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

Stability of variable-step BDF2 and BDF3 methods

Zhaoyi Li  Hong-lin Liao Department of Mathematics, Nanjing University of Aeronautics and Astronautics, Nanjing 211106, P. R. China. Email: [email protected] author. ORCID 0000-0003-0777-6832. Department of Mathematics, Nanjing University of Aeronautics and Astronautics, Nanjing 211106, China; Key Laboratory of Mathematical Modelling and High Performance Computing of Air Vehicles (NUAA), MIIT, Nanjing 211106, China. Hong-lin Liao ([email protected] and [email protected]) is supported by a grant 12071216 from National Natural Science Foundation of China.
Abstract

We prove that the two-step backward differentiation formula (BDF2) method is stable on arbitrary time grids; while the variable-step BDF3 scheme is stable if almost all adjacent step ratios are less than 2.553. These results relax the severe mesh restrictions in the literature and provide a new understanding of variable-step BDF methods. Our main tools include the discrete orthogonal convolution kernels and an elliptic-type matrix norm.
Keywords: variable-step BDF formula, stability, step-ratio condition, discrete orthogonal convolution kernels
AMS subject classifications: 65M06, 65M12

1 Introduction

In this paper, we revisit the stability of variable-step backward differentiation formulas (BDF) for the following initial value problem

dvdt=f(t,v)for 0<tT.\displaystyle\frac{\,\mathrm{d}v}{\,\mathrm{d}t}=f(t,v)\quad\text{for $0<t\leq T$}.

Choose time levels 0=t0<t1<t2<<tN=T0=t_{0}<t_{1}<t_{2}<\cdots<t_{N}=T with the variable time-step τk:=tktk1\tau_{k}:=t_{k}-t_{k-1} for 1kN1\leq k\leq N, and the maximum step size τ:=max1kNτk\tau:=\max_{1\leq k\leq N}\tau_{k}. Put the adjacent time-step ratios

rk:=τkτk1for 2kNwithr1:=0.r_{k}:=\frac{\tau_{k}}{\tau_{k-1}}\quad\text{for $2\leq k\leq N$}\quad\text{with}\quad r_{1}:=0.

For any sequences {vn}n=0N\{v^{n}\}_{n=0}^{N}, denote τvn:=(vnvn1)/τn\partial_{\tau}v^{n}:=(v^{n}-v^{n-1})/\tau_{n}. Let Πn,kv\Pi_{n,k}v (k=1,2,3k=1,2,3) be the interpolating polynomial of a function vv over the nodes tnkt_{n-k}, \cdots, tn1t_{n-1} and tnt_{n}. To describe the variable coefficients, define the following three functions

d0(x,y):=\displaystyle d_{0}(x,y):= 1+2x1+x+xy1+y+xy,\displaystyle\,\frac{1+2x}{1+x}+\frac{xy}{1+y+xy}, (1.1)
d1(x,y):=\displaystyle d_{1}(x,y):= x1+xxy1+y+xyxy21+y+xy1+x1+y,\displaystyle\,-\frac{x}{1+x}-\frac{xy}{1+y+xy}-\frac{xy^{2}}{1+y+xy}\frac{1+x}{1+y}, (1.2)
d2(x,y):=\displaystyle d_{2}(x,y):= xy21+y+xy1+x1+y,for x,y0.\displaystyle\,\frac{xy^{2}}{1+y+xy}\frac{1+x}{1+y},\qquad\text{for $x,y\geq 0$.} (1.3)

Taking vn=v(tn)v^{n}=v(t_{n}), one has the BDF1 formula D1vn:=(Πn,1v)(tn)=d0(0,0)τvnD_{1}v^{n}:=\left(\Pi_{n,1}v\right)^{\prime}(t_{n})=d_{0}(0,0)\partial_{\tau}v^{n} for n1n\geq 1, and the variable-step BDF2 formula

