξn))|,\sum_{k=0}^{k_{n}}2^{k}|\Gamma(\Lambda_{\operatorname{ev},k}(\xi_{n}))|,

respectively, and prove the same convergence rates with respect to this work measure as in Theorems 7.17 and 7.19.

7.7.3 Applications to parametric divergence-form elliptic PDEs

In this section, we apply the results in Sections 7.7.1 and 7.7.2 to parametric divergence-form elliptic PDEs (3.17). The spaces VV and WW are as in Section 3.9.

Assumption 7.23.

There are a sequence (Vn)n0(V_{n})_{n\in{\mathbb{N}}_{0}} of subspaces VnVV_{n}\subset V of dimension m\leq m, and a sequence (Pn)n0(P_{n})_{n\in{\mathbb{N}}_{0}} of linear operators from VV into VnV_{n}, and a number α>0\alpha>0 such that

Pn(v)VCvV,vPn(v)VCnαvW,n0,vW.\|P_{n}(v)\|_{V}\leq C\|v\|_{V},\quad\|v-P_{n}(v)\|_{V}\leq Cn^{-\alpha}\|v\|_{W},\quad\forall n\in{\mathbb{N}}_{0},\quad\forall v\in W. (7.64)

If Assumption 7.23 and the assumptions of Theorem 3.38 hold for the spaces W1=VW^{1}=V and W2=WW^{2}=W with some 0<q1q2<0<q_{1}\leq q_{2}<\infty, then Assumption 7.16 holds for the spaces Xi=WiX^{i}=W^{i}, i=1,2i=1,2, and the solution uL2(U,X2;γ)u\in L^{2}(U,X^{2};\gamma) to (3.17)–(3.18). Hence we obtain the following results on multilevel (fully discrete) approximations.

Theorem 7.24.

Let Assumption 7.23 hold. Let the hypothesis of Theorem 3.38 hold for the spaces W1=VW^{1}=V and W2=WW^{2}=W with some 0<q1q2<0<q_{1}\leq q_{2}<\infty and q1<2q_{1}<2. For ξ>0\xi>0, let G(ξ)G(\xi) be the set defined by (7.54) for σi;𝛎\sigma_{i;{\boldsymbol{\nu}}} as in (3.59), i=1,2i=1,2. Let α\alpha be as in (7.64). Then for every nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(G(ξn))n\dim{\mathcal{V}}(G(\xi_{n}))\leq n and

uG(ξn)uL2(U,V;γ)Cnmin(α,β),\|u-{\mathcal{I}}_{G(\xi_{n})}u\|_{L^{2}(U,V;\gamma)}\leq Cn^{-\min(\alpha,\beta)}, (7.65)

where β\beta is given by (7.58). The constant CC in (7.65) is independent of uu and nn.

Theorem 7.25.

Let Assumption 7.23 hold. Let the assumptions of Theorem 3.38 hold for the spaces W1=VW^{1}=V and W2=WW^{2}=W for some 0<q1q2<0<q_{1}\leq q_{2}<\infty with q1<4q_{1}<4. Let α\alpha be the rate as given by (7.64). For ξ>0\xi>0, let Gev(ξ)G_{\operatorname{ev}}(\xi) be the set defined by (7.60) for σi;𝛎\sigma_{i;{\boldsymbol{\nu}}} as in (3.59), i=1,2i=1,2. Then we have the following.

  • (i)

    For each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    Uv(𝒚)dγ(𝒚)𝒬Gev(ξn)vVCnmin(α,β).\left\|\int_{U}v({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}v\right\|_{V}\leq Cn^{-\min(\alpha,\beta)}. (7.66)
  • (ii)

    Let ϕV\phi\in V^{\prime} be a bounded linear functional on VV. Then for each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    |Uϕ,v(𝒚)dγ(𝒚)𝒬Gev(ξn)ϕ,v|CϕVnmin(α,β).\left|\int_{U}\langle\phi,v({\boldsymbol{y}})\rangle\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}\langle\phi,v\rangle\right|\leq C\|\phi\|_{V^{\prime}}n^{-\min(\alpha,\beta)}. (7.67)

The rate β\beta is given by

β:=(2q112)αα+δ,δ:=2q12q2.\beta:=\left(\frac{2}{q_{1}}-\frac{1}{2}\right)\frac{\alpha}{\alpha+\delta},\quad\delta:=\frac{2}{q_{1}}-\frac{2}{q_{2}}.

The constants CC in (7.66) and (7.67) are independent of uu and nn.

Proof.

From Theorem 3.38, Lemma 3.39 and Assumption 7.23 we can see that the assumptions of Theorem 7.17 hold for X1=VX^{1}=V and X2=WX^{2}=W with 0<q1/2q2/2<0<q_{1}/2\leq q_{2}/2<\infty and q1/2<2q_{1}/2<2. Hence, by applying Theorem 7.19 we prove the theorem. ∎

7.7.4 Applications to holomorphic functions

As noticed, the proof of the weighted 2\ell_{2}-summability result formulated in Theorem 3.38 employs bootstrap arguments and induction on the differentiation order of derivatives with respect to the parametric variables, for details see [8, 9]. In the log-Gaussian case, this approach and technique are too complicated and difficult for extension to more general parametric PDE problems, in particular, of higher regularity. As it has been seen in the previous sections, the approach to a unified summability analysis of Wiener-Hermite PC expansions of various scales of function spaces based on parametric holomorphy, covers a wide range of parametric PDE problems. In this section, we apply the results in Sections 7.7.1 and 7.7.2 on linear approximations and integration in Bochner spaces to approximation and numerical integration of parametric holomorphic functions based on weighted 2\ell^{2}-summabilities of the coefficient sequences of the Wiener-Hermite PC expansion.

The following theorem on weighted 2\ell_{2}-summability for (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions can be derived from Theorem 4.9 and Lemma 3.39.

Theorem 7.26.

