39
Numerical solutions of stochastic PDEs driven by arbitrary type of noise 1 Tianheng Chen 2 , Boris Rozovskii 3 and Chi-Wang Shu 4 Abstract So far the theory and numerical practice of stochastic partial differential equations (SPDEs) have dealt almost exclusively with Gaussian noise or L´ evy noise. Recently, Mikule- vicius and Rozovskii (2016) [22] proposed a distribution-free Skorokhod-Malliavin calculus framework that is based on generalized stochastic polynomial chaos expansion, and is com- patible with arbitrary driving noise. In this paper, we conduct systematic investigation on numerical results of these newly developed distribution-free SPDEs, exhibiting the efficacy of truncated polynomial chaos solutions in approximating moments and distributions. We obtain an estimate for the mean square truncation error in the linear case. The theoreti- cal convergence rate, also verified by numerical experiments, is exponential with respect to polynomial order and cubic with respect to number of random variables included. Keywords: distribution-free, stochastic PDE, stochastic polynomial chaos, Wick product, Skorokhod integral 1 Research supported by ARO grant W911NF-16-1-0103. 2 Division of Applied Mathematics, Brown University, Providence, RI 02912. E-mail: tianheng [email protected] 3 Division of Applied Mathematics, Brown University, Providence, RI 02912. E-mail: boris [email protected] 4 Division of Applied Mathematics, Brown University, Providence, RI 02912. E-mail: [email protected] 1

Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Numerical solutions of stochastic PDEs driven by arbitrary type of noise1

Tianheng Chen2, Boris Rozovskii3 and Chi-Wang Shu4

Abstract

So far the theory and numerical practice of stochastic partial differential equations

(SPDEs) have dealt almost exclusively with Gaussian noise or Levy noise. Recently, Mikule-

vicius and Rozovskii (2016) [22] proposed a distribution-free Skorokhod-Malliavin calculus

framework that is based on generalized stochastic polynomial chaos expansion, and is com-

patible with arbitrary driving noise. In this paper, we conduct systematic investigation on

numerical results of these newly developed distribution-free SPDEs, exhibiting the efficacy

of truncated polynomial chaos solutions in approximating moments and distributions. We

obtain an estimate for the mean square truncation error in the linear case. The theoreti-

cal convergence rate, also verified by numerical experiments, is exponential with respect to

polynomial order and cubic with respect to number of random variables included.

Keywords: distribution-free, stochastic PDE, stochastic polynomial chaos, Wick product,

Skorokhod integral

1Research supported by ARO grant W911NF-16-1-0103.2Division of Applied Mathematics, Brown University, Providence, RI 02912. E-mail:

tianheng [email protected] of Applied Mathematics, Brown University, Providence, RI 02912. E-mail:

boris [email protected] of Applied Mathematics, Brown University, Providence, RI 02912. E-mail: [email protected]

1

Page 2: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

1 Introduction

Stochastic partial differential equations (SPDEs) and stochastic ordinary differential equa-

tions (SODEs) describe functions driven by finite or infinite dimensional stochastic processes.

SPDEs have found a broad range of applications, including mathematical biology, financial

engineering and nonlinear filtering, to quantify the intrinsic uncertainty in these models (see

e.g., Lototsky and Rozovskii [16]). Since analytical solutions to SPDEs are rarely avail-

able, numerical methods have to be designed to solve them. The most popular approach

is the Monte Carlo method, which generates independent random sample paths via direct

discretization or averaging over characteristic lines (see e.g., Milstein and Tretyakov [23]).

However, it is well known that Monte Carlo simulation suffers from slow rate of convergence

and often requires millions of samples to reach a desired accuracy level. Various techniques

have been established to accelerate convergence and alleviate computational burden, e.g.,

importance sampling and quasi Monte Carlo sampling. For more detailed discussion on nu-

merical methods of SPDEs, we refer readers to the books [24, 37] and the references therein.

An alternative numerical approach is based on the truncation of stochastic polynomial

chaos expansion, which is well-developed for SPDEs driven by Gaussian randomness (see

e.g., reference books [8, 7]). Such methodology starts with the orthogonal representation

of Gaussian white noise by a series of i.i.d standard Gaussian random variables. According

to the Cameron-Martin theorem [3], products of Hermite polynomials of these Gaussian

variables constitute a complete L2 basis of the probability space (also known as Hermite

chaos or Wiener chaos in the literature). Moreover, the multiple Ito integral formula [10]

helps recast stochastic integrals into the Wick product form [29], leading to a system of

deterministic PDEs, called propagator (see e.g., Lototsky and Rozovskii [14]), satisfied by

the expansion coefficients. Then standard numerical solvers can be applied to the propagator

system. The Wiener chaos expansion approach separates the stochastic part (basis functions)

and the deterministic part (coefficients), and is usually recognized as the stochastic Galerkin

method (see e.g., Xiu [33]) due to its evident resemblance with the spectral Galerkin method

in numerical analysis.

For linear parabolic SPDEs, the propagator system has lower triangular structure and

hence can be solved sequentially [19, 12, 20, 14]. The authors in [12] also derived an estimate

for the truncation error, showing that the mean square error converges exponentially with

respect to expansion order and linearly with respect to number of random variables included,

but grows exponentially as time grows. A recursive multistage modification was then de-

veloped in [12] to enable long time integration and investigated numerically in [34]. As a

counterpart, a recursive version of the stochastic collocation method was thoroughly studied

2

Page 3: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

in [35, 36] as well. On the other hand, the propagator system for nonlinear SPDEs is fully

coupled and requires substantially more computational power [9, 17]. Wick-Malliavin ap-

proximation has been proposed in [21] as a decoupling technique. Numerical simulations were

performed for nonlinear problems and elliptic SPDEs with random coefficients in [28, 27] to

validate the efficacy of Wick-Malliavin approximation for polynomial nonlinearity and small

noise.

Although the stochastic polynomial chaos expansion is mostly restricted to Gaussian

random noise, the extension towards arbitrary type of noise has been explored over the last

two decades. Xiu and Karniadakis [30, 31] constructed the correspondence between types

of random variables and orthogonal polynomials in the Askey scheme [1]. Their generalized

polynomial chaos (gPC) technique turned out to be very successful for problems with random

initial/boundary condition and/or random coefficients [31, 32, 33]. However, gPC expansion

is not naturally compatible with stochastic integrals due to the lack of some vital connections,

e.g., Wick product, Skorokhod integral [26] and Malliavin calculus [18]. These building blocks

are indeed available for Levy randomness [8, 7], and Liu [11] presented some numerical

results of SPDEs driven by Poisson noise. Recently, Mikulevicius and Rozovsky built the

distribution-free Skorokhod-Malliavin calculus framework [22] directly upon gPC expansion,

giving rise to a new family of SPDEs under their arbitrary noise paradigm. For linear SPDEs,

the propagator system is independent of the type of randomness involved, and so are the

first and second moments (this is already noticed by Benth and Gjerde [2] for Poisson noise);

meanwhile, for nonlinear problems, the propagator system does vary from one type of noise

to another.

The main objective of this paper is to numerically testify the stochastic polynomial chaos

approach to solving the distribution-free SPDEs in [22]. The rest of this paper is structured

as follows. In Section 2 we summarize the distribution-free stochastic analysis infrastructure

in [22], and provide propagator systems for linear and nonlinear model problems. In Section 3

we improve the truncation error estimate in [12] for linear problems, proving that convergence

rate of mean square error is actually cubic with respect to the number of random variables.

Theorem 3.1 is a major contribution of this paper. Section 4 performs numerical experiments

on linear and nonlinear SPDEs. We will study numerical orders of convergence and the

claims in Section 3 are verified. Exact solutions or reference solutions are obtained through

some alternative approaches, including Monte Carlo simulation, moment equations and the

Fokker-Planck equation. We conduct comparisons with reference solutions, and between

different types of driving noise. Concluding remarks and possible future research directions

are given in Section 5. The computation of interaction coefficients appeared in the nonlinear

propagator system is shown in the appendix.

3

Page 4: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

2 Distribution-free stochastic analysis

Let (Ω,F , P) be a probability space and H = L2([0, T ]) for some T > 0. We define the

following driving noise N(t):

N(t) =

∞∑

k=1

mk(t)ξk (2.1)

and the stochastic process

N(t) =

∫ t

0

N(s)ds =

∞∑

k=1

( ∫ t

0

mk(s)ds)ξk (2.2)

where Ξ = ξk∞k=1 is a sequence of uncorrelated random variables with E[ξk] = 0 and

E[ξ2k] = 1 for each k ≥ 1, and mk∞k=1 is a complete orthonormal basis in H . The term

distribution-free arises from the fact that each random variable can be of any distribution.

They are not required to be identically distributed or independent. Then we aim to construct

an orthogonal basis in L2(Ω, σ(Ξ), P), in a sense that is similar to the Wiener chaos expansion

for Gaussian noise.

2.1 Polynomial chaos expansion

Denote by J the set of multi-indices α = (αk)∞k=1 of finite length |α| =

∑∞k=1 αk:

J = α = (αk)∞k=1 : αk ≥ 0, |α| <∞

Each multi-index α with |α| = n can be uniquely identified by its characteristic set Iα =

(i1α, i2α, · · · , inα), which is a vector of length n and given by

imα = k if and only if

k−1∑

j=1

αj < m ≤k∑

j=1

αj, for each 1 ≤ m ≤ n

For instance, if α = (0, 1, 0, 2, 3, 0, · · ·), Iα = (2, 4, 4, 5, 5, 5). Particularly, we let d(α) = inα,

the position of the rightmost nonzero entry in α. We also define polynomials and factorials

of multi-indices:

ξα :=∞∏

k=1

ξαk

k , α! :=∞∏

k=1

αk!

The following two assumptions are required for stochastic polynomial chaos expansion.

A1. For each finite dimensional random vector (ξi1, ξi2, · · · , ξid), the moment generating

function E[exp(θ1ξi1 + θ2ξi2 + · · ·+ θdξid)] exists for all (θ1, θ2, · · · , θd) in some neighborhood

of 0 ∈ Rd.

4

Page 5: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

A2. We have an orthogonalization Kα, α ∈ J of the system of polynomials ξα, α ∈ J such that for each n ≥ 1, Kp, |p| ≤ n spans the same linear subspace Hn as ξp : |p| ≤ n,and for each |α| = n + 1,

Kα = ξα − projectionHnξα

We rescale the basis functions Φα = cαKα so that E[ΦαΦβ ] = δαβ(α!). Obviously Φε0= 1

and Φεk= ξk where ε0 is the multi-index whose entries are all zero, and εk is the multi-index

such that its k-th entry is 1 and all other entries are zero for k ≥ 1. It is demonstrated in

[22] that Φα, α ∈ J is indeed a complete Cameron-Martin type basis.

Proposition 2.1. [22] Assume A1 and A2 hold. Then Φα, α ∈ J is a complete orthog-

onal basis of L2(Ω, σ(Ξ), P). For each η ∈ L2(Ω, σ(Ξ), P), its stochastic polynomial chaos

expansion is

η =∑

α∈J

ηαΦα, ηα =E[ηΦα]

α!

and the Parseval’s identity holds:

E[η2] =∑

α∈J

(α!)η2α

In such manner we separate the random part and the deterministic part.