D2vn:=(Πn,2v)(tn)=d0(rn,0)τvn+d1(rn,0)τvn1for n2.\displaystyle D_{2}v^{n}:=\left(\Pi_{n,2}v\right)^{\prime}(t_{n})=d_{0}(r_{n},0)\partial_{\tau}v^{n}+d_{1}(r_{n},0)\partial_{\tau}v^{n-1}\quad\text{for $n\geq 2$}. (1.4)

The variable-step BDF3 formula D3vn:=(Πn,3v)(tn)D_{3}v^{n}:=\left(\Pi_{n,3}v\right)^{\prime}(t_{n}) for n3n\geq 3,

D3vn=\displaystyle D_{3}v^{n}= d0(rn,rn1)τvn+d1(rn,rn1)τvn1+d2(rn,rn1)τvn2.\displaystyle\,d_{0}(r_{n},r_{n-1})\partial_{\tau}v^{n}+d_{1}(r_{n},r_{n-1})\partial_{\tau}v^{n-1}+d_{2}(r_{n},r_{n-1})\partial_{\tau}v^{n-2}. (1.5)

As seen, the formula (1.5) also represents a variable-order variable-step BDF formula from the first-order to third-order. The theoretical results on (1.5) are naturally valid for any combinations of the BDF-k formulas if the time-step ratios are allowed to be zero in the associated coefficients.

We consider the variable-step BDF-k\mathrm{k} time-stepping scheme for the ODE problem,

Dkvn=f(tn,vn)for knN.\displaystyle D_{\mathrm{k}}v^{n}=f(t_{n},v^{n})\quad\text{for $\mathrm{k}\leq n\leq N$.} (1.6)

Without losing the generality, assume that the starting values u1u^{1}, \cdots, uk1u^{\mathrm{k}-1} are available by some other high-order approaches, such as Runge-Kutta methods. For the stability analysis, assume further that the perturbed solution v¯n\bar{v}^{n} solves the equation Dkv¯n=f(tn,v¯n)+εnD_{\mathrm{k}}\bar{v}^{n}=f(t_{n},\bar{v}^{n})+\varepsilon^{n} with a bounded sequence {εn}\{\varepsilon^{n}\}. Then the difference v~i=v¯ivi\tilde{v}^{i}=\bar{v}^{i}-v^{i} solves

Dkv~i=f(ti,v¯i)f(ti,vi)+εifor kiN.\displaystyle D_{\mathrm{k}}\tilde{v}^{i}=f(t_{i},\bar{v}^{i})-f(t_{i},v^{i})+\varepsilon^{i}\quad\text{for $\mathrm{k}\leq i\leq N$.} (1.7)

These well-known stiff solvers have been tested to be practically valuable for differential-algebraic problems [5, 9, 10] and for the hyperbolic systems with multiscale relaxation [1]. Also, the variable-step versions are computationally efficient in capturing the multi-scale time behaviors [2, 4, 10, 11, 12, 13, 15] via adaptive time-stepping strategies. On the other hand, the stability and convergence analysis of variable-step BDF methods are difficult and remain incomplete, cf. [2, 4, 6, 7, 8, 9, 15], because they involve multiple degrees of freedom (independent time-step sizes).

One of the main defects in the existing theory is the severe mesh condition required for stability. For examples, the zero-stability of variable-step BDF2 method requires a step-ratio condition 0<rk1+22.4140<r_{k}\leq 1+\sqrt{2}\approx 2.414, see [6, 7]. By considering a specific time grid with constant step-ratio rk=rr_{k}=r, it was also shown in [6] that the zero-stability of variable-step BDF3 method requires the maximum step-ratio limit R3<(1+5)/21.618R_{3}<(1+\sqrt{5})/{2}\approx 1.618. By using an elliptic type norm, Calvo, Grande and Grigorieff [2] gave the zero-stability constraint rk1.476r_{k}\leq 1.476 in 1990. Twenty years ago, Guglielmi and Zennaro [8] applied a spectral radius approach to find the step-ratio condition rk1.501r_{k}\leq 1.501, which seems to be the best result for the variable-step BDF3 method.

