1,001
Views
4
CrossRef citations to date
0
Altmetric
Research Article

A randomized trapezoidal quadrature

ORCID Icon
Pages 680-692 | Received 14 Dec 2020, Accepted 01 May 2021, Published online: 21 May 2021

Abstract

A randomized trapezoidal quadrature rule is proposed for continuous functions which enjoy less regularity than commonly required. Indeed, we consider functions in some fractional Sobolev space. Various error bounds for the randomized rule are established while an error bound for classical trapezoidal quadrature is obtained for comparison. The randomized trapezoidal quadrature rule is shown to improve the order of convergence by half.

2010 Mathematics Subject Classifications:

1. Introduction

It is well known that the trapezoidal quadrature in classical numerical analysis is a technique for approximating Rd-valued definite integral when the integrand is at least twice differentiable. It has been well-studied by Goodwin [Citation7], Schwartz [Citation9] and Stenger [Citation10]. The case of a rough integrand was investigated in [Citation3]. More recently, a stochastic version of the trapezoidal quadrature was proposed for approximating the Itô integral where the integrator is a Brownian motion [Citation5].

Without loss of generality, we consider the time interval [0,T] and let gC2([0,T]) to be the integrand of interest, where C2([0,T]):=C2([0,T];Rd) is the space of Rd-valued continuous functions that have continuous first two derivatives, endowed with the uniform norm topology. The trapezoidal quadrature is proven to achieve a order of convergence as high as 2 for evaluating the integral I[g]:=0Tg(t)dt with finite many point evaluations [Citation4]. To implement this, we first partition the interval [0,T] into N equidistant intervals with stepsize hN=TN, i.e. (1) Πh:={tj:=jh}j=0N[0,T],(1) where the subscription N is suppressed in h for the sake of notational simplicity but assumed implicitly in all of the quantities introduced involving h. Define (2) Qh[g]:=h2i=0N1(g(ti)+g(ti+1)).(2) When g has less regularity, the trapezoidal quadrature shows a slower convergence with a sharp bound [Citation3]. To accelerate the convergence when g is ‘rougher’, we consider a randomized trapezoidal quadrature, which is inspired by the randomized version of mid-point Runge–Kutta quadrature rule [Citation8] and stochastic version of trapezoidal quadrature for Itô integral [Citation5]. In this paper, the Rd-valued integrand g is assumed to be in fractional Sobolev space Wσ,p(0,T) under Sobolev–Slobodeckij norm: (3) gWσ,p(0,T)=(0T|g(t)|pdt+0T|g˙(t)|pdt+0T0T|g˙(t)g˙(s)|p|ts|1+(σ1)pdtds)1p,(3) for σ(1,2) and p[2,). We may write gWσ,p(0,T) as gWσ,p for short. Let us define a randomized trapezoidal quadrature as follows: (4) RQhτ,n[g]:=h2i=0n1(g(ti+τih)+g(ti+τ¯ih))forn[N],(4) where {τi}i=0N1 is a sequence of independent and identically (i.i.d.) uniformly distributed random variables on a probability space (Ω,F,P), τ¯i:=1τi and [N]:={1,,N}. The main result, Theorem 3.2, shows that the convergence rate can be improved to O(Nσ12) compared to O(Nσ) achieved by the classical trapezoidal quadrature (Theorem 3.1) when gWσ,p.

The paper is organized as follows. In Section 2, we present some prerequisites from probability theory. In Section 3, we give the error estimates for both the classical trapezoidal quadrature and the randomized trapezoidal quadrature. In addition, we also investigate the error estimate in almost sure sense for the randomized trapezoidal quadrature in Theorem 3.3, which is proven still superior to the classical one. In the last section, we verify the results through several numerical experiments.

2. Preliminaries

This section is devoted to a brief review of essential probability results for audience who are not familiar with probability theory. Most of the contents are repeated material from Section 2 in [Citation8]. One may refer to [Citation2] for a more detailed introduction.

Recall that a probability space (Ω,F,P) consists of a measurable space (Ω,F) endowed with a finite measure P satisfying P(Ω)=1. A random variable X:ΩRd is called integrable if Ω|X(ω)|dP(ω)<. Then, the expectation of X is defined as E[X]:=ΩX(ω)dP(ω)=RdxdμX(x),where μX is distribution of X on its image space. We write XLp(Ω;Rd) with p[1,) if Ω|X(ω)|pdP(ω)<, where Lp(Ω;Rd) is a Banach space endowed with the norm XLp(Ω;Rd)=(E[|X|p])1p=(Ω|X(ω)|pdP(ω))1p.We will write XLp(Ω;R) as XLp(Ω) for short.