Let vv be (𝐛,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic for some 𝐛p(){\boldsymbol{b}}\in\ell^{p}(\mathbb{N}) with 0<p<10<p<1. Let s=1,2s=1,2 and τ,λ0\tau,\lambda\geq 0. Let further the sequence ϱ=(ϱj)j{\boldsymbol{\varrho}}=(\varrho_{j})_{j\in{\mathbb{N}}} be defined by

ϱj:=bjp1ξ4r!𝒃p.\varrho_{j}:=b_{j}^{p-1}\frac{\xi}{4\sqrt{r!}}\|{\boldsymbol{b}}\|_{\ell^{p}}.

Then, for any r>2s(τ+1)qr>\frac{2s(\tau+1)}{q},

𝝂s(σ𝝂v𝝂X)2M<and(p𝝂(τ,λ)σ𝝂1)𝝂sq/s(s),\sum_{{\boldsymbol{\nu}}\in\mathcal{F}_{s}}(\sigma_{{\boldsymbol{\nu}}}\|v_{\boldsymbol{\nu}}\|_{X})^{2}\leq M<\infty\quad\text{and}\quad\left(p_{{\boldsymbol{\nu}}}(\tau,\lambda)\sigma_{{\boldsymbol{\nu}}}^{-1}\right)_{{\boldsymbol{\nu}}\in\mathcal{F}_{s}}\in\ell^{q/s}(\mathcal{F}_{s}),

where q:=p1pq:=\frac{p}{1-p}, M:=δ2C(𝐛)M:=\delta^{2}C({\boldsymbol{b}}) and (σ𝛎)𝛎(\sigma_{\boldsymbol{\nu}})_{{\boldsymbol{\nu}}\in\mathcal{F}} with σ𝛎:=β𝛎(r,ϱ)1/2\sigma_{\boldsymbol{\nu}}:=\beta_{\boldsymbol{\nu}}(r,{\boldsymbol{\varrho}})^{1/2}.

To treat multilevel approximations and integration of parametric, holomorphic functions, it is appropriate to replace Assumption 7.16 by its modification.

Assumption 7.27.

Assumption 7.16 holds with item (ii) replaced with item

  • (ii’)

    For i=1,2i=1,2, vv is (𝒃i,ξ,δ,Xi)({\boldsymbol{b}}_{i},\xi,\delta,X^{i})-holomorphic for some 𝒃ipi(){\boldsymbol{b}}_{i}\in\ell^{p_{i}}(\mathbb{N}) with 0<p1p2<10<p_{1}\leq p_{2}<1.

Assumption 7.27 is a condition for fully discrete approximation of (𝒃,ξ,δ,X)({\boldsymbol{b}},\xi,\delta,X)-holomorphic functions. This is formalized in the following corollary of Theorem 7.26.

Corollary 7.28.

Assumption 7.27 implies Assumption 7.16 for qi:=pi1piq_{i}:=\frac{p_{i}}{1-p_{i}} and (σi;𝛎)𝛎(\sigma_{i;{\boldsymbol{\nu}}})_{{\boldsymbol{\nu}}\in\mathcal{F}}, i=1,2i=1,2, where

σi;𝝂:=βi;𝝂(r,ϱi)1/2,ϱi;j:=bi;jpi1ξ4r!𝒃ipi.\sigma_{i;{\boldsymbol{\nu}}}:=\beta_{i;{\boldsymbol{\nu}}}(r,{\boldsymbol{\varrho}}_{i})^{1/2},\quad\varrho_{i;j}:=b_{i;j}^{p_{i}-1}\frac{\xi}{4\sqrt{r!}}\|{\boldsymbol{b}}_{i}\|_{\ell^{p_{i}}}.

We formulate results on multilevel quadrature of parametric holomorphic functions as consequences of Corollary 7.28 and Theorems 7.17 and 7.19.

Theorem 7.29.

Let Assumption 7.27 hold for the Hilbert spaces X1X^{1} and X2X^{2} with p1<2/3p_{1}<2/3, and vL2(U,X2;γ)v\in L^{2}(U,X^{2};\gamma). For ξ>0\xi>0, let G(ξ)G(\xi) be the set defined by (7.54) for σi;𝛎\sigma_{i;{\boldsymbol{\nu}}}, i=1,2i=1,2 as given in Corollary 7.28. Then for every nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(G(ξn))n\dim{\mathcal{V}}(G(\xi_{n}))\leq n and

vG(ξn)vL2(U,X1;γ)CnR,\|v-{\mathcal{I}}_{G(\xi_{n})}v\|_{L^{2}(U,X^{1};\gamma)}\leq Cn^{-R}, (7.68)

where RR is given by the formula (7.19) and the constant CC in(7.68) is independent of vv and nn.

Theorem 7.30.

Let Assumption 7.27 hold for the Hilbert spaces X1X^{1} and X2X^{2} with p1<4/5p_{1}<4/5, and vL2(U,X2;γ)v\in L^{2}(U,X^{2};\gamma). For ξ>0\xi>0, let Gev(ξ)G_{\operatorname{ev}}(\xi) be the set defined by (7.60) for σi;𝛎\sigma_{i;{\boldsymbol{\nu}}}, i=1,2i=1,2, as given in Corollary 7.28. Then we have the following.

  • (i)

    For each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    Uv(𝒚)dγ(𝒚)𝒬Gev(ξn)vX1CnR.\left\|\int_{U}v({\boldsymbol{y}})\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}v\right\|_{X^{1}}\leq Cn^{-R}. (7.69)
  • (ii)

    Let ϕ(X1)\phi\in(X^{1})^{\prime} be a bounded linear functional on X1X^{1}. Then for each nn\in{\mathbb{N}} there exists a number ξn\xi_{n} such that dim𝒱(Gev(ξn))n\dim{\mathcal{V}}(G_{\operatorname{ev}}(\xi_{n}))\leq n and

    |Uϕ,v(𝒚)dγ(𝒚)𝒬Gev(ξn)ϕ,v|Cϕ(X1)nR,\left|\int_{U}\langle\phi,v({\boldsymbol{y}})\rangle\,\,\mathrm{d}\gamma({\boldsymbol{y}})-\mathcal{Q}_{G_{\operatorname{ev}}(\xi_{n})}\langle\phi,v\rangle\right|\leq C\|\phi\|_{(X^{1})^{\prime}}n^{-R}, (7.70)

where the convergence rate RR is given by the formula (7.26) and the constants CC in (7.69) and (7.70) are independent of vv and nn.