We aim to relax the existing mesh conditions for the perturbed stability of variable-step BDF2 and BDF3 methods according to the following definition.

Definition 1 (Stability).

Assume that the nonlinear function f(t,v)f(t,v) is a Lipschitz-continuous function with the Lipschitz constant Lf>0L_{f}>0. The k\mathrm{k}-step time-stepping (1.6) with unequal time-steps is stable if there exists a real number τ0\tau_{0} and a fixed constant CC (independent of the step sizes τn\tau_{n}) such that the numerical solution of (1.7) fulfills

maxknN|v~i|C(i=0k1|v~i|+τi=1k1|τv~i|+maxkiN|εi|)for all τnτ0.\displaystyle\max_{\mathrm{k}\leq n\leq N}\big{|}\tilde{v}^{i}\big{|}\leq C\left(\sum_{i=0}^{\mathrm{k}-1}\big{|}\tilde{v}^{i}\big{|}+\tau\sum_{i=1}^{\mathrm{k}-1}\big{|}\partial_{\tau}\tilde{v}^{i}\big{|}+\max_{\mathrm{k}\leq i\leq N}\big{|}\varepsilon^{i}\big{|}\right)\quad\text{for all $\tau_{n}\leq\tau_{0}$.}

This definition is practically reasonable in the sense that one always uses some high-order scheme to start the k\mathrm{k}-step time-stepping (1.6). In the uniform time-step case, it reduces into the classical zero-stability, cf. [3, Definition 2.2],

maxknN|v~i|C(i=0k1|v~i|+maxkiN|εi|)for all τnτ0.\displaystyle\max_{\mathrm{k}\leq n\leq N}\big{|}\tilde{v}^{i}\big{|}\leq C\left(\sum_{i=0}^{\mathrm{k}-1}\big{|}\tilde{v}^{i}\big{|}+\max_{\mathrm{k}\leq i\leq N}\big{|}\varepsilon^{i}\big{|}\right)\quad\text{for all $\tau_{n}\leq\tau_{0}$.}

Definition 1 would be as useful as the original zero-stability for the numerical analysis of variable-step, multi-step methods. Actually, it was proven in [14] that the step-ratio condition rk<3.561r_{k}<3.561 is sufficient to the stability of variable-step BDF2 method for linear parabolic problem. In a recent work [12], this step-ratio condition was slightly improved to a new one, rk<4.864r_{k}<4.864. These step-ratio conditions are weaker than the classical zero-stability restriction rk<2.414r_{k}<2.414, but both of them are are sufficient for the stability of BDF2 method in the sense of Definition 1.

By developing a new framework with discrete orthogonal convolution (DOC) kernels, we apply Definition 1 to examine the ratio-step conditions for the stability of BDF2 and BDF3 methods in solving the ODE problems. The main contribution is two-fold:

  • Theorem 2.1 shows that the BDF2 method is stable on arbitrary time grids.

  • Theorem 3.1 shows that the variable-step BDF3 method is stable if the adjacent step ratios rk<R3,r_{k}<R_{3}, where R32.553R_{3}\approx 2.553 is the unique positive root of

    R36+R354R348R3310R326R32=0.R_{3}^{6}+R_{3}^{5}-4R_{3}^{4}-8R_{3}^{3}-10R_{3}^{2}-6R_{3}-2=0.