We say that a family of Rd-valued random variables (Xm)mN is a discrete-time stochastic process if we interpret the index m as a time parameter. A crucial concept in our main proof is martingales, which is a special case of the discrete-time stochastic process with many nice properties. If (Xm)mN is an independent family of integrable random variables satisfying E[Xm]=0 for each mN, then the stochastic process defined by the partial sums Sn:=m=1nXm,nN,is a discrete-time martingale. One of the most important inequalities for martingales is the Burkholder–Davis–Gundy inequality. In this paper, we need its discrete-time version.

Theorem 2.1

Burkholder–Davis–Gundy

For each p(1,) there exist positive constants cp and Cp such that for every discrete time martingale (Xn)nN and for every nN we have cp[X]n1/2Lp(Ω)maxj{1,,n}|Xj|Lp(Ω)Cp[X]n1/2Lp(Ω),where [X]n=|X1|2+k=2n|XkXk1|2 denotes the quadratic variation of (Xn)nN up to n.

3. Trapezoidal quadratures for a rougher integrand

This section investigates the errors from trapezoidal rules for approximating integral of gWσ,p. The error bound from the classical trapezoidal rule is obtained in Section 3.1 and the ones from the randomized trapezoidal rule is in Section 3.2.

3.1. Classical trapezoidal quadrature for gWσ,p

Theorem 3.1

If gWσ,p(0,T) for σ[1,2), then we have (5) |I[g]Qh[g]|CT11phσgWσ,p(0,T),(5) where C is a constant that only depends on p.

Proof.

To show Equation (Equation5), we follow [Citation5] to rewrite (6) g(ti)+g(ti+1)=2g(ti+12)+ti+12tig˙(s)ds+ti+12ti+1g˙(s)ds,(6) where ti+12:=12(ti+ti+1). Then, the LHS of Equation (Equation5) can be rewritten as I[g]Qh[g]=i=0N1E1i,i+1+i=0N1E2i,i+1,where (7) E1i,i+1:=titi+1(g(t)g(ti+12))dt=12titi+1ti+12tg˙(r)drdt(7) and (8) E2i,i+1:=12titi+1(ti+12tig˙(s)ds+ti+12ti+1g˙(s)ds)dt.(8) Regarding E1i,i+1, first note that Vi:=1h(g(ti+1)g(ti))titi+1(tti+12)dt=1htiti+1ti+12ttiti+1g˙(s)dsdrdt=0.Then, we can rewrite E1i,i+1 as (9) E1i,i+1=E1i,i+1Vi=1htiti+1titi+1ti+12tg˙(r)drdsdtVi=1hi=0N1titi+1titi+1ti+12t(g˙(r)g˙(s))drdsdt.(9) Thus evaluating i=0N1E1i,i+1 under Lp norm gives (10) |i=0N1E1i,i+1|i=0N1titi+1titi+1|g˙(r)g˙(s)|drdsi=0N1h2q(titi+1titi+1|g˙(r)g˙(s)|pdrds)1p,(10) where the second line is deduced by applying Hölder's inequality twice, and 1q:=11p. For the case σ=1 and any p2, we may directly apply the discrete Hölder's inequality to the last term above: i=0N1h2q(titi+1titi+1|g˙(r)g˙(s)|pdrds)1pCi=0N1h2q+1p(titi+1|g˙(r)|pdr)1pCh(i=0N1h)1q(i=0N1titi+1|g˙(r)|pdr)1p=ChT11pgW1,p.For the case σ(1,2) and any p2, we may first make use of the definition of Wσ,p and then apply the discrete Hölder's inequality: i=0N1h2q(titi+1titi+1|g˙(r)g˙(s)|pdrds)1pi=0N1h2q+1p+σ1(titi+1titi+1|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)1p=hσi=0N1h1q(titi+1titi+1|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)1phσT11pgW1,p.For term E2i,i+1, we can follow a similar argument in [Citation5] to show that (11) E2i,i+1=12htiti+1titi+1(ti+12ti(g˙(s)g˙(r))dr+ti+12ti+1(g˙(s)g˙(r))dr)dsdt.(11) Indeed, note that (titi+12)+(ti+1ti+12)=0,If defining a new process Pi:=titi+122htiti+1titi+1g˙(r)drdt+ti+1ti+122htiti+1titi+1g˙(r)drdt,then Equation (Equation11) can be obtained through the fact that E2i,i+1=E2i,i+1Pi.Thus applying a similar argument as for i=0N1E1i,i+1, we can show that (12) |i=0N1E2i,i+1|CT11phσgWσ,p[0,T].(12) Finally, we can conclude that |I[g]Qh[g]||i=0N1E1i,i+1|+|i=0N1E2i,i+1|CT11phσgWσ,p[0,T].