For the special case where ξk∞k=1 are i.i.d. random variables, if there exists an orthog-

onal set of univariate polynomials ϕn(ξ1)∞n=0 such that E[ϕn(ξ1)ϕm(ξ1)] = δmnn!, and the

moment generating function of ξ1 is finite near zero, the polynomial chaos basis functions

are simply products of these univariate polynomials

Φα =

∞∏

k=1

ϕαk(ξk) (2.3)

Then A1 and A2 are satisfied and Φα, α ∈ J forms a complete orthogonal basis. Table

2.1 shows some common types of random distributions and their corresponding polynomial

chaos basis functions (see e.g., [30]). It may be necessary to shift and rescale ξk∞k=1 to

achieve zero mean and unit variance. Two specific examples are provided below.

Example 2.1.1. If ξk∞k=1 are i.i.d. standard Gaussian random variables, ϕn(ξ1)∞n=0

are probabilists’ Hermite polynomials Hen(x)∞n=0. Here E[ϕ2n(ξ)] = n! is automatically

satisfied. The driving noise N(t) =∑∞

k=1 mk(t)ξk is indeed the Gaussian white noise W (t),

and the driving process N(t) is the standard Wiener process W (t) (see e.g., [8]).

5

Page 6: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Table 2.1: Correspondence between random distribution and polynomial chaos basis func-tions for an i.i.d. sequence of random variables.

Distribution of ξ1 Polynomial chaos basisContinuous Gaussian Hermite

Gamma LagurreBeta JacobiUniform Legendre

Discrete Poisson CharlierBinomial KrawtchoukNegative binomial MeixnerHypergeometric Hahn

Example 2.1.2. If ξk∞k=1 are i.i.d. uniformly distributed on [−√

3,√

3], ϕn(ξ1)∞n=0 are

the rescaled version of Legendre polynomials Ln(x)∞n=0:

ϕn(ξ1) =√

(2n + 1)n!Ln

( ξ1√3

)(2.4)

to ensure E[ϕ2n(ξ1)] = n!. Analogous to the Wiener process, the driving stochastic process

N(t) has zero mean and uncorrelated increments. However, N(t) is non-Gaussian as its

characteristic function is

E[iθN(t)] = E

[exp

(iθ

∞∑

k=1

( ∫ t

0

mk(s)ds)ξk

)]=

∞∏

k=1

sin(√

3θ( ∫ t

0mk(s)ds

))

√3θ

( ∫ t

0mk(s)ds

)

Remark 2.1. We intentionally use the weaker assumption of uncorrelated random variables

to incorporate Levy randomness, whose chaos expansion basis functions are not polynomials

of simple random variables. Such complexity is beyond the scope of this paper, and we will

always consider i.i.d. random variables in our numerical experiments.

2.2 Wick product and Skorokhod integral

This subsection briefly explains the languages to construct distribution-free stochastic inte-

grals. For the opposite direction, i.e. the stochastic (Malliavin) derivative, we refer interested

readers to [22] for more details. We will work on the space of generalized random variables

written as formal chaos expansion series:

D′(E) :=

u =∑

α∈J

uαΦα : uα ∈ E

and the space of square integrable general random variables:

D(E) :=u =

α∈J

uαΦα : uα ∈ E,∑

α∈J

α!‖uα‖2E <∞

6

Page 7: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

where E is a given Hilbert space. For instance, if E = H = L2([0, T ]), D′(H) consists of

generalized stochastic processes u = u(t) =∑

α∈J uα(t)Φα such that each uα ∈ L2([0, T ]),

andD(H) equals D′(H)∩L2(Ω×[0, T ]), the subspace of square integrable stochastic processes

in D′(H).

Let us first introduce the Wick product ⋄, a convolution type binary operator on expan-

sion coefficients:

Φα ⋄ Φβ = Φα+β, u ⋄ v =∑

α∈J

β∈J

uαvβΦα+β for u, v ∈ D′(R)

Then for u = u(t) ∈ D′(H), its Skorokhod integral δ(u) ∈ D′(R) is denoted by

δ(u) :=

∫ T

0

u(t) ⋄ N(t)dt =∑

α∈J

∞∑

k=1

(∫ T

0

uα(t)mk(t)dt)Φα+εk

(2.5)

Especially, if u = u(t) ∈ H is a deterministic function,

δ(u) =∞∑

k=1

(∫ T

0

u(t)mk(t)dt)ξk, E[δ(u)2] =

(∫ T

0

u(t)mk(t)dt)2

= ‖u‖2H (2.6)

Therefore, δ can be regarded as an isometric embedding from H into L2(Ω).

Proposition 2.2. Suppose u ∈ C([0, T ]) is continuous. Consider a sequence of partitions

of [0, T ]:

∆i = [ti,j−1, ti,j] : 1 ≤ j ≤ i− 1, ti,0 = 0, ti,i = T

such that |∆i| → 0. We have

i∑

j=1

u(ti,j−1)(N(ti,j)−N(ti,j−1))→ δ(u)

in L2(Ω). In other words, δ(u) is the limit of discrete sums in Ito’s sense.

Proof. In fact,

i∑

j=1

u(ti,j−1)(N(ti,j)−N(ti,j−1)) =i∑

j=1

u(ti,j−1)( ∞∑

k=1

( ∫ ti,j

ti,j−1

mk(t)dt)ξk

):= δ(ui)

where

ui(t) =i∑

j=1

1[ti,j−1,ti,j)(t)u(ti,j−1)

and 1 is the indicator function. As |∆i| → 0, it is easy to verify that ui → u in H . Hence

according to the isometric property, δ(ui) converges to δ(u) in L2(Ω).

7

Page 8: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

There is an equivalent way to characterize the Skorokhod integral in terms of multiple

integrals. For nonnegative n, let Hn := L2([0, T ]n) and Hn be the family of symmetric

functions in Hn. We use t(n) as the short hand notation of (t1, t2, · · · , tn). For each multi-

index α with |α| = n, we set

Eα(t(n)) =∑

σ∈Gn

mi1α(tσ(1))mi2α

(tσ(2)) · · ·minα(tσ(n)) (2.7)

where Gn is the permutation group on 1, 2, · · · , n. Notice that Eα, |α| = n is a complete

orthogonal basis of Hn and ‖Eα‖2Hn = n!α!. Then we are ready to establish multiple integrals

on Hn. For f =∑

|α|=n fαEα(t(n)) ∈ Hn, let

In(f) := n!∑

|α|=n

fαΦα ∈ D(R)

In(f) is square integrable due to the fact that

E[In(f)2] =∑

|α|=n

α!(n!fα)2 = n!‖f‖2Hn <∞

Now for u ∈ D(R), we can write down its chaos expansion with regard to multiple integrals:

u =∞∑

n=0

In(fn(t(n))), fn(t(n)) :=1

n!

|α|=n

uαEα(t(n)) (2.8)

By the same token, for u = u(t) ∈ D(H),

u(t) =

∞∑

n=0

In(fn(t, t(n))), fn(t, t(n)) :=1

n!

|α|=n

uα(t)Eα(t(n)) (2.9)

Definition 2.1. u = u(t) ∈ D(H) with expansion (2.9) is called adapted if

supp fn(t, ·) ∈ [0, t]n, for each t ∈ [0, T ]

The proposition below points out the close connection between Skorokhod integral and

the classic Ito integral, which is crucial to deriving deterministic propagators for Ito type

SPDEs. Proofs can be found in [7].

Proposition 2.3. For u = u(t) ∈ D(H) with expansion (2.9), if δ(u) ∈ D(R), the multiple

integral expansion of δ(u) is

δ(u) =∞∑

n=0

In+1(fn(t(n+1))) (2.10)

8

Page 9: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

where the standard symmetrization g for g ∈ Hn is taken to be

g(t(n)) :=1

n!

σ∈Gn

g(tσ(1), tσ(2), · · · , tσ(n)) (2.11)

Furthermore, in the case that ξk∞k=1 are i.i.d. standard Gaussian variables, if u is also

adapted, the Skorokhod integral coincides with the Ito integral:

δ(u) =

∫ T

0

u(t) ⋄ W (t)dt =

∫ T

0

u(t)dW (t) (2.12)

2.3 SPDE model problems

In this subsection we will look into examples of linear and nonlinear distribution-free SPDEs.

We will take linear parabolic SPDE as the linear model problem, and stochastic Burgers

equation as the nonlinear model problem. In both examples, the solution u = u(t, x) lives in

D(L2([0, T ]×D)), where D ∈ Rd is a smooth finite domain equipped with periodic boundary

condition. The polynomial chaos expansion of u(t, x) reads

u(t, x) =∑

α∈J

uα(t, x)Φα, uα ∈ L2([0, T ]2 ×D)

Example 2.3.1. Consider the following homogeneous linear parabolic SPDE:

∂tu(t, x) = Lu(t, x) +Mu(t, x) ⋄ N(t), (t, x) ∈ (0, T ]×D

u(0, x) = u0(x), x ∈ D(2.13)

where

Lu(t, x) =

d∑

i=1

d∑

j=1

aij(x)∂i∂ju(t, x) +

d∑

i=1

bi(x)∂iu(t, x) + c(x)u(t, x) (2.14)

Mu(t, x) =d∑

i=1

αi(x)∂iu(t, x) + β(x)u(t, x) (2.15)

and ∂i is the i-th spatial partial derivative. If ξk∞k=1 are i.i.d. standard Gaussian variables,

according to Proposition 2.3, (2.13) is equivalent to the Ito type SPDE

du(t, x) = Lu(t, x)dt +Mu(t, x)dW (t), (t, x) ∈ (0, T ]×D (2.16)

or the Stratonovich type SPDE

du(t, x) = Lu(t, x)dt +Mu(t, x) dW (t), (t, x) ∈ (0, T ]×D (2.17)

where Lu = Lu − 12MMu. We assume that the coefficients in L and M are smooth and

bounded, L is uniformly elliptic, and u0(x) is deterministic and bounded. These assumptions

9

Page 10: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

are sufficient for a unique square integrable solution u ∈ D(L2([0, T ]×D)) (see e.g., [20, 15]).

As we will see later, the propagator system is independent of the type of noise. Therefore

these well-posedness requirements remain the same in the distribution-free setting.

Recall the definition of the Skorokhod integral (2.5). We come up with the propagator

system by comparing the expansion coefficients on both sides of (2.13):

∂tuα(t, x) = Luα(t, x) +∑

εk≤α

Muα−εk(t, x)mk(t), (t, x) ∈ (0, T ]×D

uα(0, x) = u0(x)1α=ε0, x ∈ D

(2.18)

It is a system of linear parabolic deterministic PDEs, with a lower-triangular and sparse

structure, i.e. a multi-index with order n only talks to itself and multi-indices with order

n − 1. As a result, we can solve the system sequentially, and coefficients with the same

order can be updated in parallel. Additionally, the system does not depend on the type

of randomness involved, which implies the computational overhead from changes of noise is

almost negligible. The propagator is solved once and for all.

Numerical discretization of (2.18) usually follows the method of lines technique, in which

we start with standard spatial discretization schemes, transforming the propagator system

into a larger system of ODEs. Suppose that the spatial discretization has M degrees of

freedom, and A and B are M ×M difference matrices of L andM. The ODE system is

u′α(t) = Auα(t) +

εk≤α

mk(t)Buα−εk(t), t ∈ (0, T ]

uα(0) = u01α=ε0

(2.19)

where uα(t) ∈ RM is the vector uα(x, t) evaluated at those degrees of freedom. Then Runge-

Kutta type ODE solvers can be directly adopted.