We start from a novel viewpoint by writing the BDF-k\mathrm{k} (k=2,3\mathrm{k}=2,3) formula as a convolution summation of local difference quotients (this viewpoint is quite different from that in previous works [12, 14])

Dkvn:=j=1ndnj(k,n)τvjfor nk,\displaystyle D_{\mathrm{k}}v^{n}:=\sum_{j=1}^{n}d^{(\mathrm{k},n)}_{n-j}\partial_{\tau}v^{j}\quad\text{for $n\geq\mathrm{k}$,} (1.8)

where the first superscript index k\mathrm{k} in the discrete kernels dnj(k,n)d^{(\mathrm{k},n)}_{n-j} denotes the order of method. From the BDF2 formula (1.4) with n2n\geq 2, we have

dj(2,n):=dj(rn,0)for j=0,1,anddj(2,n):=0for nj+13.\displaystyle d^{(2,n)}_{j}:=d_{j}(r_{n},0)\;\;\text{for $j=0,1$,}\quad\text{and}\quad d^{(2,n)}_{j}:=0\quad\text{for $n\geq j+1\geq 3$}. (1.9)

Similarly, from the BDF3 formula (1.5) with n3n\geq 3, we have

dj(3,n):=dj(rn,rn1)for j=0,1,2,anddj(3,n):=0for nj+14.\displaystyle d^{(3,n)}_{j}:=d_{j}(r_{n},r_{n-1})\;\;\text{for $j=0,1,2$,}\quad\text{and}\quad d^{(3,n)}_{j}:=0\quad\text{for $n\geq j+1\geq 4$}. (1.10)

Here and hereafter, assume that the summation k=ij\sum_{k=i}^{j}\cdot to be zero and the product k=ij\prod_{k=i}^{j}\cdot to be one if the index i>ji>j. As for the BDF-k\mathrm{k} kernels dnj(k,n)d^{(\mathrm{k},n)}_{n-j} with any fixed indexes nn, we recall a class of discrete orthogonal convolution (DOC) kernels {ϑnj(k,n)}j=kn\big{\{}\vartheta_{n-j}^{(\mathrm{k},n)}\big{\}}_{j=\mathrm{k}}^{n} by a recursive procedure, also see [14, 12],

ϑ0(k,n):=1d0(k,n)andϑnj(k,n):=1d0(k,j)i=j+1nϑni(k,n)dij(k,i)for kjn1.\displaystyle\vartheta_{0}^{(\mathrm{k},n)}:=\frac{1}{d^{(\mathrm{k},n)}_{0}}\quad\text{and}\quad\vartheta_{n-j}^{(\mathrm{k},n)}:=-\frac{1}{d^{(\mathrm{k},j)}_{0}}\sum_{i=j+1}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}d^{(\mathrm{k},i)}_{i-j}\quad\text{for $\mathrm{k}\leq j\leq n-1$.} (1.11)

Obviously, the DOC kernels ϑnj(k,n)\vartheta_{n-j}^{(\mathrm{k},n)} satisfy the following discrete orthogonality identity

i=jnϑni(k,n)dij(k,i)δnjfor any kjn,\displaystyle\sum_{i=j}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}d^{(\mathrm{k},i)}_{i-j}\equiv\delta_{nj}\quad\text{for any $\mathrm{k}\leq j\leq n$,} (1.12)

where δnj\delta_{nj} is the Kronecker delta symbol with δnj=0\delta_{nj}=0 if jnj\neq n. Furthermore, with the identity matrix Im×mI_{m\times m} (m:=nk+1m:=n-\mathrm{k}+1), the above discrete orthogonality identity (1.12) also implies

𝚯k𝐃k=Im×m,\mathbf{\Theta}_{\mathrm{k}}\mathbf{D}_{\mathrm{k}}=I_{m\times m},

where the two m×mm\times m matrices 𝐃k\mathbf{D}_{\mathrm{k}} and 𝚯k\mathbf{\Theta}_{\mathrm{k}} are defined by