For the classical trapezoidal quadrature (CTQ), Theorem 3.1 claims that its order of convergence would be the same as the regularity of the integrand. For the boundary case, when gW1,p, the order is 1.

3.2. Randomized trapezoidal rules for gWσ,p

For the randomized trapezoidal quadrature (Equation4), the proof follows a similar argument as in Theorem 4.2 [Citation8].

Theorem 3.2

Define In:=0tng(t)dt for n[N] for gWσ,p with σ[1,2) and p2. Then RQhτ,n[g]Lp(Ω;Rd) and is an unbiased estimator of In[g], i.e. E[RQhτ,n[g]]=In[g]. Moreover, it holds true that (13) I[g]RQhτ,N[g]Lp(Ω;Rd)Cp|T|p22ph12+σgWσ,p(0,T),(13) where Cp is a constant that depends only on p.

Proof.

First, due to gWσ,p we have gLp([0,T];Rd)<. Recall that τiU(0,1) for each i[N1]{0}. Then, it follows that h2(g(ti+τih)Lp(Ω;Rd)p+g(ti+τ¯ih)Lp(Ω;Rd)p)=titi+1|g(t)|pdt<.Hence RQhτ,n[g]Lp(Ω) for n[N]. To show RQhτ,n[g] is unbiased, we need to examine each term in RHS of Equation (Equation4) through spelling out the expectation and changing variable, i.e. h2E[g(ti+τih)]=h201g(ti+rh)dr=12titi+1g(t)dtand h2E[g(ti+τ¯ih)]=h201g(ti+(1r)h)dr=12titi+1g(t)dt.Summing these terms up gives that RQhτ,n[g] is unbiased for In[g]. Furthermore, if define the error term like (14) En:=In[g]RQhτ,n[g]=12i=0n1titi+1(2g(t)g(ti+τih)g(ti+τ¯ih))dt,(14) then each summand is a mean-zero random variable, i.e. E[titi+1(2g(t)g(ti+τih)g(ti+τ¯ih))dt]=0.Note that the summands are mutually independent due to the independence of {τi}i=0N1. In addition, it is easy to show EnLp(Ω;Rd). Therefore, En is a Lp-martingale. Then applying the discrete version of the Burkholder–Davis–Gundy inequality leads to (15) maxn|En|Lp(Ω)Cp[En]N12Lp(Ω)=Cp2(i=0N1|titi+1(2g(t)g(ti+τih)g(ti+τ¯ih))dt|2)12Lp(Ω)Cp(i=0N1|titi+1(g(t)g(ti+τih)))dt|2)12Lp(Ω)+Cp(i=0N1|titi+1(g(t)g(ti+τ¯ih))dt|2)12Lp(Ω),(15) where in the second line we substitute the quadratic variation [En]N. Due to symmetric property, it is easy to see we only need to handle the first term on the RHS of Equation (Equation15). Note that (16) Cp(i=0N1|titi+1(g(t)g(ti+τih)))dt|2)12Lp(Ω)=Cpi=0N1|titi+1(g(t)g(ti+τih)))dt|2Lp2(Ω)12Cp(i=0N1titi+1(g(t)g(ti+τih)))dtLp(Ω;Rd)2)12.(16) Then we have that Cp(i=0N1titi+1(g(t)g(ti+τih)))dtLp(Ω;Rd)2)12=Cp(i=0N1titi+1ti+τihtg˙(s)dsdtLp(Ω;Rd)2)12Cp(i=0N1titi+1ti+τiht|g˙(s)|dsdtLp(Ω)2)12Cph(i=0N1|titi+1|g˙(s)|ds|2)12Cph(i=0N1h2q|titi+1|g˙(s)|pds|2p)12.When p = 2, the term on the right-hand side above can be directly bounded by (17) Cph(i=0N1h2q|titi+1|g˙(s)|pds|2p)12Cph32gW1,p,(17) where 1q+1p=1. For p>2, we may apply discrete Hölder inequality and get (18) Cph(i=0N1h2q|titi+1|g˙(s)|pds|2p)12Cph(i=0N1h2qpp2)p22p(i=0N1titi+1|g˙(s)|pds)1pCph1+(2pq(p2)1)p22p|T|p22pgW1,p=Cph32|T|p22pgW1,p.