In practice, it is impossible to handle the infinite system (2.18) or (2.19), and a finite

truncation is always necessary. For K, N ≥ 0, define the truncated multi-index set

JN,K := α ∈ J : |α| ≤ N, d(α) ≤ K, #(JN,K) =

(N + K

N

)

That is, JN,K contains multi-indices whose polynomial order is no more than N , and number

of random variables is no more than K. The size of JN,K grows rapidly with respect to both

N and K. The truncated solutions of (2.18) and (2.19) are

uN,K(t, x) :=∑

α∈JN,K

uα(t, x)Φα, uN,K(t) :=∑

α∈JN,K

uα(t)Φα (2.20)

Approximation error introduced by finite truncation will be further analyzed in Section 3.

10

Page 11: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Example 2.3.2. Here we restrict our interest to the one-dimensional case such that D ⊂ R.

Consider the stochastic Burgers equation with additive noise [9]:

∂tu(t, x) + ∂x

(u2

2

)= µ∂2

xu + σ(x)N(t), (t, x) ∈ (0, T ]×D

u(0, x) = u0(x), x ∈ D(2.21)

where µ is a positive constant and σ(x) is a periodic forcing function. The corresponding

Ito type SPDE in the case of i.i.d standard Gaussian noise is

du(t, x) =(µ∂2

xu− ∂x

(u2

2

))dt + σ(x)dW (t), (t, x) ∈ (0, T ]×D

u(0, x) = u0(x), x ∈ D(2.22)

By assuming that u0(x) is deterministic and σ, u0 ∈ L2(D), we make sure that (2.21) has a

unique square integrable solution (see e.g., [6]). However, such result does not generalize to

the distribution-free setting as the propagator system varies for different driving noises.

In order to figure out the propagator system, we have to expand u2 as stochastic poly-

nomial chaos expansion:

u2 =∑

α∈J

( ∑

β∈J

p∈J

B(α, β, p)uβup

)Φα (2.23)

where

B(α, β, p) =E[ΦαΦβΦp]

E[Φ2α]

=E[ΦαΦβΦp]

α!(2.24)

are interaction coefficients. Hence the propagator equations are

∂tuα(t, x) +1

2

β∈J

p∈J

B(α, β, p)∂x(uβup) = µ∂2xuα + σ(x)

∞∑

k=1

1α=εkmk(t), (t, x) ∈ (0, T ]×D

uα(0, x) = u0(x)1α=ε0, x ∈ D(2.25)

It is a fully coupled system of nonlinear PDEs, whose interaction coefficients B(α, β, p)

depend on the type of driving noise. Compared with the linear case, (2.25) lacks sparsity,

and must be recalculated each time we change distribution. Both features make the nonlinear

problem more challenging and expensive to simulate. In Appendix A we will present the

generic procedure to calculate interaction coefficients, and explicit formulas for some special

types of distribution.

For a truncated multi-index set JN,K , the approximated solution is

uN,K(t, x) :=∑

α∈JN,K

uα(t, x)Φα

11

Page 12: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

where uα : α ∈ JN,K satisfies the truncated propagator system

∂tuα(t, x)+1

2

β∈JN,K

p∈JN,K

B(α, β, p)∂x(uβup) = µ∂2xuα+σ(x)

∞∑

k=1

1α=εkmk(t), (t, x) ∈ (0, T ]×D

(2.26)

We note that uα is not uα due to aliasing error. The method of lines technique can also be

applied to the numerical discretization of (2.26).

Remark 2.2. In principle, the propagator system can be determined explicitly as long as

we only have polynomial nonlinearity. We expand power functions as tensor products in

a way that is similar to (2.23). The expansion of nonpolynomial functions is much more

challenging. Several methods are presented in [5] to perform general function evaluations on

polynomial chaos series.

3 Error estimate

For the sake of simplicity, we will focus on the truncation error analysis of the ODE system

(2.19), which can be viewed as the propagator system of the following linear SODE system:

u′(t) = Au(t) + Bu(t) ⋄ N(t), t ∈ (0, T ]

u(0) = u0

(3.1)

where u ∈ D(H) and H = (L2([0, T ]))M . A and B are constant M ×M matrices. The

reason for such simplification is twofold. Since (3.1) is the spatial discretization of (2.13), it

is the equation we are actually dealing with in numerical simulations. Besides, all arguments

in this section can be generalized to (2.13) in a straightforward manner, but with more

technical considerations. We simply replace the Euclidean norm with appropriate Sobolev

norms and impose regularity assumptions on L andM (see e.g., [36]).

Theorem 3.1. Suppose that mk∞k=1 is the trigonometric basis

m1(t) =

√1

T, mk(t) =

√2

Tcos

((k − 1)πt

T

), k ≥ 2 (3.2)

Let λA = ‖A‖2 and λB = ‖B‖2 be the matrix norms. Then the mean square error of the

truncated solution uN,K(T ) from (2.20) is bounded by the estimate

E[|uN,K(T )−u(T )|2] ≤ e(2λA+λ2B)T

((λ2BT )N+1

(N + 1)!+

16λ2Aλ2

BT 3

π4(K − 12)3

(5+3λ2AT 2+6λ4

BT 2))|u0|2 (3.3)

12

Page 13: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Remark 3.1. From (3.3), we conclude that the mean square truncation error converges at

an exponential rate with respect to N , and at a cubic rate with respect to K. We improve

the error estimate in [12] where the authors only proved linear rate with respect to K. Cu-

bic convergence result can also be found in [9] for a special example of stochastic Burgers

equation. However, the approximation error increases exponentially in time. Long time sim-

ulation might be impractical. At least more expansion coefficients are required to balance out

error growth.

The proof is primarily along the lines in [12]. We first prove a lemma to extract the

analytical solution of the propagator system.

Lemma 3.1. Suppose uα(t), α ∈ J solves the propagator system (2.19). For each n ≥ 0

and α ∈ J with |α| = n, the explicit formula of uα(t) is

uα(t) =1

α!

∫ (t,n)

Fn(t, t(n))Eα(t(n))dt(n) (3.4)

where Eα(t(n)) is from (2.7) and

Fn(t, t(n)) := e(t−tn)ABe(tn−tn−1)AB · · ·Bet1Au0

∫ (t,n)

g(t(n))dt(n) :=

∫ t

0

∫ tn

0

· · ·∫ t2

0

g(t(n))dt1 · · · dtn−1dtn

Notice that Fn is not related to the fn in (2.9) as Fn is not symmetric.

Proof. We prove by induction on n. If n = 0, uε0(t) = etAu0. (3.4) is obviously correct. Now

for n ≥ 1 and |α| = n, we assume that (3.4) holds for all β ∈ J with |β| < n. By Duhamel’s

principle,

uα(t) =

∫ t

0

e(t−s)A( ∑

εk≤α

mk(s)Buα−εk(s)

)ds

=1

α!

∫ t

0

e(t−s)A( ∑

εk≤α

αkmk(s)B

∫ (s,n−1)

Fn−1(s, t(n−1))Eα−εk

(t(n−1))dt(n−1))ds

=1

α!

∫ (t,n)

Fn(t, t(n))( ∑

εk≤α

αkmk(tn)Eα−εk(t(n−1))

)dt(n)

=1

α!

∫ (t,n)

Fn(t, t(n))Eα(t(n))dt(n)

where we use the identity

Eα(t(n)) =∑

εk≤α

αkmk(tn)Eα−εk(t(n−1))

Hence (3.4) is satisfied by any α.

13

Page 14: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Proof of Theorem 3.1. According to Parseval’s identity, we decompose the truncation error

as

E[|uN,K(T )− u(T )|2] =∞∑

n=N+1

|α|=n

α!|uα(T )|2 +∞∑

k=K+1

N∑

n=1

|α|=n,d(α)=k

α!|uα(T )|2

We only need to show the two inequalities below:

∞∑

n=N+1

|α|=n

α!|uα(T )|2 ≤ e(2λA+λ2B

)T (λ2BT )N+1

(N + 1)!|u0|2 (3.5)

∞∑

k=K+1

N∑

n=1

|α|=n,d(α)=k

α!|uα(T )|2 ≤ e(2λA+λ2B)T 16λ2

Aλ2BT 3

π4(K − 12)3

(5 + 3λ2AT 2 + 6λ4

BT 2)|u0|2 (3.6)

As for (3.5), set Fn(T, ·) to be the standard symmetrization of Fn(T, ·) (Fn is extended

with zero value outside the simplex t(n) : 0 ≤ t1 ≤ · · · ≤ tn ≤ T), we have

uα(T ) =1

α!

∫ (T,n)

Fn(T, t(n))Eα(t(n))dt(n) =1

α!

[0,T ]nFn(T, t(n))Eα(t(n))dt(n)

Since Eα, |α| = n is an orthogonal basis of Hn,

|α|=n

α!|uα(T )|2 =∑

|α|=n

1

α!

∣∣∣∫

[0,T ]nFn(T, t(n))Eα(t(n))dt(n)

∣∣∣2

= n!‖Fn(T, ·)‖2Hn = (n!)2

∫ (T,n)

|Fn(T, t(n))|2dt(n)

=

∫ (T,n)

|Fn(T, t(n))|2dt(n)

(3.7)

For any given t(n),

|Fn(T, t(n))| ≤ eλA(T−tn)λBeλA(tn−tn−1) · · ·λBeλAt1 |u0| = eλAT λnB|u0|

Plugging this into (3.7) yields

∞∑

n=N+1

|α|=n

α!|uα(T )|2 ≤( ∞∑

n=N+1

e2λAT (λ2BT )n

n!

)|u0|2 ≤ e(2λA+λ2

B)T (λ2

BT )N+1

(N + 1)!|u0|2

which exactly recovers (3.5). Here we use the mean-value form of the remainder term of

Taylor’s expansion:∞∑

n=N+1

xn

n!= eθx xN+1

(N + 1)!for some θ ∈ [0, 1]

14

Page 15: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Proof of (3.6) is more involved. For any α with |α| = n and d(α) = k,

∫ (T,n)

Fn(T, t(n))Eα(t(n))dt(n) =

n∑

j=1

∫ (T,n−1) (∫ tj+1

tj−1

Fn(T, t(n))mk(tj)dtj

)Eα−εk

(t(n\j))dt(n\j)

=n∑

j=1

∫ (T,n−1) (∫ tj

tj−1

Fn(T, t(n\j,s))mk(s)ds)Eα−εk

(t(n−1))dt(n−1)

(3.8)

where t(n\j) is the short hand notation of (t1, · · · , tj−1, tj+1, · · · , tn) and t(n\j,s) is the short

hand notation of (t1, · · · , tj−1, s, tj, · · · , tn−1). We also adopt the convention t0 = 0, tn+1 = T .