Remark 7.31.

We comment on the relation of the results of Theorems 7.5 and 7.6 to the results of [43] which are presented in Theorems 7.24 and 7.25, on multilevel approximation of solutions to parametric divergence-form elliptic PDEs with log-Gaussian inputs.

Specifically, in [43], by combining spatial and parametric approximability in the spatial domain and weighted 2\ell^{2}-summability of the V:=H10(D)V:=H^{1}_{0}({D}) and WW norms of Wiener-Hermite PC expansion coefficients obtained in [9, 8], the author constructed linear non-adaptive methods of fully discrete approximation by truncated Wiener-Hermite PC expansion and polynomial interpolation approximation as well as fully discrete weighted quadrature for parametric and stochastic elliptic PDEs with log-Gaussian inputs, and proved the convergence rates of approximation by them. The results in [43] are based on Assumption 7.23 that requires the existence of a sequence (Pn)n0(P_{n})_{n\in{\mathbb{N}}_{0}} of linear operators independent of 𝐲{\boldsymbol{y}}, from H10(D)H^{1}_{0}({{D}}) into nn-dimensional subspaces VnH10(D)V_{n}\subset H^{1}_{0}({{D}}) such that

Pn(v)H10C1vH10andvPn(v)H10C2nαvW\|P_{n}(v)\|_{H^{1}_{0}}\leq C_{1}\|v\|_{H^{1}_{0}}\ \ \text{and}\ \ \|v-P_{n}(v)\|_{H^{1}_{0}}\leq C_{2}n^{-\alpha}\|v\|_{W}

for all n0n\in{\mathbb{N}}_{0} and for all vWv\in W, where the constants C1,C2C_{1},C_{2} are independent of nn. The assumption of PnP_{n} being independent of 𝐲{\boldsymbol{y}} is however typically not satisfied if Pn(u(𝐲))=un(𝐲)P_{n}(u({\boldsymbol{y}}))=u^{n}({\boldsymbol{y}}) is a numerical approximation to u(𝐲)u({\boldsymbol{y}}) (such as, e.g., a Finite-Element or a Finite-Difference discretization).

In contrast, the present approximation rate analysis is based on quantified, parametric holomorphy of the discrete approximations ulu^{l} to uu as in Assumption 7.2. For example, assume that u:UH01(D)u:U\to H_{0}^{1}({{D}}) is the solution of the parametric PDE

div(a(𝒚)u(𝒚))=f-\operatorname{div}(a({\boldsymbol{y}})\nabla u({\boldsymbol{y}}))=f

for some fL2(D)f\in L^{2}({{D}}) and a parametric diffusion coefficient a(𝒚)L(D)a({\boldsymbol{y}})\in L^{\infty}({{D}}) such that

essinf𝒙Da(𝒚,𝒙)>0𝒚U.\underset{{\boldsymbol{x}}\in{{D}}}{\operatorname{ess\,inf}}\,a({\boldsymbol{y}},{\boldsymbol{x}})>0\ \ \ \forall{\boldsymbol{y}}\in U.

Then ul:UH01(D)u^{l}:U\to H_{0}^{1}({{D}}) could be a numerical approximation to uu, such as the FEM solution: for every ll\in\mathbb{N} there is a finite dimensional discretization space XlH01(D)X_{l}\subseteq H_{0}^{1}({{D}}), and

Dul(𝒚)a(𝒚)vd𝒙=Dfvd𝒙\int_{{{D}}}\nabla u^{l}({\boldsymbol{y}})^{\top}a({\boldsymbol{y}})\nabla v\,\mathrm{d}{\boldsymbol{x}}=\int_{{{D}}}fv\,\mathrm{d}{\boldsymbol{x}}

for every vXlv\in X_{l} and for every 𝒚U{\boldsymbol{y}}\in U. Hence ul(𝒚)u^{l}({\boldsymbol{y}}) is the orthogonal projection of u(𝒚)u({\boldsymbol{y}}) onto XlX_{l} w.r.t. the inner product

v,wa(𝒚):=Dva(𝒚)wd𝒙\langle v,w\rangle_{a({\boldsymbol{y}})}:=\int_{{D}}\nabla v^{\top}a({\boldsymbol{y}})\nabla w\,\mathrm{d}{\boldsymbol{x}}

on H01(D)H_{0}^{1}({{D}}). We may write this as ul(𝒚)=Pl(𝒚)u(𝒚)u^{l}({\boldsymbol{y}})=P_{l}({\boldsymbol{y}})u({\boldsymbol{y}}), for a 𝒚{\boldsymbol{y}}-dependent projector

Pl(𝒚):H01(D)Xl.P_{l}({\boldsymbol{y}}):H_{0}^{1}({{D}})\to X_{l}.

This situation is covered by Assumption 7.2.

The preceding comments can be extended to the results on multilevel approximation of holomorphic functions in Theorems 7.5 and 7.6 to the results in Theorems 7.29 and 7.30. On the other hand, as noticed above, the convergence rates in Theorems 7.29 and 7.30 are slightly better than those obtained in Theorems 7.5 and 7.6.

8 Conclusions

We established holomorphy of parameter-to-solution maps

Eau=𝒰(a)XE\ni a\mapsto u={\mathcal{U}}(a)\in X

for linear, elliptic, parabolic, and other PDEs in various scales of function spaces EE and XX, including in particular standard and corner-weighted Sobolev spaces. Our discussion focused on non-compact parameter domains which arise from uncertain inputs from function spaces expressed in a suitable basis with Gaussian distributed coefficients. We introduced and used a form of quantified, parametric holomorphy in products of strips to show that this implies summability results of coefficients of the Wiener-Hermite PC expansion of such infinite parametric functions. Specifically, we proved weighted 2\ell^{2}-summability and p\ell^{p}-summability results for Wiener-Hermite PC expansions of certain parametric, deterministic solution families {u(𝒚):𝒚U}X\{u({\boldsymbol{y}}):{\boldsymbol{y}}\in U\}\subset X, for a given “log-affine” parametrization (3.18) of admissible random input data aEa\in E.