(18) Now, we have shown Bound (Equation13) when σ=1. For Bound (Equation13) under σ>1, we first note that Equation (Equation6) remains true if replacing ti by ti+τih and ti+1 by ti+τ¯ih, i.e. (19) g(ti+τih)+g(ti+τ¯ih)=2g(ti+12)+ti+12ti+τihg˙(s)ds+ti+12ti+τ¯ihg˙(s)ds.(19) Thus, the second line of Equation (Equation15) can be further split as the follows: Cp2(i=0N1|titi+1(2g(t)g(ti+τih)g(ti+τ¯ih))dt|2)12Lp(Ω)Cp(i=0N1|titi+1(g(t)g(ti+12))dt|2)12Lp(Ω)+Cp(i=0N1|titi+1ti+12ti+τihg˙(s)dsdt+titi+1ti+12ti+τ¯ihg˙(s)dsdt|2)12Lp(Ω).Similar as in the proof of Theorem 3.1, we introduce E1i,i+1 defined in Equation (Equation7) and (20) E2i,i+1(τ):=12titi+1(ti+12ti+τhg˙(s)ds+ti+12ti+τ¯hg˙(s)ds)dt.(20) As in the proof of Theorem 3.1, E1i,i+1 can be handled through the equivalent form Equation (Equation9) and E2i,i+1(τ) can be treated in a similar way as Equation (Equation11) by replacing ti by ti+τih and ti+1 by ti+τ¯ih in the inner integral of Equation (Equation11), i.e. (21) E2i,i+1=12htiti+1titi+1(ti+12ti+τih(g˙(s)g˙(r))dr+ti+12ti+τ¯ih(g˙(s)g˙(r))dr)dsdt.(21) Thus Cp2(i=0N1|titi+1(2g(t)g(ti+τih)g(ti+τ¯ih))dt|2)12Lp(Ω)Cp(iN1|E1i,i+1|2)12Lp(Ω)+Cp(iN1|E2i,i+1|2)12Lp(Ω)=Cph(i=0N1|titi+1titi+1ti+12t(g˙(r)g˙(s))drdsdt|2)12Lp(Ω)+Cp2h(i=0N1|titi+1titi+1(ti+12ti+τih(g˙(s)g˙(r))dr+ti+12ti+τ¯ih(g˙(s)g˙(r))dr)dsdt|2)12Lp(Ω),where the first term on the right-hand side from Equation (Equation9) and the second term is due to Equation (Equation21). Let us now deal with the first term, the second term can be handled in the same way. Following a similar argument in (Equation16), we have that Cph(i=0N1|titi+1titi+1ti+12t(g˙(r)g˙(s))drdsdt|2)12Lp(Ω)Cph(i=0N1titi+1titi+1ti+12t(g˙(r)g˙(s))drdsdtLp(Ω;Rd)2)12Cp(i=0N1|titi+1ti+12t|g˙(r)g˙(s)|drds|2)12Cp(i=0N1h2q(titi+1(ti+12t|g˙(r)g˙(s)|dr)pds)2p)12Cp(i=0N1h4q(titi+1ti+12t|g˙(r)g˙(s)|pdrds)2p)12Cp(i=0N1h4q+2p+2(σ1)(titi+1ti+12t|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)2p)12=Cphσ(i=0N1h2q(titi+1ti+12t|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)2p)12,where we apply Hölder's inequality in Line 4 and 5. Similarly as in (Equation17), for p = 2 we have that Cphσ(i=0N1h2q(titi+1ti+12t|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)2p)12Cphσ+12gWσ,p.Applying discrete Hölder inequality for p>2 as in (Equation18), we have that Cphσ(i=0N1h2q(titi+1ti+12t|g˙(r)g˙(s)|p|rs|1+(σ1)pdrds)2p)12Cphσ+12Tp22pgWσ,p.Altogether we have achieved Bound (Equation13).