Define

M1k (t) :=

∫ t

0

mk(s)ds =

√2T

(k − 1)πsin

((k − 1)πt

T

)

M2k (t) :=

∫ t

0

M1k (s)ds =

√2T 3

(k − 1)2π2

(1− cos

((k − 1)πt

T

))

and

Fjn(T, t(n)) :=

∂Fn

∂tj(T, t(n)) = e(T−tn)AB · · · e(tj+1−tj)A(BA−AB)e(tj−tj−1)A · · ·Bet1Au0

Fjjn (T, t(n)) :=

∂2Fn

∂t2j(T, t(n)) = e(T−tn)AB · · · e(tj+1−tj)A(A2B+BA2−2ABA)e(tj−tj−1)A · · ·Bet1Au0

The following estimates are right at hand:

|Fjn(T, t(n\j,s))| ≤ 2eλAT λAλn

B|u0|, |Fjjn (T, t(n\j,s))| ≤ 4eλAT λ2

AλnB|u0|, ∀1 ≤ j ≤ n and t(n\j,s)

(3.9)

M1k (0) = M1

k (T ) = M2k (0) = 0 , |M2

k (T )| ≤√

8T 3

(k − 1)2π2,

∫ T

0

(M2k (t))2dt =

3T 4

(k − 1)4π4

(3.10)

Then we perform integration-by-parts twice on the inner integral of (3.8) and obtain

∫ (T,n)

Fn(T, t(n))Eα(t(n))dt(n) :=

∫ (T,n−1)

Gn,k(T, t(n−1))Eα−εk(t(n−1))dt(n−1)

:=

∫ (T,n−1)

(G1n,k + G2

n,k + G3n,k)(T, t(n−1))Eα−εk

(t(n−1))dt(n−1)

where

Gn,k(T, t(n−1)) = G1n,k(T, t(n−1)) + G2

n,k(T, t(n−1)) + G3n,k(T, t(n−1))

and

G1n,k(T, t(n−1)) =

n∑

j=1

(Fn(T, t(n\j,s))M1

k (s)∣∣∣s=tj

s=tj−1

)

15

Page 16: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

G2n,k(T, t(n−1)) = −

n∑

j=1

(Fj

n(T, t(n\j,s))M2k (s)

∣∣∣s=tj

s=tj−1

)

G3n,k(T, t(n−1)) =

n∑

j=1

( ∫ tj

tj−1

Fjjn (T, t(n\j,s))M2

k (s)ds)

Since M1k (0) = M1

k (T ) = 0 and Fn(T, t(n\j,s); s = tj) = Fn(T, t(n\(j+1),s); s = tj) for 1 ≤ j ≤n− 1, G1

n,k(T, t(n−1)) = 0. By (3.9) and (3.10), the other two terms are bounded by

|G2n,k(T, t(n−1))| ≤ 2eλAT λAλn

B|u0|(2

n−1∑

j=1

|M2k (tj)|+ |M2

k (T )|)

≤ 4eλAT λAλnB|u0|

( n−1∑

j=1

|M2k (tj)|+

√2T 3

(k − 1)2π2

)

|G3n,k(T, t(n−1))| ≤ 4eλAT λ2

AλnB|u0|

(∫ T

0

|M2k (t)|dt

)≤ 4eλAT λ2

AλnB|u0|

T

∫ T

0

(M2k (t))2dt

= 4eλAT λ2Aλn

B|u0|√

3T 5

(k − 1)2π2

Similar to the idea in (3.7),

|α|=n,d(α)=k

α!|uα(T )|2 =∑

|α|=n,d(α)=k

1

α!

∣∣∣∫ (T,n)

Fn(T, t(n))Eα(t(n))dt(n)∣∣∣2

=∑

|α|=n,d(α)=k

1

α!

∣∣∣∫ (T,n−1)

Gn,k(T, t(n−1))Eα−εk(t(n−1))dt(n−1)

∣∣∣2

≤∑

|β|=n−1

1

β!

∣∣∣∫ (T,n−1)

Gn,k(T, t(n−1))Eβ(t(n−1))dt(n−1)∣∣∣2

=

∫ (T,n−1)

|Gn,k(T, t(n−1))|2dt(n−1) ≤∫ (T,n−1) (

|G2n,k(T, t(n−1))|+ |G3

n,k(T, t(n−1))|)2

dt(n−1)

≤16e2λAT λ2Aλ2n

B

∫ (T,n−1) ( n−1∑

j=1

|M2k (tj)|+

√2T 3

(k − 1)2π2+ λA

√3T 5

(k − 1)2π2

)2

dt(n−1)

≤48e2λAT λ2Aλ2n

B

∫ (T,n−1) ((n− 1)

( n−1∑

j=1

(M2k (tj))

2)

+2T 3

(k − 1)4π4+ λ2

A

3T 5

(k − 1)4π4

)dt(n−1)

The remaining part of proof is clear. Since∑n−1

j=1 (M2k (tj))

2 is a symmetric function,

∫ (T,n−1) ( n−1∑

j=1

(M2k (tj))

2)dt(n−1) =

1

(n− 1)!

[0,T ]n−1

( n−1∑

j=1

(M2k (tj))

2)dt(n−1)

=n− 1

(n− 1)!

3T n+2

(k − 1)4π4

16

Page 17: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Therefore

|α|=n,d(α)=k

α!|uα(T )|2 ≤ 48e2λAT λ2Aλ2n

B

(n− 1)!(k − 1)4π4

(3(n− 1)2T n+2 + 2T n+2 + 3λ2

AT n+4)

Summing over n and k yields:

∞∑

k=K+1

N∑

n=1

|α|=n,d(α)=k

α!|uα(T )|2

≤e2λAT 48λ2A

π4

( ∞∑

k=K+1

1

(k − 1)4

)( N∑

n=1

λ2nB (3(n− 1)2T n+2 + 2T n+2 + 3λ2

AT n+4)

(n− 1)!

)|u0|2

≤e(2λA+λ2B

)T 16λ2Aλ2

BT 3

π4(K − 12)3

(5 + 3λ2AT 2 + 6λ4

BT 2)|u0|2

where we use the inequalities

∞∑

k=K+1

1

(k − 1)4≤

∞∑

k=K

∫ k+ 1

2

k− 1

2

1

x4dx =

∫ ∞

K− 1

2

1

x4dx =

1

3(K − 12)3

N∑

n=1

xn−1

(n− 1)!≤

∞∑

n=0

xn

n!= ex

N∑

n=1

(n− 1)2xn−1

(n− 1)!= x +

N−3∑

n=0

(n + 2)xn+2

(n + 1)!≤ x +

∞∑

n=0

2xn+2

n!≤ ex(1 + 2x2)

We have finished the proofs of (3.5) and (3.6). Then (3.3) immediately follows.

Remark 3.2. The proof of (3.5) is independent of the choice of mk∞k=1, the convergence

rate with respect to N is always exponential. The proof of (3.6) relies on the trigonometric

basis assumption. The crucial property is (3.10), which enables cubic convergence. In fact,

the proof will work for any orthonormal basis such that

M1k (T ) = 0, M2

k (T ) = O(k−2), ‖M2k‖H = O(k−2), ∀k ≥ 2

For example, consider the scaled Legendre basis

mk(t) =

√2k − 1

TLk−1

(2t

T− 1

)(3.11)

We are able to show that

M1k (t) =

1

2

√T

2k − 1

(Lk

(2t

T− 1

)− Lk−2

(2t

T− 1

))

17

Page 18: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

M2k (t) =

1

4(2k + 1)

√T 3

2k − 1

(Lk+1

(2t

T− 1

)− Lk−1

(2t

T− 1

))

− 1

4(2k − 3)

√T 3

2k − 1

(Lk−1

(2t

T− 1

)− Lk−3

(2t

T− 1

))

where Lk is taken to be 0 for negative k. Then M1k (T ) = 0 for any k ≥ 2 and M2

k (T ) = 0 for

any k ≥ 3. We also have ‖M2k‖H = O(k−2). Therefore for Legendre basis, the convergence

rate with respect to K is still cubic.

Remark 3.3. If A and B commute such that AB = BA, then

Fn(T, t(n)) = e(T−tn)Ae(tn−tn−1)A · · · et1ABnu0 = eTABnu0

is a constant vector that does not depend on t(n). Consequently, for any |α| = n,

uα(T ) =eTABnu0

α!

∫ (T,n)

Eα(t(n))dt(n) =eTABnu0

n!α!

[0,T ]nEα(t(n))dt(n)

=eTABnu0

α!

[0,T ]nmi1α

(t1) · · ·minα(tn)dt(n) =

eTABnu0

α!M1

i1α(T ) · · ·M1

inα(T )

For trigonometric basis (and Legendre basis), M1k (T ) = 0 for any k ≥ 2. That is, uα(T ) = 0

whenever d(α) = inα ≥ 2. It is enough to fix K = 1 and only consider the truncation on N .

The resulting error estimate is simply

E[|uN,1(T )− u(T )|2] ≤ e(2λA+λ2B

)T (λ2BT )N+1

(N + 1)!|u0|2 (3.12)

4 Numerical experiments

We first introduce some post-processing techniques for the numerical solution written as

polynomial chaos expansion:

uN,K(t, x) =∑

α∈JN,K

uα(t, x)Φα

Moments can be computed directly, the first and second moments are

E[uN,K(t, x)] = uε0(t, x), E[u2

N,K(t, x)] =∑

α∈JN,K

α!(uα(t, x))2 (4.1)

For linear problems, the first two moments remain the same for all kinds of randomness since

the propagator system does not change. The third and fourth moments are given by

E[u3N,K(t, x)] =

α∈JN,K

α!uα(t, x)( ∑

β∈JN,K

p∈JN,K

B(α, β, p)uβ(t, x)up(t, x))

(4.2)

18

Page 19: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

E[u4N,K(t, x)] =

α∈J2N,K

α!(( ∑

β∈JN,K

p∈JN,K

B(α, β, p)uβ(t, x)up(t, x))2

(4.3)

In the computation of fourth moment, we use the fact that the expansion order of u2N,K(t, x)

is at most 2N . Due to the emergence of interaction coefficients B(α, β, p), higher moments

always depend on the type of randomness. Other statistics can be computed via random

sampling. We simply generate L i.i.d. copies of (ξ1, · · · , ξK), denoted by ξ(l)1 , · · · , ξ(l)

K for

each 1 ≤ l ≤ L. The sample points of uN,K(t, x) are:

u(l)N,K(t, x) :=

J∈JN,K

uα(t, x)Φα(ξ(l)1 , · · · , ξ(l)

K )

Then for any function f , the expectation E[f(uN,K(t, x))] is approximated by sample mean.

We can also plot the normalized histogram of these sample points to visualize empirical

distribution.

In this section, we always assume that ξk∞k=1 are i.i.d. random variables. To be more

specific, we will test three types of randomness: Gaussian noise with Hermite chaos (Ex-

ample 2.1.1), uniform noise with Legendre chaos (Example 2.1.2), and Beta(12, 1

2) noise with

Chebyshev chaos. In the last situation, ξk∞k=1 are supported on [−√

2,√

2] with probability

density function

ρ(ξ1) =

√2− ξ2

1

2π, ξ1 ∈ [−

√2,√

2]

The univariate chaos basis functions are rescaled Chebyshev polynomials Tn(x)∞n=0:

ϕ0(ξ1) = 1, ϕn(ξ1) =√

2n!Tn

( ξ1√2

), n ≥ 1 (4.4)

Now we proceed to solve distribution-free linear and nonlinear SPDEs numerically. Prop-

agator systems are integrated in time with fourth order Runge-Kutta method. Time step is

small enough so that error from temporal discretization is negligible. mk(t)∞k=1 is taken to

be the trigonometric basis (3.2). The results of the scaled Legendre basis (3.11) are almost

indistinguishable from trigonometric basis and will not be reported.