𝐃k:=(d0(k,k)d1(k,k+1)d0(k,k+1)dnk(k,n)dnk1(k,n)d0(k,n))and𝚯k:=(ϑ0(k,k)ϑ1(k,k+1)ϑ0(k,k+1)ϑnk(k,n)ϑnk1(k,n)ϑ0(k,n)).\mathbf{D}_{\mathrm{k}}:=\left(\begin{array}[]{cccc}d_{0}^{(\mathrm{k},\mathrm{k})}&&&\\ d_{1}^{(\mathrm{k},\mathrm{k}+1)}&d_{0}^{(\mathrm{k},\mathrm{k}+1)}&&\\ \vdots&\vdots&\ddots&\\ d_{n-\mathrm{k}}^{(\mathrm{k},n)}&d_{n-\mathrm{k}-1}^{(\mathrm{k},n)}&\cdots&d_{0}^{(\mathrm{k},n)}\\ \end{array}\right)\;\;\text{and}\;\;\mathbf{\Theta}_{\mathrm{k}}:=\left(\begin{array}[]{cccc}\vartheta_{0}^{(\mathrm{k},\mathrm{k})}&&&\\ \vartheta_{1}^{(\mathrm{k},\mathrm{k}+1)}&\vartheta_{0}^{(\mathrm{k},\mathrm{k}+1)}&&\\ \vdots&\vdots&\ddots&\\ \vartheta_{n-\mathrm{k}}^{(\mathrm{k},n)}&\vartheta_{n-\mathrm{k}-1}^{(\mathrm{k},n)}&\cdots&\vartheta_{0}^{(\mathrm{k},n)}\\ \end{array}\right).

Obviously, one has

𝐃k𝚯k=Im×m,\mathbf{D}_{\mathrm{k}}\mathbf{\Theta}_{\mathrm{k}}=I_{m\times m},

which implies the following mutual orthogonality identity

i=jndni(k,n)ϑij(k,i)δnjfor any kjn.\displaystyle\sum_{i=j}^{n}d_{n-i}^{(\mathrm{k},n)}\vartheta^{(\mathrm{k},i)}_{i-j}\equiv\delta_{nj}\quad\text{for any $\mathrm{k}\leq j\leq n$.} (1.13)

We will use this identity (1.13) to study the decaying property of DOC kernels ϑnj(k,n)\vartheta_{n-j}^{(\mathrm{k},n)}.

We will derive an equivalent convolution form of the difference equation (1.7). By exchanging the summation order and using (1.12), one has

i=knϑni(k,n)Dkv~i=\displaystyle\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}D_{\mathrm{k}}\tilde{v}^{i}= i=knϑni(k,n)j=1k1dij(k,i)τvj+i=knϑni(k,n)j=kidij(k,i)τvj\displaystyle\,\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}\sum_{j=1}^{\mathrm{k}-1}d^{(\mathrm{k},i)}_{i-j}\partial_{\tau}v^{j}+\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}\sum_{j=\mathrm{k}}^{i}d^{(\mathrm{k},i)}_{i-j}\partial_{\tau}v^{j}
=\displaystyle= j=1k1τvji=knϑni(k,n)dij(k,i)+j=knτvji=jnϑni(k,n)dij(k,i)\displaystyle\,\sum_{j=1}^{\mathrm{k}-1}\partial_{\tau}v^{j}\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}d^{(\mathrm{k},i)}_{i-j}+\sum_{j=\mathrm{k}}^{n}\partial_{\tau}v^{j}\sum_{i=j}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}d^{(\mathrm{k},i)}_{i-j}
=\displaystyle= kn[v]+τvnfor nk,\displaystyle\,\mathcal{I}_{\mathrm{k}}^{n}[v]+\partial_{\tau}v^{n}\qquad\text{for $n\geq\mathrm{k}$,} (1.14)

where kn[v]\mathcal{I}_{\mathrm{k}}^{n}[v] represents the starting effect on the numerical solution at tnt_{n},

kn[v]:=\displaystyle\mathcal{I}_{\mathrm{k}}^{n}[v]:= j=1k1τvji=knϑni(k,n)dij(k,i)for nk.\displaystyle\,\sum_{j=1}^{\mathrm{k}-1}\partial_{\tau}v^{j}\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}d^{(\mathrm{k},i)}_{i-j}\qquad\text{for $n\geq\mathrm{k}$.} (1.15)