We introduced and analyzed constructive, deterministic, sparse-grid (“stochastic collocation”) algorithms based on univariate Gauss-Hermite points, to efficiently sample the parametric, deterministic solutions in the possibly infinite-dimensional parameter domain U=U={\mathbb{R}}^{\infty}. The sparsity of the coefficients of Wiener-Hermite PC expansion was shown to entail corresponding convergence rates of the presently developed sparse-grid sampling schemes. In combination with suitable Finite Element discretizations in the physical, space(-time) domain (which include proper mesh-refinements to account for singularities in the physical domain) we proved convergence rates for abstract, multilevel algorithms which employ different combinations of sparse-grid interpolants in the parametric domain with space(-time) discretizations at different levels of accuracy in the physical domain.

The presently developed, holomorphic setting was also shown to apply to the corresponding Bayesian inverse problems subject to PDE constraints: here, the density of the Bayesian posterior with respect to a Gaussian random field prior was shown to generically inherit quantified holomorphy from the parametric forward problem, thereby facilitating the use of the developed sparse-grid collocation and integration algorithms also for the efficient deterministic computation of Bayesian estimates of PDEs with uncertain inputs, subject to noisy observation data.

Our approximation rate bounds are free from the curse-of-dimensionality and only limited by the PC coefficient summability. They will therefore also be relevant for convergence rate analyses of other approximation schemes, such as Gaussian process emulators or neural networks (see, e.g., [104, 46, 44, 99] and references there).


Acknowledgments. The work of Dinh Dũng and Van Kien Nguyen is funded by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under Grant No. 102.01-2020.03. A part of this work was done when Dinh Dũng was visiting the Forschungsinstitut für Mathematik (FIM) of the ETH Zürich invited by Christoph Schwab, and when Dinh Dũng and Van Kien Nguyen were at the Vietnam Institute for Advanced Study in Mathematics (VIASM). They would like to thank the FIM and VIASM for providing a fruitful research environment and working condition. Dinh Dũng thanks Christoph Schwab for invitation to visit the FIM and for his hospitality.