For the purpose of validation and comparison, we also adopt several techniques to com-

pute reference solutions.

• Moment equations: the ODE of first few moments, available to linear Ito type SPDEs.

• Fokker-Planck equation: the PDE of probability density function, available to low-

dimensional Ito type SODEs.

• Monte Carlo simulation: the most commonly used and least restrictive approach, avail-

able to additive noise and/or Ito type SPDEs. Main drawbacks are high computational

cost and low accuracy.

19

Page 20: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

The first two methods are free from sampling error, and will be selected whenever possible.

Monte Carlo simulation serves as a backup option.

Example 4.3 (Linear SODE). Suppose that u = u(t) ∈ D(H) satisfies the following linear

SODEu′(t) = u(t) + 1 + u(t) ⋄ N(t), t ∈ [0, T ]

u(0) = 1(4.5)

For Gaussian noise, it is equivalent to the Ito type SODE

du(t) = (u(t) + 1)dt + u(t)dW (t), t ∈ [0, T ] (4.6)

By Ito’s formula, we are able to derive its moment equations

dE[u(t)]

dt= E[u(t)] + 1

dE[u2(t)]

dt= 3E[u2(t)] + 2E[u(t)]

dE[u3(t)]

dt= 6E[u3(t)] + 3E[u2(t)]

dE[u4(t)]

dt= 10E[u4(t)] + 6E[u3(t)]

(4.7)

The analytical solution to (4.7) is

E[u(t)] = 2et − 1

E[u2(t)] =7

3e3t − 2et +

2

3

E[u3(t)] =37

15e6t − 7

3e3t +

6

5et − 1

3

E[u4(t)] =38

15e10t − 37

15e6t +

4

3e3t − 8

15et +

2

15

(4.8)

For uniform and Beta noise, the first and second moments are the same, but we do not have

explicit formulas for the third and fourth moments.

We evolve the propagator ODE system up to end time T = 1 with time step δt = 10−4.

In order to examine the convergence rates of mean square truncation error, ideally we should

compute E[|uN,K − u∞,K|2] to single out the error induced by N , and E[|uN,K − uN,∞|2] to

single out the error induced by K. In practice we use u20,K to approximate u∞,K, and uN,50

to approximate uN,∞. Figure 4.1 contains the semi-log plot of E[|uN,K(1)−u20,K(1)|2] versus

N with K = 1, and the log-log plot of E[|uN,K(1) − uN,50(1)|2] versus K with N = 1, 2, 3.

The numerical rate of convergence with respect to N is evidently exponential. The plot with

respect to K has a zigzag shape (especially for N = 1), but the average slope is close to 3.

20

Page 21: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

1 2 3 4 5 6 7 8 9 10

N

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

101

[|u

N,1(1)−u20,1(1)|2

]

(a) Convergence with respect to N

1 2 3 4 5 6 7 8 9 10 11 12

K

10−5

10−4

10−3

10−2

10−1

100

[|u

N,K

(1)−uN,5

0(1)|2

]

1 : 3

N=1

N=2

N=3

(b) Convergence with respect to K

Figure 4.1: Example 4.3: Plots of mean square truncation error with respect to N and K.Left panel shows the semi-log plot of E[|uN,1(1) − u20,1(1)|2] versus N for N = 1, · · · , 10.Right panel shows the log-log plot of E[|uN,K(1) − uN,50(1)|2] versus K for N = 1, 2, 3 andK = 1, · · · , 12.

Table 4.1: Example 4.3: Values and numerical convergence orders of eN,K :=E[|uN,K(1) − uN,50(1)|2] for N = 1, 2, 3 and K = 2, 4, · · · , 12. The orders are given bylog(

eN,K

eN,K+2)/ log(K+2

K).

KN = 1 N = 2 N = 3

error order error order error order2 7.997e-3 - 2.080e-2 - 2.677e-2 -4 9.659e-4 3.049 2.201e-3 3.241 2.831e-3 3.2416 2.816e-4 3.040 6.102e-4 3.164 7.799e-4 3.1808 1.173e-4 3.044 2.480e-4 3.129 3.158e-4 3.14310 5.939e-5 3.050 1.238e-4 3.114 1.572e-4 3.12612 3.398e-5 3.062 7.018e-5 3.113 8.895e-5 3.123

The cubic convergence rate is more clearly seen in Table 4.1, where we only compare even

values of K to average out the zig-zag profile.

Next we compute moments of the truncated solution. Table 4.2 lists the first four central

moments of uN,K(1) with all three types of randomness, by taking N = K = 4, N = K = 6

and N = K = 8. For Gaussian noise, moments of u(1) are also included according to

(4.8), while for uniform noise and Beta(12, 1

2) noise, third and fourth moments of u(1) are not

available. Two conclusions can be drawn from the table. Comparing the central moments

of uN,K(1) and u(1), we see that the variance can be approximated well with relatively few

chaos expansion terms, but more terms are needed to resolve higher moments. Comparing

among types of driving noise, we notice the large discrepancy in third and fourth moments,

21

Page 22: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

despite the fact that they share the same mean and variance. It is probably related to

the kurtosis of different distributions. The fourth moment of ξ1 is 3 for standard Gaussian

distribution, 95

for uniform distribution, and 32

for Beta(12, 1

2) distribution. Higher kurtosis

in ξk∞k=1 leads to higher kurtosis in uN,K .

Table 4.2: Example 4.3: Comparison of central moments of uN,K(1) and u(1). We takeN = K = 4, N = K = 6 and N = K = 8. Higher moments of u(1) are only available toGaussian noise.

Type of noise Type of moment u(1) u4,4(1) u6,6(1) u8,8(1)Gaussian Variance 22.413 22.313 22.410 22.413

Third central moment 565.548 487.838 558.223 565.138Fourth central moment 41759.97 22914.36 37479.48 41233.50

Uniform Variance 22.413 22.313 22.410 22.413Third central moment - 208.415 220.604 221.203Fourth central moment - 3080.52 3446.20 3474.01

Beta Variance 22.413 22.313 22.410 22.413Third central moment - 150.739 159.566 160.011Fourth central moment - 1789.49 1957.34 1970.63

Empirical distribution is a more intuitive way to describe random variables. Figure 4.2

demonstrates the empirical probability densities of u4,4(1), u6,6(1) and u8,8(1) for all three

types of randomness. All densities are estimated by normalized histograms with 103 bins

out of 107 i.i.d samples of uN,K(1). For Gaussian noise, we can also find out the probability

density function of u(t), denoted by ρ(u, t). The governing PDE of ρ(t, u), i.e. Fokker-Planck

equation, follows from the Ito form (4.6):

∂tρ(t, u) = −∂u((u + 1)ρ) + ∂2u

(u2

2ρ), (t, u) ∈ (0, 1]× (0,∞)

ρ(0, u) = δ(u− 1), u ∈ (0,∞)(4.9)

where δ is the Dirac delta function. Substituting v = log u, we simplify (4.9) and get

∂tρ(t, v) =(1

2− e−v

)∂vρ +

1

2∂2

vρ, (t, v) ∈ (0, 1]× R

ρ(0, v) = δ(v), v ∈ R

(4.10)

(4.10) is a standard convection-diffusion equation. We choose the local discontinuous Galerkin

(LDG) method [4] as the numerical solver. The computational domain is [−7, 7] with zero

boundary condition, and divided into 501 quadratic elements. The numerical solution of

Fokker-Planck equation at t = 1 is also displayed in Figure 4.2. The empirical densities of

u4,4(1) and u6,6(1) slightly deviates from the Fokker-Planck solution, and the empirical den-

sity of u8,8(1) agrees with the Fokker-Planck solution very well. As for the comparison among

22

Page 23: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

three types of noise, their density patterns are qualitatively different. The distributions with

Gaussian noise spread out and have long tails. The density profiles with Beta(12, 1

2) noise

are mostly constrained in a narrow region, and the density profiles with uniform noise lie

somewhere in between. This also explains the difference of higher moments in Table 4.2.

0 5 10 15 20

u

0

0. 05

0. 1

0. 15

0. 2

0. 25

Density

N=4, K=4

N=6, K=6

N=8, K=8

Fokker-Planck

(a) Gaussian noise

0 5 10 15 20

u

0

0. 1

0. 2

0. 3

0. 4

0. 5

Density

N=4, K=4

N=6, K=6

N=8, K=8

(b) Uniform noise

0 5 10 15 20

u

0

0. 2

0. 4

0. 6

0. 8

1

1. 2

Density

N=4, K=4

N=6, K=6

N=8, K=8

(c) Beta(1

2, 1

2) noise

Figure 4.2: Example 4.3: Normalized histograms of uK,N(1) out of 107 i.i.d samples. Wetake N = K = 4, N = K = 6 and N = K = 8. For Gaussian distribution, the black dashedline represents the numerical solution of Fokker-Planck equation (4.10). Values larger than20 are discarded. Number of bins is 103.

In summary, this example studies a very simple linear SODE so that reference solutions

of moments and density function can be acquired without much effort. By linearity, we only

need to solve the propagator once and save the expansion coefficients for post-processing.

The mean square truncation error results in Figure 4.1 and Table 4.1 indicate exponential

convergence with respect to N and cubic convergence with respect to K, as predicted by

23

Page 24: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

Theorem 3.1. Table 4.2 and Figure 4.2 highlight the contrast of higher moments and empir-

ical distributions with different noises. We also observe that although the mean square error

converges rapidly, high order chaos expansion terms are beneficial to the approximation of

higher moments and distribution.

Example 4.4 (Linear parabolic PDE). We solve the linear parabolic PDE in Example 2.3.1.

Consider the one-dimensional space region D = [0, 2π] with periodic boundary condition.

The initial data is u0(x) = cos x. Differential operators L andM are set to be [35, 36]

Lu = 0.145∂2xu + 0.1 sin x∂xu, Mu = 0.5∂xu

Fourier collocation method with M = 32 collocation points is used for spatial discretization.

Let xiMi=1 be the set of equidistant collocation points such that xi = 2π(i−1)M

. Then u(t, ·) is

identified by the vector evaluated at xiMi=1:

u(t) =[u(1)(t) . . . u(M)(t)

]T:=

[u(t, x1) · · · u(t, xM)

]T

After spatial discretization, L andM turn into difference matrices A and B. We only need

to work on the linear SODE system (3.1).

For Gaussian noise, it is possible to obtain moment equations of (3.1). Direct application

of Ito’s formula yields

dE[u(i)u(j)]

dt=

M∑

l=1

(AilE[u(j)u(l)] + AjlE[u(i)u(l)]) +

M∑

l=1

M∑

r=1

BilBjrE[u(l)u(r)], 1 ≤ i, j ≤ M

(4.11)

It is a M2-dimensional ODE system describing the evolution of covariance matrix. The

covariance matrix is not influenced by the type of driving noise, as a result of linearity.

Higher moment equations are written in a similar fashion, but they are computationally

formidable in that for d-th moment, we have to tackle the full tensor product system with

Md dimensions.