For a fixed integrand, the randomized quadrature rule (RTQ) improves the order of convergence by 12 through incorporating randomness compared to Theorem 3.1. One may also be interested in the almost sure convergence of RTQ. Indeed, the argument from [Citation8, Theorem 3.2] can be directly adapted here:

Theorem 3.3

Almost sure convergence

Assume that conditions from Theorem 3.2 are satisfied. Let (hm)mN(0,1) be an arbitrary sequence of step sizes with m=1hm<. Then, there exists a nonnegative random variable m0:ΩN{0} and a measurable set AF with P(A)=1 such that for all ωA and mm0(ω), then for every ϵ(0,12) there exist a nonnegative random variable m0ϵ:ΩN0 and a measurable set AϵF with P(Aϵ)=1 such that such that for all ωA and mm0(ω) we have (22) maxn{0,1,,Nhm}|In[g]RQhmτ,n[g](ω)|hm12+γϵ,(22) where Nhm:=Thm, i.e. the integer part of Thm.

Theorem 3.3 ensures that RTQ can achieve a slightly better order of pathwise convergence in the almost sure sense compared to CTQ when stepsize is adequately small.

4. Numerical experiments

In this section, we assessed the proposed method via different experiments. For simplicity, we fix T = 1.

4.1. Example 1

Consider the function: (23) gγ(t):=tγ,(23) where γ{54,32,74}, gγWσ,2(0,T), for all ϵ(1,12+γ) (Sobolev's inequality in [Citation1]). The curves of gγ with different values in γ can be found in Figure . The true solution was easily obtained as 1γ+1. The numerical approximations were calculated for both kinds of trapezoidal quadrature with step sizes h{2i:i=5,,10} and then compared to the true solution for errors. For RTQ, we computed errors in L2 norm via Monte Carlo method and also computed pathwise error, i.e.error from one realization.

Figure 1. Function values for gγ under different choices for γ.

Figure 1. Function values for gγ under different choices for γ.

The results of our simulations are shown in Figure  and Table . Across all different values of γ, RTQ gave the higher order of convergence compared to CTQ. When γ increased from 54 to 74, the order of convergence for RTQ increased eventually to a number very close to 2.5. Note that the order of convergence for CTQ are not beyond 2 for all γ values. All the performances were superior to theoretical order of convergences shown in Theorems 3.1 and 3.2. We also examined the computational efficiency of both methods (lower right in Figure ). Though incorporating randomness increased computational expense, RTQ quickly offsetted its cost with its higher accuracy.

Figure 2. Error plots for approximating I[gγ] via variants of trapezoidal rule under different choices for γ (upper left: γ=54; upper right: γ=32; lower left: γ=74) and time cost plot for γ=32 (lower right).

Figure 2. Error plots for approximating I[gγ] via variants of trapezoidal rule under different choices for γ (upper left: γ=54; upper right: γ=32; lower left: γ=74) and time cost plot for γ=32 (lower right).

Table 1. Order of convergences for simulating I[gγ].

4.2. Example 2

Consider the function: (24) gB(t):=0tB(s)ds,fort[0,T],(24) where B(s) is a realization of standard Brownian motion (BM) (cf. Section 3.1 in [Citation6]). It is well known that BC12ϵ for arbitrary small ϵ>0, therefore gBW32ϵ,p for p>1. Figure  illustrates how one BM path looks like and the curve of its gB.We are interested in approximating I[gB].

Figure 3. One realization of standard Brownian motion and function values for the corresponding gB.

Figure 3. One realization of standard Brownian motion and function values for the corresponding gB.

Due to the nature of BM, it is not easy to obtain the exact value of gB. To approximating terms gB(tn), one simply applies the Euler method, i.e. gB(tn)=0tnB(s)ds=i=0n1titi+1B(s)dshi=0n1B(ti).For CTQ, for a fixed stepsize h[0,1], we have that Qh[gB]=h2n=0N1(gB(tn)+gB(tn+1))=hn=0N1gB(tn)+h2n=0N1(gB(tn+1)gB(tn))=hn=0N1gB(tn)+h2gB(tN)=hn=1NgB(tn)h2gB(tN)h2n=0N1i=0nB(ti)h22i=0N1B(ti).For RTQ, define the corresponding i.i.d. uniform distributed sequence is {τjh}jN, we have a similar expression: (25) RQhτ,N[gB]=h2n=0N1(gB(tn+τnhh)+gB(tn+τ¯nhh))=hn=0N1gB(tn)+h2n=0N1(tntn+τnhhB(s)ds+tntn+τ¯nhhB(s)ds)hn=0N1gB(tn)+h2n=0N1(τnhh2(B(tn)+B(tn+τnhh))+τ¯nhh2(B(tn)+B(tn+τ¯nhh))),(25) where the last term can be further expanded as h2n=0N1(τnhh2(B(tn)+B(tn+τnhh))+τ¯nhh2(B(tn)+B(tn+τ¯nhh)))=h24n=0N1B(tn)+h24n=0N1(τnhB(tn+τnhh)+τ¯nhB(tn+τ¯nhh)).Note that to deduce the third line of Equation (Equation25), we make use of CTQ rather than the Euler method. The reason for this is that using the Euler method will result in the same expression as Qh[gB]. It is easy to see the difference between expressions for CTQ and RTQ lies in the last two terms of the equation above.