References

  • [1] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965.
  • [2] R. A. Adams and J. J. F. Fournier, Sobolev Spaces, second ed., Pure and Applied Mathematics (Amsterdam), vol. 140, Elsevier/Academic Press, Amsterdam, 2003.
  • [3] R. J. Adler, The Geometry of Random Fields, John Wiley & Sons, Ltd., Chichester, 1981, Wiley Series in Probability and Mathematical Statistics.
  • [4] R. Andreev and A. Lang, Kolmogorov-Chentsov theorem and differentiability of random fields on manifolds, Potential Anal. 41 (2014), no. 3, 761–769.
  • [5] A. Ayache and M. S. Taqqu, Rate optimality of wavelet series approximations of fractional Brownian motion, J. Fourier Anal. Appl. 9 (2003), no. 5, 451–471.
  • [6] I. Babuška, R. B. Kellogg, and J. Pitkäranta, Direct and inverse error estimates for finite elements with mesh refinements, Numer. Math. 33 (1979), no. 4, 447–471.
  • [7] 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), no. 3, 1005–1034.
  • [8] M. Bachmayr, A. Cohen, D. Dũng, and Ch. Schwab, Fully discrete approximation of parametric and stochastic elliptic PDEs, SIAM J. Numer. Anal. 55 (2017), no. 5, 2151–2186.
  • [9] M. Bachmayr, A. Cohen, R. DeVore, and G. Migliorati, Sparse polynomial approximation of parametric elliptic PDEs. Part II: Lognormal coefficients, ESAIM Math. Model. Numer. Anal. 51 (2017), no. 1, 341–363.
  • [10] M. Bachmayr, A. Cohen, and G. Migliorati, Sparse polynomial approximation of parametric elliptic PDEs. Part I: affine coefficients, ESAIM Math. Model. Numer. Anal. 51 (2017), 321–339.
  • [11] M. Bachmayr, A. Cohen, and G. Migliorati, Sparse polynomial approximation of parametric elliptic PDEs. Part I: Affine coefficients, ESAIM Math. Model. Numer. Anal. 51 (2017), no. 1, 321–339.
  • [12]  , Representations of Gaussian random fields and approximation of elliptic PDEs with lognormal coefficients, J. Fourier Anal. Appl. 24 (2018), no. 3, 621–649.
  • [13] M. Bachmayr and A. Djurdjevac, Multilevel representations of isotropic Gaussian random fields on the sphere, 2022, drac034, https://doi.org/10.1093/imanum/drac034.
  • [14] M. Bachmayr, I. G. Graham, V. K. Nguyen, and R. Scheichl, Unified analysis of periodization-based sampling methods for Matérn covariances, SIAM J. Numer. Anal. 58 (2020), no. 5, 2953–2980.
  • [15] M. Bachmayr and V. K. Nguyen, Identifiability of diffusion coefficients for source terms of non-uniform sign, Inverse Probl. Imaging 13 (2019), no. 5, 1007–1021.
  • [16] J. Beck, R. Tempone, F. Nobile, and L. Tamellini, On the optimal polynomial approximation of stochastic PDEs by Galerkin and collocation methods, Math. Models Methods Appl. Sci. 22 (2012), no. 9, 1250023, 33.
  • [17] R. Bellman, Dynamic Programming, Princeton Landmarks in Mathematics, Princeton University Press, Princeton, NJ, 2010, Reprint of the 1957 edition, With a new introduction by Stuart Dreyfus.
  • [18] J. Bergh and J. Löfström, Interpolation Spaces. An Introduction, Springer-Verlag, Berlin-New York, 1976, Grundlehren der Mathematischen Wissenschaften, No. 223.
  • [19] Marcel Bieri, Roman Andreev, and Christoph Schwab, Sparse tensor discretization of elliptic SPDEs, SIAM J. Sci. Comput. 31 (2009/10), no. 6, 4281–4304. MR 2566594
  • [20] D. Boffi, F. Brezzi, and M. Fortin, Mixed Finite Element Methods and Applications, Springer Series in Computational Mathematics, vol. 44, Springer, Heidelberg, 2013.
  • [21] V. I. Bogachev, Gaussian Measures, Mathematical Surveys and Monographs, vol. 62, American Mathematical Society, Providence, RI, 1998.
  • [22] A. Bonito, A. Demlow, and J. Owen, A priori error estimates for finite element approximations to eigenvalues and eigenfunctions of the Laplace-Beltrami operator, SIAM J. Numer. Anal. 56 (2018), no. 5, 2963–2988.
  • [23] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, third ed., Texts in Applied Mathematics, vol. 15, Springer, New York, 2008.
  • [24] C. Băcuţă, H. Li, and V. Nistor, Differential operators on domains with conical points: precise uniform regularity estimates, Rev. Roum. Math. Pure Appl. 62 (2017), 383–411.
  • [25] C. Băcuţă, V. Nistor, and L. T. Zikatanov, Improving the rate of convergence of ‘high order finite elements’ on polygons and domains with cusps, Numer. Math. 100 (2005), no. 2, 165–184.
  • [26]  , Improving the rate of convergence of high-order finite elements on polyhedra. I. A priori estimates, Numer. Funct. Anal. Optim. 26 (2005), no. 6, 613–639.
  • [27]  , Improving the rate of convergence of high-order finite elements on polyhedra. II. Mesh refinements and interpolation, Numer. Funct. Anal. Optim. 28 (2007), no. 7-8, 775–824.
  • [28] J. Charrier, Strong and weak error estimates for elliptic partial differential equations with random coefficients, SIAM J. Numer. Anal. 50 (2012), no. 1, 216–246.
  • [29] J. Charrier and A. Debussche, Weak truncation error estimates for elliptic PDEs with lognormal coefficients, Stoch. Partial Differ. Equ. Anal. Comput. 1 (2013), no. 1, 63–93.
  • [30] 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), no. 1, 322–352. MR 3033013
  • [31] P. Chen, Sparse quadrature for high-dimensional integration with Gaussian measure, ESAIM Math. Model. Numer. Anal. 52 (2018), no. 2, 631–657.
  • [32] A. Chkifa, A. Cohen, and Ch. Schwab, High-dimensional adaptive sparse polynomial interpolation and applications to parametric PDEs, Found. Comput. Math. 14 (2014), no. 4, 601–633.
  • [33] P. G. Ciarlet, The Finite Element Method for Elliptic Problems, North-Holland Publishing Co., Amsterdam-New York-Oxford, 1978, Studies in Mathematics and its Applications, Vol. 4.
  • [34] Z. Ciesielski, Hölder conditions for realizations of Gaussian processes, Trans. Amer. Math. Soc. 99 (1961), 403–413.
  • [35] G. Cleanthous, A. G. Georgiadis, A. Lang, and E. Porcu, Regularity, continuity and approximation of isotropic Gaussian random fields on compact two-point homogeneous spaces, Stochastic Process. Appl. 130 (2020), no. 8, 4873–4891.
  • [36] K. A. 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), no. 1, 3–15. MR 2835612
  • [37] A. Cohen and R. DeVore, Approximation of high-dimensional parametric PDEs, Acta Numer. 24 (2015), 1–159.
  • [38] A. Cohen, R. Devore, and Ch. Schwab, Convergence rates of best NN-term Galerkin approximations for a class of elliptic sPDEs, Found. Comput. Math. 10 (2010), no. 6, 615–646.
  • [39]  , Analytic regularity and polynomial approximation of parametric and stochastic elliptic PDE’s, Anal. Appl. (Singap.) 9 (2011), no. 1, 11–47.