The propagator system is run up to T = 5 with Runge-Kutta time step δt = 10−3. Figure

4.3 depicts the mean square truncation error with respect to N and K. Same as the previous

example, we compute E[‖uN,K(5, ·)− u20,K(5, ·)‖2l2] as the proxy of error induced by N , and

E[‖uN,K(5, ·)−uN,50(5, ·)‖2l2] as the proxy of error induced by K, where the discrete L2 norm

for v ∈ H is defined by

‖v‖2l2 :=2π

M

M∑

i=1

(v(i))2 =2π

M

M∑

i=1

(v(xi))2

Once again, the N -version convergence shows exponential rate, and the K-version conver-

gence shows cubic rate.

24

Page 25: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

1 2 3 4 5 6 7 8 9 10

N

10−5

10−4

10−3

10−2

10−1

100

[‖(

uN,1−u20

,1)(5,·)‖

2 l2]

(a) Convergence with respect to N

1 2 3 4 5 6 7 8 9 10 11 12

K

10−7

10−6

10−5

10−4

10−3

10−2

[‖(

uN,K

−uN,5

0)(5,·)‖

2 l2]

1 : 3

N=1

N=2

N=3

(b) Convergence with respect to K

Figure 4.3: Example 4.4: Plots of mean square truncation error in discrete l2 norm withrespect to N and K. Left panel shows the semi-log plot of E[‖uN,1(5, ·)−u20,1(5, ·)‖2l2] versusN for N = 1, · · · , 10. Right panel shows the log-log plot of E[‖uN,K(5, ·) − uN,50(5, ·)‖2l2]versus K for N = 1, 2, 3 and K = 1, · · · , 12.

The first four central moments of uN,K(5, ·) with all three noises are plotted in Figure

4.4. We consider N = 4, 8 and K = 4. Variance of u(5, ·) is also exhibited as reference by

solving the moment equation (4.11) through fourth order Runge-Kutta method with time

step δt = 10−3. For Gaussian noise, higher moments can be approximated by Monte Carlo

sampling. We use the second order weak scheme (see Chapter 2 of [24]) to discretize the

Ito integral and generate sample paths. Let up be the sample of u at p-th time step. The

update rule is

up+1 = up + δtAup +δt2

2A2up +

√δtζpBup +

δt

2(ζ2

p − 1)B2up +

√δt3

2ζp(AB + BA)up

where ζp are i.i.d. standard Gaussian random variables. Third and fourth central moments

of the Monte Carlo solution with 106 samples and time step δt = 10−3 are also shown in

Figure 4.4. We observe that both u4,4 and u8,4 predict the variance sufficiently well, but

only u8,4 succeeds in resolving higher moments. As for the comparison among three types

of noise, their third central moments have different structures. Fourth central moments look

similar in shape but different in magnitude.

Empirical distributions at the second collocation point x2 are illustrated in Figure 4.5.

For Gaussian noise, we plot normalized histograms of u4,4(5, x2) and u8,4(5, x2) out of 107

samples in the left panel, as well as the histogram of 106 Monte Carlo samples for reference.

Number of bins is 103. Here solving Fokker-Planck equation is not feasible as it depends on

M spatial variables. We underline that the distribution of u(5, x2) is supported in [−1, 1], as

25

Page 26: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

0 π/2 π 3π/2 2π

x

0

0. 1

0. 2

0. 3

0. 4

Variance

N=4, K=4

N=8, K=4

Moment equation

(a) Variance

0 π/2 π 3π/2 2π

x

−0. 2

−0. 1

0

0. 1

0. 2

Third central moment

Gaussian: N=4, K=4

Gaussian: N=8, K=4

Gaussian: Monte Carlo

Uniform: N=4, K=4

Uniform: N=8, K=4

Beta: N=4, K=4

Beta: N=8, K=4

(b) Third central moment

0 π/2 π 3π/2 2π

x

0

0. 1

0. 2

0. 3

0. 4

Fourth central moment

Gaussian: N=4, K=4

Gaussian: N=8, K=4

Gaussian: Monte Carlo

Uniform: N=4, K=4

Uniform: N=8, K=4

Beta: N=4, K=4

Beta: N=8, K=4

(c) Fourth central moment

Figure 4.4: Example 4.4: Central moments of uK,N(5, ·). We take N = 4, 8 and K = 4. Blackdashed lines are reference solutions. Variance is computed via moment equation (4.11) (forall types of noise), and higher moments are approximated by Monte Carlo method with 106

samples (only for Gaussian noise).

a result of averaging over characteristic lines [23]. The empirical distribution of Monte Carlo

sampling is indeed inside [−1, 1] and highly rightly skewed. Almost all samples of u4,4(5, x2)

and u8,4(5, x2) fall into [−1, 1] as well. We discard the few outlier samples in the figure.

The empirical density function of u4,4 underestimates the position of right peak, and the

empirical density function u8,4 is almost on top of the reference solution. For uniform noise

and Beta(12, 1

2) noise, we plot the empirical density functions of u10,4(5, x2) and u12,4(5, x2)

in the right panel. The profiles of u10,4 and u12,4 nearly coincide with each other, which

suggests that we achieve reasonable approximations of the true density functions. These

density patterns look dramatically different from the Gaussian noise case, in that they are

26

Page 27: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

neither supported in [−1, 1] nor rightly skewed. Such distinction is consistent with Figure

4.4, where the skewness (third central moment) at x2 is negative for Gaussian noise, and

positive for other two noises.

−1 −0. 5 0 0. 5 1

u

0

1

2

3

4

5

6

7

8

Density

N=4, K=4

N=8, K=4

Monte Carlo

(a) Gaussian noise

−0. 5 0 0. 5 1 1. 5

u

0

0. 5

1

1. 5

2

2. 5

Density

Uniform: N=10, K=4

Uniform: N=12, K=4

Beta: N=10, K=4

Beta: N=12, K=4

(b) Uniform and Beta(1

2, 1

2) noise

Figure 4.5: Example 4.4: Normalized histograms of uN,K(5, x2) out of 107 i.i.d samples. Inthe left panel, we take N = 4, 8 and K = 4. Values outside [−1, 1] are discarded, and theblack dashed line represents the normalized histogram of Monte Carlo simulation with 106

samples. In the right panel, we take N = 10, 12 and K = 4. Number of bins is 103.

In this example, we analyze a one-dimensional linear parabolic SPDE with relatively long

evolution time. Here solving moment equations is only practical for the second moment. We

resort to Monte Carlo simulation to produce other reference solutions. Our observations

are roughly parallel to the previous example. The truncation error of the second moment

converges at rates predicted by Theorem 3.1. Higher moments and empirical distributions

are more difficult to characterize, and highly depend on the type of underlying randomness.

Example 4.5 (Passive scalar equation). We move on to two-dimensional linear transport

type SPDE. Consider the following distribution-free passive scalar equation, driven by two

independent noises N1(t) and N2(t), and equipped with periodic boundary condition.

∂tu(t, x, y) =1

2(M2

1 +M22)u +M1u ⋄ N1(t) +M2u ⋄ N2(t), (t, x, y) ∈ [0, T ]× [0, 2π]2

u(0, x, y) = u0(x, y) = sin(2x) sin(y), (x, y) ∈ [0, 2π]2

(4.12)

where

M1u = cos(x + y)(∂xu− ∂yu), M2u = sin(x + y)(∂xu− ∂yu)

N1(t) =

∞∑

k=1

mk(t)ξ1k, N2(t) =

∞∑

k=1

mk(t)ξ2k

27

Page 28: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

and ξdk : d = 1, 2, k ≥ 1 is an i.i.d. sequence of random variables. We introduce the

Cartesian product of multi-index sets:

J 2 := α = (α1, α2) : α1, α2 ∈ J , J 2N,K := α = (α1, α2) : α1, α2 ∈ JN,K

The polynomial chaos basis functions are

Φα :=∞∏

k=1

ϕα1k(ξ1

k)∞∏

k=1

ϕα2k(ξ2

k)

Under the extended nomenclature, the propagator system has the form

∂tuα(t, x, y) =1

2(M2

1 +M22)uα +

ε1k≤α

M1uα−ε1kmk(t) +

ε2k≤α

M2uα−ε2kmk(t), (t, x, y) ∈ [0, T ]× [0, 2π]2

uα(0, x, y) = sin(2x) sin(y)1α=ε0, (x, y) ∈ [0, 2π]2

(4.13)

where ε1k = (εk, ε0) and ε2

k = (ε0, εk). The truncated solution is still defined as

uN,K(t, x, y) :=∑

α∈J 2N,K

uα(t, x, y)Φα (4.14)

We employ Fourier collocation method with M = 64 collocation points in each dimension for

the spatial discretization of (4.13). Equidistant collocation points are denoted by xi = yi =2π(i−1)

M. The propagator system is then computed up to T = 0.2 with time step δt = 5×10−4.

For Gaussian noise, (4.12) reduces to the passive scalar equation in Stratonovich version

[13, 34]

du(t, x, y) =M1u dW1(t) +M2u dW2(t), (t, x, y) ∈ [0, T ]× [0, 2π]2 (4.15)

We are able to work out the analytical solution based on tracing back characteristic lines

[23]. u(T, x, y) has the following representation

u(T, x, y) = u0(Xx,y(0), Yx,y(0)) (4.16)

where Xx,y(t) and Yx,y(t) satisfy the system of backward (characteristic) SODEs

dXx,y(t) = cos(Xx,y + Yx,y)←−−dW1(t) + sin(Xx,y + Yx,y)

←−−dW2(t), t ∈ [0, T ]

dYx,y(t) = − cos(Xx,y + Yx,y)←−−dW1(t)− sin(Xx,y + Yx,y)

←−−dW2(t), t ∈ [0, T ]

Xx,y(T ) = x, Yx,y(T ) = y

(4.17)

The definition of backward Ito integral←−−dW (t) can also be found in [23]. Summing the two

equations, we realize that Xx,y(t) + Yx,y(t) is constant over time. Therefore

Xx,y(0) = x− cos(x+ y)η1− sin(x+ y)η2, Yx,y(0) = y +cos(x+ y)η1 +sin(x+ y)η2 (4.18)

28

Page 29: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

where η1 = W1(T ) and η2 = W2(T ) are independent Gaussian variables. Then the analytical

solution is

u(T, x, y) = sin(2(x−cos(x+y)η1−sin(x+y)η2)) cos(y+cos(x+y)η1 +sin(x+y)η2) (4.19)

We note that the distribution of u(T, x, y) is again supported in [−1, 1]. Since η1 =√

Tξ11

and η2 =√

Tξ21 , the solution just depends on ξ1

1 and ξ21 . In terms of truncation, we can

only take K = 1 and adjust the value of N . Such simplification is actually a direct result of

Remark 3.3. It is easy to show thatM1 andM2 commute with each other, so that they also

commute with 12(M2

1 +M22). Commutativity is the reason why K = 1 is enough. Monte

Carlo sampling of (4.19) is trivial. Moments of u(T, x, y) can be computed very accurately

through Gauss-Hermite quadrature rule. We pick 50 quadrature points in each dimension

to establish reference solutions.

The N -version convergence of mean square truncation error is displayed in Figure 4.6.

We plot values of E[‖uN,1(0.2, ·, ·)−u(0.2, ·, ·)‖2l2] in logarithm scale for N = 1, · · · , 10, where

the discrete L2 norm is

‖v‖2l2 :=4π2

M2

M∑

i=1

M∑

j=1

(v(xi, yj))2

The convergence rate is clearly exponential.

1 2 3 4 5 6 7 8 9 10

N

10−5

10−4

10−3

10−2

10−1

100

101

[‖(

uN,1−

u)(0.