To compute the reference solution, we first sampled a BM path with a small stepsize href=214. Then, we generated an i.i.d. standard uniformly distributed sequence {τj}jN, and sampled B((j+τj)href), which is determined by property of Brownian bridge (cf. Section 3.1 in [Citation6]), i.e. B((j+τj)href)N(τ¯jB((j+1)href)+τjB(jhref),τjτ¯jhref),where N(μ,σ2) is normal distribution with mean μ and variance σ2, and τ¯j:=1τj for all j. The reference solution was thus computed via CTQ on grid points consisting of {jhref}jN as well as these intermediate {(j+τj)href}jN.

The reason for including randomness at this early stage is that this allows an easier sampling procedure for {τjh}jN on coarser grids of stepsize h. For instance, if h=2href and consider interval [t0,t0+h], then t0+τ0href and t0+href+τ1href are in the same interval. Thus τ0h can be determined from t0+τ0hh=1U0(0)(t0+τ0href)+1U0(1)(t0+href+τ1href),where 1() is the indicator function, U0U{0,1}, i.e. a discrete uniform distribution on the integers 0 and 1.

The numerical approximations were calculated for both trapezoidal quadratures with larger step sizes h{2i:i=5,,10} and then compared to the reference solution for errors. The results of our simulations are shown in Figure . RTQ gave the higher order of pathwise convergence compared to CTQ and gained a minor advantage in absolute error. Both the performances are consistent with the theoretical order of convergences shown in Theorems 3.1 and 3.3. We, in the meantime, examined the computational efficiency of both methods. Due to additional terms involved for RTQ in Equation (Equation25), its time cost roughly doubled that of CTQ at the same stepsize. In this case, unfortunately, the slight odds of RTQ in accuracy did not offset its cost.

Figure 4. Error plot (left) and time cost plot (right) for approximating I[gB] using CTQ and RTQ.

Figure 4. Error plot (left) and time cost plot (right) for approximating I[gB] using CTQ and RTQ.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by the Alan Turing Institute under the Engineering and Physical Sciences Research Council (EPSRC) grant EP/N510129/1 and by EPSRC though the project EP/S026347/1, titled ‘Unparameterized multi-modal data, high order signatures, and the mathematics of data science’.

References

  • R.A. Adams and J.J. Fournier, Sobolev Spaces, Elsevier, Amsterdam, 2003.
  • R.B. Ash and C.A. Doleans-Dade, Probability and Measure Theory, Academic Press, Burlington, MA, 2000.
  • D. Cruz-Uribe and C.J. Neugebauer, Sharp error bounds for the trapezoidal rule and Simpson's rule, J. Inequal. Pure Appl. Math. 3(4) (2002), pp. 1–222.
  • P.J. Davis and P. Rabinowitz, Methods of Numerical Integration, Courier Corporation, Mineola, NY, 2007.
  • M. Eisenmann and R. Kruse, Two quadrature rules for stochastic Ito-integrals with fractional Sobolev regularity, Commun. Math. Sci. 16(8) (2018), pp. 2125–2146.
  • P. Glasserman, Monte Carlo Methods in Financial Engineering, Springer Science & Business Media, New York, 2013.
  • E.T. Goodwin, The evaluation of integrals of the form ∫−∞∞f(x)exp⁡(−x)dx, Proc. Camb. Phil. Soc. 45 (1949), pp. 241–245.
  • R. Kruse and Y. Wu, Error analysis of randomized Runge–Kutta methods for differential equations with time-irregular coefficients, Comput. Methods Appl. Math. 17(3) (2017), pp. 479–498.
  • C. Schwartz, Numerical integration of analytic functions, J. Comput. Phys. 4(1) (1969), pp. 19–29.
  • F. Stenger, Integration formulae based on the trapezoidal formula, IMA J. Appl. Math. 12(1) (1973), pp. 103–114.