  • [40] A. Cohen, Ch. Schwab, and J. Zech, Shape holomorphy of the stationary Navier-Stokes Equations, SIAM J. Math. Anal. 50 (2018), no. 2, 1720–1752.
  • [41] R. R. Coifman and M. Maggioni, Diffusion wavelets for multiscale analysis on graphs and manifolds, Wavelets and splines: Athens 2005, Mod. Methods Math., Nashboro Press, Brentwood, TN, 2006, pp. 164–188.
  • [42] D. Dũng, Linear collocation approximation for parametric and stochastic elliptic PDEs, Sb. Math. 210 (2019), no. 4, 565–588.
  • [43]  , Sparse-grid polynomial interpolation approximation and integration for parametric and stochastic elliptic PDEs with lognormal inputs, ESAIM Math. Model. Numer. Anal. 55 (2021), 1163–1198.
  • [44]  , Collocation approximation by deep neural ReLU networks for parametric elliptic PDEs with lognormal inputs, Sb. Math. 214 (2023), no. 4.
  • [45]  , Erratum to: “Sparse-grid polynomial interpolation approximation and integration for parametric and stochastic elliptic PDEs with lognormal inputs”, [Erratum to: ESAIM: M2AN 55(2021) 1163–1198], ESAIM Math. Model. Numer. Anal. 57 (2023), 893–897.
  • [46] D. Dũng, V. K. Nguyen, and D. T. Pham, Deep ReLU neural network approximation of parametric and stochastic elliptic PDEs with lognormal inputs, 2021, arXiv:2111.05854.
  • [47] M. Dashti, K. J. H. Law, A. M. Stuart, and J. Voss, MAP estimators and their consistency in Bayesian nonparametric inverse problems, Inverse Probl. 29 (2013), no. 9, 095017, 27.
  • [48] M. Dashti and A. M. Stuart, The Bayesian approach to inverse problems, Handbook of uncertainty quantification. Vol. 1, 2, 3, Springer, Cham, 2017, pp. 311–428.
  • [49] A. Demlow, Higher-order finite element methods and pointwise error estimates for elliptic problems on surfaces, SIAM J. Numer. Anal. 47 (2009), no. 2, 805–827.
  • [50] C. R. Dietrich and G. N. Newsam, Fast and exact simulation of stationary Gaussian processes through circulant embedding of the covariance matrix, SIAM J. Sci. Comput. 18 (1997), no. 4, 1088–1107.
  • [51] J. Dölz, H. Harbrecht, and Ch. Schwab, Covariance regularity and \cal{H}-matrix approximation for rough random fields, Numer. Math. 135 (2017), no. 4, 1045–1071.
  • [52] O. G. Ernst, B. Sprungk, and L. Tamellini, Convergence of sparse collocation for functions of countably many Gaussian random variables (with application to elliptic PDEs), SIAM J. Numer. Anal. 56 (2018), no. 2, 877–905.
  • [53] L. C. Evans, Partial Differential Equations, second ed., Graduate Studies in Mathematics, vol. 19, American Mathematical Society, Providence, RI, 2010.
  • [54] F. D. Gaspoz and P. Morin, Convergence rates for adaptive finite elements, IMA J. Numer. Anal. 29 (2009), no. 4, 917–936.
  • [55] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics in Mathematics, Springer-Verlag, Berlin, 2001, Reprint of the 1998 edition.
  • [56] E. Giné and R. Nickl, Mathematical Foundations of Infinite-dimensional Statistical Models, Cambridge Series in Statistical and Probabilistic Mathematics, [40], Cambridge University Press, New York, 2016.
  • [57] C. Gittelson, J. Könnö, Ch. Schwab, and R. Stenberg, The multi-level Monte Carlo finite element method for a stochastic Brinkman problem, Numer. Math. 125 (2013), no. 2, 347–386.
  • [58] C. J. Gittelson, Representation of Gaussian fields in series with independent coefficients, IMA J. Numer. Anal. 32 (2012), no. 1, 294–319.
  • [59] I. G. Graham, F. Y. Kuo, J. A. Nichols, R. Scheichl, Ch. Schwab, and I. H. Sloan, Quasi-Monte Carlo finite element methods for elliptic PDEs with lognormal random coefficients, Numer. Math. 131 (2014), 329–368.
  • [60] I. G. Graham, F. Y. Kuo, D. Nuyens, R. Scheichl, and I.H. Sloan, Analysis of circulant embedding methods for sampling stationary random fields, SIAM J. Numer. Anal. 56 (2018), 1871–1895.
  • [61] P. Grisvard, Elliptic Problems in Nonsmooth Domains, Classics in Applied Mathematics, vol. 69, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2011, Reprint of the 1985 original [ MR0775683], With a foreword by Susanne C. Brenner.
  • [62] B. Q. Guo and I. Babuška, On the regularity of elasticity problems with piecewise analytic data, Adv. in Appl. Math. 14 (1993), no. 3, 307–347.
  • [63] H. Harbrecht, M. Peters, and M. Siebenmorgen, Multilevel accelerated quadrature for PDEs with log-normally distributed diffusion coefficient, SIAM-ASA J. Uncertain. Quantif. 4 (2016), no. 1, 520–551.
  • [64] G. H. Hardy, Note on a theorem of Hilbert, Math. Z. 6 (1920), no. 3-4, 314–317.
  • [65] Ch. Heil, A Basis Theory Primer, expanded ed., Applied and Numerical Harmonic Analysis, Birkhäuser/Springer, New York, 2011.
  • [66] L. Herrmann, M. Keller, and Ch. Schwab, Quasi-Monte Carlo Bayesian estimation under Besov priors in elliptic inverse problems, Math. Comp. 90 (2021), no. 330, 1831–1860.
  • [67] L. Herrmann and Ch. Schwab, Multilevel quasi-Monte Carlo integration with product weights for elliptic PDEs with lognormal coefficients, ESAIM Math. Model. Numer. Anal. 53 (2019), no. 5, 1507–1552.
  • [68]  , QMC integration for lognormal-parametric, elliptic PDEs: local supports and product weights, Numer. Math. 141 (2019), no. 1, 63–102.
  • [69] E. Hille, Contributions to the theory of Hermitian series. II. The representation problem, Trans. Amer. Math. Soc. 47 (1940), 80–94.
  • [70] V. H. Hoang, J. H. Quek, and Ch. Schwab, Analysis of a multilevel Markov chain Monte Carlo finite element method for Bayesian inversion of log-normal diffusions, Inverse Probl. 36 (2020), no. 3, 035021, 46.
  • [71] V. H. Hoang and Ch. Schwab, Sparse tensor Galerkin discretization of parametric and random parabolic PDEs—analytic regularity and generalized polynomial chaos approximation, SIAM J. Math. Anal. 45 (2013), no. 5, 3050–3083.
  • [72]  , NN-term Wiener chaos approximation rate for elliptic PDEs with lognormal Gaussian random inputs, Math. Models Methods Appl. Sci. 24 (2014), no. 4, 797–826.
  • [73]  , Convergence rate analysis of MCMC-FEM for Bayesian inversion of log-normal diffusion problems, Tech. Report 2016-19, Seminar for Applied Mathematics, ETH Zürich, Switzerland, 2016.
  • [74] J. Indritz, An inequality for Hermite polynomials, Proc. Amer. Math. Soc. 12 (1961), 981–983.
  • [75] S. Janson, Gaussian Hilbert Spaces, Cambridge Tracts in Mathematics, vol. 129, Cambridge University Press, Cambridge, 1997.