2,·,

·)‖2 l2]

Figure 4.6: Example 4.5: Semi-log plot of mean square truncation error E[‖uN,1(0.2, ·, ·)−u(0.2, ·, ·)‖2

l2] for N = 1, · · · , 10.

Next we fix N = 8 and pay attention to higher moments. Figure 4.7 presents contour

plots for the third and fourth central moments of u8,1(0, 2, ·, ·). For Gaussian noise, we also

provide third and fourth central moments of u(0.2, ·, ·) using Gauss-Hermite quadrature. The

agreement between the truncated solution and the reference solution is quite satisfactory. For

29

Page 30: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

uniform noise and Beta(12, 1

2) noise, the corresponding contour plots have different patterns,

especially for the third central moment.

We can also detect the impact of driving noise by checking empirical distributions. In

Figure 4.8 we demonstrate normalized histograms of uN,1 at the collocation point (x6, y6)

out of 107 samples. For Gaussian noise, we choose N = 4, 8, together with the reference

distribution generated by 107 samples of (4.19). Samples outside [−1, 1] are discarded.

Similar to Figure 4.5, u8,1 outperforms u4,1 in approximating the highly rightly skewed true

distribution. For uniform noise and Beta(12, 1

2) noise, we consider N = 8, 10. The empirical

density functions of u8,1 and u10,1 are mostly overlapping, so that they can be thought as

credible approximations. Once again we notice the fact that different driving noises lead to

strikingly different empirical density profiles.

Example 4.6 (Stochastic Burgers equation). We consider the stochastic Burgers equation

in Example 2.3.2. The space region is D = [0, 1] with periodic boundary condition. The

parameters are µ = 0.005 and σ(x) = 12cos(4πx), and the initial data is u0(x) = 1

2(e(2πx) −

1.5) sin(2π(x + 0.37)). We apply the Fourier collocation method with M = 128 collocation

points for spatial discretization. The end time is set to be T = 0.8 with Runge-Kutta time

step δt = 10−3.

For such additive noise, Monte Carlo simulation is well suited for any type of distribution.

We generate sample paths of N(t) by truncating the infinite sum up to K = 50. For a fixed

sample path, we solve the resulting deterministic Burgers equation using Fourier collocation

method and fourth order Runge-Kutta time stepping, with M = 128 collocation points and

time step δt = 10−3. Moments and empirical distributions of Monte Carlo samples will be

chosen as reference solutions. We take 106 samples.

The convergence of mean square truncation error is given in Figure 4.9. We again plot

E[‖uN,K(0.8, ·) − u20,K(0.8, ·)‖2l2] to represent N -version convergence, and E[‖uN,K(0.8, ·) −

uN,50(0.8, ·)‖2l2] to represent K-version convergence. Due to nonlinearity, these truncation

errors rely on the underlying randomness. For the N -version convergence, we only plot results

with uniform and Beta(12, 1

2) noise as the interaction coefficients of Hermite polynomials grow

exponentially with respect to N , causing the numerical computation to blow up for large N .

For the K-version convergence, the plots are nearly identical for three noises, so that we only

present the plot with Gaussian noise. We emphasize that the numerical convergence rate is

still exponential with respect to N and cubic with respect to K, even though Theorem 3.1

is only proved for the linear case.

Then we fix K = 8. Third and fourth central moments of u2,8(0.8, ·) and u5,8(0.8, ·)are drawn in Figure 4.10. As the profiles with different noises are close to each other, we

only show the zoomed-in view between the 21-st collocation point and the 50-th collocation

30

Page 31: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

0 π/2 π 3π/2 2π0

π/2

π

3π/2

−0.072

−0.048

−0.024

0.000

0.024

0.048

0.072

(a) Third central moment: Gaussian noise(reference solution)

0 π/2 π 3π/2 2π0

π/2

π

3π/2

0.004

0.020

0.036

0.052

0.068

0.084

0.100

0.116

(b) Fourth central moment: Gaussian noise(reference solution)

0 π/2 π 3π/2 2π0

π/2

π

3π/2

−0.072

−0.048

−0.024

0.000

0.024

0.048

0.072

(c) Third central moment: Gaussian noise0 π/2 π 3π/2 2π

0

π/2

π

3π/2

0.008

0.024

0.040

0.056

0.072

0.088

0.104

0.120

(d) Fourth central moment: Gaussian noise

0 π/2 π 3π/2 2π0

π/2

π

3π/2

−0.036

−0.024

−0.012

0.000

0.012

0.024

0.036

(e) Third central moment: uniform noise0 π/2 π 3π/2 2π

0

π/2

π

3π/2

0.003

0.015

0.027

0.039

0.051

0.063

0.075

0.087

(f) Fourth central moment: uniform noise

0 π/2 π 3π/2 2π0

π/2

π

3π/2

−0.0300

−0.0225

−0.0150

−0.0075

0.0000

0.0075

0.0150

0.0225

0.0300

(g) Third central moment: Beta(1

2, 1

2) noise

0 π/2 π 3π/2 2π0

π/2

π

3π/2

0.004

0.016

0.028

0.040

0.052

0.064

0.076

0.088

0.100

(h) Fourth central moment: Beta(1

2, 1

2) noise

Figure 4.7: Example 4.5: Third and fourth central moments of u8,1(0.2, ·, ·). First two plotsare reference solutions with Gaussian noise. 30 equally spaced contour levels are used for allplots.

31

Page 32: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

−1 −0. 5 0 0. 5 1

u

0

5

10

15

20

25

Density

N=4

N=8

Monte Carlo

(a) Gaussian noise

−0. 5 0 0. 5 1

u

0

1

2

3

4

Density

Uniform: N=8

Uniform: N=10

Beta: N=8

Beta: N=10

(b) Uniform and Beta(1

2, 1

2) noise

Figure 4.8: Example 4.5: Normalized histograms of uN,1(0.2, x6, y6) out of 107 i.i.d samples.In the left panel, we take N = 4, 8. Values outside [−1, 1] are discarded, and the blackdashed line represents the normalized histogram of 107 Monte Carlo samples of (4.19). Inthe right panel, we take N = 8, 10. Number of bins is 103.

1 2 3 4 5 6 7 8 9 10

N

10−11

10−10

10−9

10−8

10−7

10−6

10−5

10−4

10−3

[‖(

uN,1−

u20

,1)(0.

8,·)‖

2 l2]

Uniform noise

Beta(12, 12) noise

(a) Convergence with respect to N

1 2 3 4 5 6 7 8 9 10 11 12

K

10-6

10-5

10-4

10-3

10-2

10-1

[‖(

uN,K

−uN,5

0)(0.

8,·)‖

2 l2]

1 : 3

Gaussian noise: N=1

Gaussian noise: N=2

(b) Convergence with respect to K

Figure 4.9: Example 4.6: Plots of mean square truncation error in discrete l2 norm withrespect to N and K. Left panel shows the semi-log plot of E[‖uN,1(0.8, ·) − u20,1(0.8, ·)‖2l2]versus N with uniform and Beta(1

2, 1

2) noise for N = 1, · · · , 10. Right panel shows the

log-log plot of E[‖uN,K(5, ·)− uN,50(5, ·)‖2l2] versus K with Gaussian noise for N = 1, 2 andK = 1, · · · , 12.

point. Central moments of Monte Carlo solution are also plotted for comparison. We note

that for all noises, u2,8 results in inaccurate approximations, and u5,8 leads to much better

performance. The plots of empirical density functions of u2,8 and u5,8 out of 107 samples at

x30 are provided in Figure 4.11. From the figure we can also see how u5,8 is superior to u2,8

32

Page 33: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

in agreeing with the reference distributions.

0. 2 0. 25 0. 3 0. 35

x

−0. 01

0

0. 01

Third central moment

Gaussian: N=2, K=8

Gaussian: N=5, K=8

Gaussian: Monte Carlo

Uniform: N=2, K=8

Uniform: N=5, K=8

Uniform: Monte Carlo

Beta: N=2, K=8

Beta: N=5, K=8

Beta: Monte Carlo

(a) Third central moment

0. 2 0. 25 0. 3 0. 35

x

0

0. 01

0. 02

Fourth central moment

Gaussian: N=2, K=8

Gaussian: N=5, K=8

Gaussian: Monte Carlo

Uniform: N=2, K=8

Uniform: N=5, K=8

Uniform: Monte Carlo

Beta: N=2, K=8

Beta: N=5, K=8

Beta: Monte Carlo

(b) Fourth central moment

Figure 4.10: Example 4.6: Third and fourth central moments of uN,K(0.8, ·) between x21 andx50. We take N = 2, 5 and K = 8. Dashed lines are reference solutions computed by MonteCarlo simulation with 106 samples.

5 Concluding remarks

In this paper, we explore the numerical solutions of SPDEs by truncating the stochastic

polynomial chaos expansion series under the distribution-free framework. We generalize the

definition of Wick product and Skorokhod integral to arbitrary driving noise. Then Wick-

Skorokhod type SPDEs are not limited to Gaussian randomness or Levy randomness. More

importantly, for linear SPDEs, the propagator system, and even the first two moments or

the solution, are the same for different noises. The computational burden of solving the

propagator system is purely off-line. The only on-line work is post-processing. However,

the propagator system of nonlinear SPDE changes with noise as interaction terms come into

play.

Analysis of the mean square truncation error is carried out for linear problems. We

prove exponential convergence with respect to polynomial order, and cubic convergence with

respect to the number of random variables. The cubic rate arises from repeated integration-

by-parts and special properties of the orthonormal basis mk(t)∞k=1. We need to assume

trigonometric basis or Legendre basis.

We conduct systematic investigation on the numerical results of linear and nonlinear

SPDEs with different driving noises. Numerical rates of convergence are consistent with

our theoretical analysis. Higher moments and density function can also be approximated

33

Page 34: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

−1 −0. 5 0 0. 5 1

u

0

1

2Density

N=2, K=8

N=5, K=8

Monte Carlo

(a) Gaussian noise

−1 −0. 5 0 0. 5 1

u

0

0. 5

1

1. 5

Density

N=2, K=8

N=5, K=8

Monte Carlo

(b) Uniform noise

−1 −0. 5 0 0. 5 1

u

0

0. 5

1

1. 5

Density

N=2, K=8

N=5, K=8

Monte Carlo

(c) Beta(1

2, 1

2) noise

Figure 4.11: Example 4.6: Normalized histograms of uN,K(0.8, x30) out of 107 i.i.d samples.We take N = 2, 5 and K = 8. Black dashed line represents the normalized histogram ofMonte Carlo simulation with 106 samples. Number of bins is 103.

effectively with sufficiently many polynomial chaos expansion terms. However, we should also

recognize some drawbacks and unsolved problems, which gives hints on our future research.

(i) To the best of our knowledge, the limiting procedure of distribution-free Skorokohd

integral is unclear. Proposition 2.2 is only for deterministic processes, and Proposition

2.3 is only for Gaussian (and Levy) noise. Then we can only come up with reference

solutions for Ito type SPDEs with Gaussian (and Levy) noise and / or SPDEs with

additive noise. Further work is required for better understanding of the distribution-

free stochastic analysis.

(ii) We do not focus on long time integration in this paper, but the exponential growth of

34

Page 35: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

error with respect to time is seen both theoretically and numerically. Proper techniques

should be devised to mitigate the impact of time evolution. We remark that direct

generalization of the multi-stage methodology in [12, 11, 36] is specious as the driving

process N(t) may not have independent increments.