  • [76] D. Kamilis and N. Polydorides, Uncertainty quantification for low-frequency, time-harmonic Maxwell equations with stochastic conductivity models, SIAM-ASA J. Uncertain. Quantif. 6 (2018), no. 4, 1295–1334.
  • [77] K. Karhunen, Zur Spektraltheorie stochastischer Prozesse, Ann. Acad. Sci. Fennicae Ser. A. I. Math.-Phys. 1946 (1946), no. 34, 7.
  • [78] Y. Kazashi, Quasi-Monte Carlo integration with product weights for elliptic PDEs with log-normal coefficients, IMA J. Numer. Anal. 39 (2019), no. 3, 1563–1593.
  • [79] G. Kerkyacharian, S. Ogawa, P. Petrushev, and D. Picard, Regularity of Gaussian processes on Dirichlet spaces, Constr. Approx. 47 (2018), no. 2, 277–320.
  • [80] M. Kohr, S. Labrunie, H. Mohsen, and V. Nistor, Polynomial estimates for solutions of parametric elliptic equations on complete manifolds, Stud. Univ. Babeş-Bolyai Math. 67 (2022), no. 2, 369–382.
  • [81] V. A. Kozlov and J. Rossmann, Asymptotics of solutions of second order parabolic equations near conical points and edges, Bound. Value Probl. (2014), 2014:252, 37.
  • [82] A. Kufner, Weighted Sobolev Spaces, Teubner-Texte zur Mathematik [Teubner Texts in Mathematics], vol. 31, BSB B. G. Teubner Verlagsgesellschaft, Leipzig, 1980, With German, French and Russian summaries.
  • [83] F. Y. Kuo, R. Scheichl, Ch. Schwab, I. H. Sloan, and E. Ullmann, Multilevel Quasi-Monte Carlo methods for lognormal diffusion problems, Math. Comp. 86 (2017), no. 308, 2827–2860.
  • [84] A. Lang and Ch. Schwab, Isotropic Gaussian random fields on the sphere: regularity, fast simulation and stochastic partial differential equations, Ann. Appl. Probab. 25 (2015), no. 6, 3047–3094.
  • [85] M. Lassas, E. Saksman, and S. Siltanen, Discretization-invariant Bayesian inversion and Besov space priors, Inverse Probl. Imaging 3 (2009), no. 1, 87–122.
  • [86] H. Li, An anisotropic finite element method on polyhedral domains: interpolation error analysis, Math. Comp. 87 (2018), no. 312, 1567–1600.
  • [87] M. A. Lifshits, Gaussian Random Functions, Mathematics and its Applications, vol. 322, Kluwer Academic Publishers, Dordrecht, 1995.
  • [88] H. Luschgy and G. Pagès, Expansions for Gaussian processes and Parseval frames, Electron. J. Probab. 14 (2009), no. 42, 1198–1221.
  • [89] B. Matérn, Spatial Variation, second ed., Lecture Notes in Statistics, vol. 36, Springer-Verlag, Berlin, 1986, With a Swedish summary.
  • [90] V. Maz’ya and J. Rossmann, Elliptic Equations in Polyhedral Domains, Mathematical Surveys and Monographs, vol. 162, American Mathematical Society, Providence, RI, 2010.
  • [91] G. A. Muñoz, Y. Sarantopoulos, and A. Tonge, Complexifications of real Banach spaces, polynomials and multilinear maps, Studia Math. 134 (1999), no. 1, 1–33.
  • [92] F. J. Narcowich, P. Petrushev, and J. D. Ward, Localized tight frames on spheres, SIAM J. Math. Anal. 38 (2006), no. 2, 574–594.
  • [93] J. Nečas, Sur une méthode pour résoudre les équations aux dérivées partielles du type elliptique, voisine de la variationnelle, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3) 16 (1962), 305–326.
  • [94] R. Nickl, Bayesian non-linear statistical inverse problems, Lecture Notes ETH Zurich, EMS Publ., 2023 (to appear).
  • [95] Victor Nistor, Hassan Mohsen, and Simon Labrunie, Polynomial bounds for the solutions of parametric transmission problems on smooth, bounded domains, working paper or preprint, February 2023.
  • [96] R. Scheichl, A. M. Stuart, and A. L. Teckentrup, Quasi-Monte Carlo and multilevel Monte Carlo methods for computing posterior expectations in elliptic inverse problems, SIAM-ASA J. Uncertain. Quantif. 5 (2017), no. 1, 493–518.
  • [97] Ch. Schwab and R. Stevenson, Space-time adaptive wavelet methods for parabolic evolution problems, Math. Comput. 78 (2009), 1293–1318.
  • [98] Ch. Schwab and A. M. Stuart, Sparse deterministic approximation of Bayesian inverse problems, Inverse Probl. 28 (2012), no. 4, 045003, 32.
  • [99] Ch. Schwab and J. Zech, Deep learning in high dimension: Neural network expression rates for analytic functions in L2(d,γd){L}^{2}(\mathbb{R}^{d},\gamma_{d}), SIAM-ASA J. Uncertain. Quantif. 11 (2023), no. 1, 199–234.
  • [100] M. L. Stein, Interpolation of Spatial Data, Springer Series in Statistics, Springer-Verlag, New York, 1999, Some theory for Kriging.
  • [101] I. Steinwart and C. Scovel, Mercer’s theorem on general domains: on the interaction between measures, kernels, and RKHSs, Constr. Approx. 35 (2012), no. 3, 363–417.
  • [102] S. B. Stečkin, On absolute convergence of orthogonal series, Dokl. Akad. Nauk SSSR (N.S.) 102 (1955), 37–40.
  • [103] A. M. Stuart and A. L. Teckentrup, Posterior consistency for Gaussian process approximations of Bayesian posterior distributions, Math. Comp. 87 (2018), no. 310, 721–753.
  • [104]  , Posterior consistency for Gaussian process approximations of Bayesian posterior distributions, Math. Comp. 87 (2018), no. 310, 721–753.
  • [105] G. Szegő, Orthogonal Polynomials, fourth ed., American Mathematical Society, Providence, R.I., 1975, American Mathematical Society, Colloquium Publications, Vol. XXIII.
  • [106] A. L. Teckentrup, R. Scheichl, M. B. Giles, and E. Ullmann, Further analysis of multilevel Monte Carlo methods for elliptic PDEs with random coefficients, Numer. Math. 125 (2013), no. 3, 569–600. MR 3117512
  • [107] H. Triebel, Interpolation Theory, Function Spaces, Differential Operators, second ed., Johann Ambrosius Barth, Heidelberg, 1995.
  • [108] C. Truesdell, A First Course in Rational Continuum Mechanics. Vol. 1, second ed., Pure and Applied Mathematics, vol. 71, Academic Press, Inc., Boston, MA, 1991, General concepts.
  • [109] N. Wiener, The Homogeneous Chaos, Amer. J. Math. 60 (1938), no. 4, 897–936.
  • [110] J. Wloka, Partial Differential Equations, Cambridge University Press, Cambridge, 1987, Translated from the German by C. B. Thomas and M. J. Thomas.
  • [111] J. Zech, Sparse-Grid Approximation of High-Dimensional Parametric PDEs, 2018, Dissertation 25683, ETH Zürich, http://dx.doi.org/10.3929/ethz-b-000340651.
  • [112] J. Zech, D. Dũng, and Ch. Schwab, Multilevel approximation of parametric and stochastic PDEs, Math. Models Methods Appl. Sci. 29 (2019), 1753–1817.
  • [113] J. Zech and Y. Marzouk, Sparse approximation of triangular transports on bounded domains, 2020, arXiv:2006.06994.
  • [114] J. Zech and Ch. Schwab, Convergence rates of high-dimensional Smolyak quadrature, ESAIM: Math. Model. Numer. Anal. 54 (2020), no. 4, 1259–1307.

List of symbols

Cs(D)C^{s}({D}) Space of ss-Hölder continuous functions on D{D}
Δ\Delta Laplace operator
div\operatorname{div} Divergence operator
D{D} Domain in d\mathbb{R}^{d}
D\partial{D} Boundary of the domain D{D}
𝒆j{\boldsymbol{e}}_{j} The multiindex (δij)j0(\delta_{ij})_{j\in\mathbb{N}}\in\mathbb{N}_{0}^{\infty}
γ\gamma Gaussian measure
γd\gamma_{d} Gaussian measure on d\mathbb{R}^{d}
\nabla Gradient operator
H(γ)H(\gamma) Cameron-Martin space of the Gaussian measure γ\gamma
HkH_{k} kkth normalized probabilistic Hermite polynomial
Hs(D)H^{s}({D}) Ws,2(D)W^{s,2}({D})
H10(D)H^{1}_{0}({D}) Space of functions uW1,2(D)u\in W^{1,2}({D}) such that u|D=0u|_{\partial{D}}=0
Hs0(D)H^{s}_{0}({{D}}) Hs(D)H10(D)H^{s}({{D}})\cap H^{1}_{0}({D})
H1(D)H^{-1}({{D}}) Dual space of H10(D)H^{1}_{0}({D})
𝐈Λ\mathbf{I}_{\Lambda} Smolyak interpolation operator
𝐈ML𝐥\mathbf{I}^{\rm ML}_{\mathbf{l}} Multilevel Smolyak interpolation operator
(z)\Im(z) Imaginary part of the complex number zz
𝒦sϰ(D){\mathcal{K}}^{s}_{\varkappa}({D}) Space of functions u:Du:{D}\to{\mathbb{C}} such that rD|𝜶|ϰD𝜶uL2(D)r_{D}^{|{\boldsymbol{\alpha}}|-\varkappa}D^{\boldsymbol{\alpha}}u\in L^{2}({D}) for all |𝜶|s|{\boldsymbol{\alpha}}|\leq s
λd\lambda_{d} Lebesgue measure on d\mathbb{R}^{d}
Λ\Lambda Set of multiindices
μ^\hat{\mu} Fourier transform of the measure μ\mu
Lp(Ω)L^{p}(\Omega) Space of Lebesgue measurable, pp-integrable functions on Ω\Omega
Lp(Ω,μ)L^{p}(\Omega,\mu) Space of μ\mu-measurable, pp-integrable functions on Ω\Omega
Lp(Ω,X;μ)L^{p}(\Omega,X;\mu) Space of functions u:ΩXu:\Omega\to X such that uXLp(Ω,μ)\|u\|_{X}\in L^{p}(\Omega,\mu)
L(Ω)L^{\infty}(\Omega) Space of Lebesgue measurable, essentially bounded functions on Ω\Omega
p(I)\ell^{p}(I) Space of sequences (yj)jI(y_{j})_{j\in I} such that (jI|yj|p)1/p<(\sum_{j\in I}|y_{j}|^{p})^{1/p}<\infty
fX\|f\|_{X} Norm of ff in the space XX
𝐐Λ\mathbf{Q}_{\Lambda} Smolyak quadrature operator
𝐐ML𝐥\mathbf{Q}^{\rm ML}_{\mathbf{l}} Multilevel Smolyak quadrature operator
rDr_{D} Smooth function D+{D}\to{\mathbb{R}}_{+} which equals |𝒙𝒄||{\boldsymbol{x}}-{\boldsymbol{c}}| in the vicinity of each corner of DD
(z)\Re(z) Real part of the complex number zz
UU \mathbb{R}^{\infty}
Ws,q(D)W^{s,q}({D}) Sobolev spaces of integer order ss and integrability qq on D{D}
𝒲s(D){\mathcal{W}}^{s}_{\infty}({D}) Space of functions u:Du:{D}\to{\mathbb{C}} such that rD|𝜶|D𝜶uL(D)r_{D}^{|{\boldsymbol{\alpha}}|}D^{\boldsymbol{\alpha}}u\in L^{\infty}({D}) for all |𝜶|s|{\boldsymbol{\alpha}}|\leq s

List of abbreviations

BIPBIP Bayesian inverse problems
BVPBVP boundary value problem
FEFE finite element
FEMFEM finite element method
GMGM Gaussian measure
GRFGRF Gaussian random fields
IBVPIBVP initial boundary value problem
KLKL Karhunen-Loève
MCMC Monte-Carlo
ONBONB orthonormal basis
PCPC polynomial chaos
PDEPDE partial differential equations
QMCQMC quasi-Monte Carlo
RKHSRKHS reproducing kernel Hilbert space
RVRV random variable
UQUQ uncertainty quantification
\printindex