(iii) The propagator system usually consists of PDEs of the same type but with different

data (e.g. different coefficients, different given right hand sides, etc.). It can be ex-

pected that for a large number of expansion terms, the application of reduced basis

method [25] may significantly reduce the computational cost while maintaining desired

accuracy.

Acknowledgement. The authors would like to thank Michael Tretyakov and Zhongqiang

Zhang for helpful discussions on the relation between commutativity and K-version conver-

gence.

A Interaction coefficients B(α, β, p)

We still assume that ξk∞k=1 are i.i.d. random variables. Then the interaction coefficient

B(α, β, p) can be decomposed into

B(α, β, p) =E[ΦαΦβΦp]

α!=

∞∏

k=1

E[ϕαk(ξk)ϕβk

(ξk)ϕpk(ξk)]

αk!:=

∞∏

k=1

b(αk, βk, pk) (A.1)

It suffices to compute b(i, j, l) for any i, j, l ≥ 0. According to orthogonality,

ϕj(ξ)ϕl(ξ) =∞∑

i=0

E[ϕi(ξ)ϕj(ξ)ϕl(ξ)]

i!ϕi(ξ) =

∞∑

i=0

b(i, j, l)ϕi(ξ) (A.2)

Hence b(i, j, l) is the i-th expansion coefficient of ϕj(ξ)ϕl(ξ) in terms of ϕn(ξ)∞n=0. In

particular, for the three types of noises and corresponding orthogonal polynomials considered

in Section 4, there are explicit formulas for these expansion coefficients.

• For Gaussian noise and Hermite chaos. ϕn(ξ) = Hen(ξ). Since

Hej(x)Hel(x) =

minj,l∑

r=0

j!l!

(j − r)!(l − r)!r!Hej+l−2r(x) (A.3)

we have

b(i, j, l) =

j!l!

(j−r)!(l−r)!r!if i = j + l − 2r and r ≤ mini, j

0 otherwise(A.4)

35

Page 36: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

• For uniform noise and Legendre chaos, ϕn(ξ) =√

(2n + 1)n!Ln(ξ/√

3). Define

λn :=Γ(n + 1/2)

n!Γ(1/2)=

∏n−1m=0(m + 1/2)

n!

Then the expansion of Lj(x)Ll(x) is

Lj(x)Ll(x) =

minj,l∑

r=0

2(j + l − 2r) + 1

2(j + l − r) + 1

λrλi−rλj−r

λi+j−r

Lj+l−2r(x) (A.5)

Thus

b(i, j, l) =

√(2i+1)(2j+1)(2l+1)

2(j+l−r)+1

√j!l!i!

λrλi−rλj−r

λi+j−rif i = j + l − 2r and r ≤ mini, j

0 otherwise

(A.6)

• For Beta(12, 1

2) noise and Chebyshev chaos, ϕn(ξ) =

√cnn!Tn(ξ/

√2) where c0 = 1 and

cn = 2 for n ≥ 1. Since Chebyshev polynomials are essentially cosine functions,

Tj(x)Tl(x) =1

2Tj+l(x) +

1

2T|j−l|(x) (A.7)

Thus

b(i, j, l) =

1 if i = j, l = 0 or i = l, j = 0

12

√cjcl

ci

√j!l!i!

if j, l > 0 and i = j + l or i = |j − l|0 otherwise

(A.8)

Here the expansion coefficients have a sparse pattern. For fixed j and l, there are at

most two values of i such that b(i, j, l) is nonzero.

In general, we compute b(i, j, l) by matching the monomial coefficients on the both sides

of (A.2) (see e.g., [38]). Suppose that

ϕn(ξ) =n∑

m=0

Pm,nξm

According to (A.2), for i > j + l, b(i, j, l) = 0, and b(i, j, l) : 0 ≤ i ≤ j + l satisfies the

following linear system

j+l∑

i=0

b(i, j, l)Pm,i =

mini,j∑

r=max0,i−l

Pr,jPi−r,l (A.9)

It is very easy to solve (A.9) as Pm,nj+lm,n=0 is a upper triangular matrix. This procedure is

applicable to any set of orthogonal polynomials.

36

Page 37: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

References

[1] R. Askey and J. A. Wilson. Some Basic Hypergeometric Orthogonal Polynomials that

Generalize Jacobi polynomials. American Mathematical Society, 1985.

[2] F. E. Benth and J. Gjerde. A remark on the equivalence between Poisson and Gaussian

stochastic partial differential equations. Potential Analysis, 8:179–193, 1998.

[3] R. H. Cameron and W. T. Martin. The orthogonal development of non-linear functionals

in series of Fourier-Hermite functionals. Annals of Mathematics, 385–392, 1947.

[4] B. Cockburn and C.-W. Shu. The local discontinuous Galerkin method for time-

dependent convection-diffusion systems. SIAM Journal on Numerical Analysis, 35:2440–

2463, 1998.

[5] B. J. Debusschere, H. N. Najm, P. P. Pebay, O. M. Knio, R. O. Ghanem and O. P.

Le Maıtre. Numerical challenges in the use of polynomial chaos representations for

stochastic processes. SIAM Journal on Scientific Computing, 26:698–719, 2004.

[6] G. Da Prato, A. Debussche and R. Temam. Stochastic Burgers’ equation. Nonlinear

Differential Equations and Applications NoDEA, 1:389–402, 1994.

[7] G. Di Nunno, B. Øksendal and F. Proske. Malliavin Calculus for Levy Processes with

Applications to Finance. Springer, 2009.

[8] H. Holden, B. Øksendal, J. Ubøe and T. Zhang. Stochastic Partial Differential Equa-

tions. Springer, 1996.

[9] T. Y. Hou, W. Luo, B. Rozovskii and H. Zhou. Wiener chaos expansions and numerical

solutions of randomly forced equations of fluid mechanics. Journal of Computational

Physics, 216:687–706, 2006.

[10] K. Ito. Multiple Wiener integral. Journal of the Mathematical Society of Japan, 3:157–

169, 1951.

[11] H. Liu. On spectral approximations of stochastic partial differential equations driven by

Poisson noise. Ph.D. thesis, University of Southern California, 2007.

[12] S. Lototsky, R. Mikulevicius, and B. Rozovskii. Nonlinear filtering revisited: a spectral

approach. SIAM Journal on Control and Optimization, 35:435–461, 1997.

37

Page 38: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

[13] S. Lototsky and B. Rozovskii. Passive scalar equation in a turbulent incompressible

Gaussian velocity field. Russian Mathematical Surveys, 59:297–312, 2004.

[14] S. Lototsky and B. Rozovskii. Stochastic differential equations: a Wiener chaos ap-

proach. In From Stochastic Calculus to Mathematical Finance, Springer, 2006, pp.433–

506.

[15] S. Lototsky and B. Rozovskii. Wiener chaos solutions of linear stochastic evolution

equations. The Annals of Probability, 638–662, 2006.

[16] S. Lototsky and B. Rozovskii. Stochastic Partial Differential Equations. Springer, 2017.

[17] W. Luo. Wiener chaos expansion and numerical solutions of stochastic partial differen-

tial equations. Ph.D. thesis, California Institute of Technology, 2006.

[18] P. Malliavin. Stochastic calculus of variation and hypoelliptic operators. In Proc. Intern.

Symp. SDE Kyoto 1976, Kinokuniya, 1978, pp.195–263.

[19] R. Mikulevicius and B. Rozovskii. Separation of observations and parameters in nonlin-

ear filtering. In Decision and Control, Proceedings of the 32nd IEEE Conference, IEEE,

1993, pp.1564–1569.

[20] R. Mikulevicius and B. Rozovskii. Linear parabolic stochastic PDE and Wiener chaos.

SIAM Journal on Mathematical Analysis, 29:452–480, 1998.

[21] R. Mikulevicius and B. Rozovskii. On unbiased stochastic Navier–Stokes equations.

Probability Theory and Related Fields, 154:787–834, 2012.

[22] R. Mikulevicius and B. Rozovskii. On distribution free Skorokhod–Malliavin calculus.

Stochastics and Partial Differential Equations: Analysis and Computations, 4:319–360,

2016.

[23] G. Milstein and M. Tretyakov. Solving parabolic stochastic partial differential equations

via averaging over characteristics. Mathematics of Computation, 78:2075–2106, 2009.

[24] G. Milstein and M. Tretyakov. Stochastic Numerics for Mathematical Physics. Springer,

2013.

[25] A. Quarteroni, A. Manzoni and F. Negri. Reduced Basis Methods for Partial Differential

Equations: an Introduction. Springer, 2015.

[26] A. V. Skorokhod. On a generalization of a stochastic integral. Theory of Probability &

Its Applications, 20:219–233, 1976.

38

Page 39: Abstract - Brown University...able, numerical methods have to be designed to solve them. The most popular approach is the Monte Carlo method, which generates independent random sample

[27] D. Venturi, X. Wan, R. Mikulevicius, B. Rozovskii, and G. Karniadakis. Wick–Malliavin

approximation to nonlinear stochastic partial differential equations: analysis and simu-

lations. In Proc. R. Soc. A, The Royal Society, 2013, pp.1–20.

[28] X. Wan and B. Rozovskii. The Wick–Malliavin approximation of elliptic problems with

log-normal random coefficients. SIAM Journal on Scientific Computing, 35:A2370–

A2392, 2013.

[29] G. C. Wick. The evaluation of the collision matrix. Physical Review, 80:268–272, 1950.

[30] D. Xiu and G. Karniadakis. The Wiener–Askey polynomial chaos for stochastic differ-

ential equations. SIAM Journal on Scientific Computing, 24:619–644, 2002.

[31] D. Xiu and G. Karniadakis. Modeling uncertainty in flow simulations via generalized

polynomial chaos. Journal of Computational Physics, 187:137–167, 2003.

[32] D. Xiu and G. Karniadakis. Supersensitivity due to uncertain boundary conditions.

International Journal for Numerical Methods in Engineering, 61:2114–2138, 2004.

[33] D. Xiu. Numerical Methods for Stochastic Computations: a Spectral Method Approach.

Princeton University Press, 2010.

[34] Z. Zhang, B. Rozovskii, M. Tretyakov and G. Karniadakis. A multistage Wiener chaos

expansion method for stochastic advection-diffusion-reaction equations. SIAM Journal

on Scientific Computing, 34:A914–A936, 2012.

[35] Z. Zhang, M. Tretyakov, B. Rozovskii and G. Karniadakis. A recursive sparse grid col-

location method for differential equations with white noise. SIAM Journal on Scientific

Computing, 36:A1652–A1677, 2014.

[36] Z. Zhang, M. Tretyakov, B. Rozovskii and G. Karniadakis. Wiener chaos versus stochas-

tic collocation methods for linear advection-diffusion-reaction equations with multiplica-

tive white noise. SIAM Journal on Numerical Analysis, 53:153–183, 2015.

[37] Z. Zhang and G. Karniadakis. Numerical Methods for Stochastic Partial Differential

Equations with White Noise. Springer, 2017.

[38] M. Zheng, B. Rozovskii and G. Karniadakis. Adaptive Wick–Malliavin approxima-

tion to nonlinear SPDEs with discrete random variables. SIAM Journal on Scientific

Computing, 37:A1872–A1890, 2015.

39