Multiplying both sides of the difference equation (1.7) by the DOC kernels ϑni(k,n)\vartheta_{n-i}^{(\mathrm{k},n)}, and summing ii from k\mathrm{k} to nn, we apply (1) to get

τv~n=kn[v~]+i=knϑni(k,n)[f(ti,v¯i)f(ti,vi)+εi]for knN.\displaystyle\partial_{\tau}\tilde{v}^{n}=-\mathcal{I}_{\mathrm{k}}^{n}[\tilde{v}]+\sum_{i=\mathrm{k}}^{n}\vartheta_{n-i}^{(\mathrm{k},n)}\left[f(t_{i},\bar{v}^{i})-f(t_{i},v^{i})+\varepsilon^{i}\right]\quad\text{for $\mathrm{k}\leq n\leq N$.} (1.16)

In next two sections, we will present the stability analysis for the variable-step BDF2 and BDF3 methods, respectively, via this discrete convolution form (1.16) of the perturbed equation. Numerical experiments on the graded and random time meshes are included in Section 4 to support our theory. Some concluding remarks end this article.

2 Unconditional stability of BDF2 method

For the BDF2 method, the associated DOC kernels are positive and decay exponentially.

Lemma 2.1.

The DOC kernels ϑnj(2,n)\vartheta_{n-j}^{(2,n)} in (1.11) have an explicit formula

ϑnj(2,n)=1d0(2,j)i=j+1nri1+2ri>0 for 2jn,\displaystyle\vartheta_{n-j}^{(2,n)}=\frac{1}{d^{(2,j)}_{0}}\prod_{i=j+1}^{n}\frac{r_{i}}{1+2r_{i}}>0\quad\text{ for $2\leq j\leq n$},

and satisfy

j=2nϑnj(2,n)=1i=2nri1+2ri<1for n2.\displaystyle\sum_{j=2}^{n}\vartheta_{n-j}^{(2,n)}=1-\prod_{i=2}^{n}\frac{r_{i}}{1+2r_{i}}<1\quad\text{for $n\geq 2$.}
Proof.

For any fixed jj, by taking n=jn=j in the identity (1.13), one has ϑ0(2,j)=1/d0(2,j)\vartheta_{0}^{(2,j)}=1/d^{(2,j)}_{0} for j2j\geq 2. According to the definition (1.9), one has dni(2,n)=0d^{(2,n)}_{n-i}=0 for ni+2n\geq i+2. The identity (1.13) gives

ϑmj(2,m)=d1(2,m)d0(2,m)ϑm1j(2,m1)=d1(rm,0)d0(rm,0)ϑm1j(2,m1)=rm1+2rmϑm1j(2,m1)\displaystyle\vartheta_{m-j}^{(2,m)}=-\frac{d^{(2,{m})}_{1}}{d^{(2,m)}_{0}}\vartheta_{m-1-j}^{(2,m-1)}=-\frac{d_{1}(r_{m},0)}{d_{0}(r_{m},0)}\vartheta_{m-1-j}^{(2,m-1)}=\frac{r_{m}}{1+2r_{m}}\vartheta_{m-1-j}^{(2,m-1)}

for mj+12m\geq j+1\geq 2. It yields immediately

ϑnj(2,n)=\displaystyle\vartheta_{n-j}^{(2,n)}= ϑn2j(2,n2)i=n1nri1+2ri==ϑ0(2,j)i=j+1nri1+2ri for 2jn.\displaystyle\,\vartheta_{n-2-j}^{(2,n-2)}\prod_{i=n-1}^{n}\frac{r_{i}}{1+2r_{i}}=\cdots=\vartheta_{0}^{(2,j)}\prod_{i=j+1}^{n}\frac{r_{i}}{1+2r_{i}}\quad\text{ for $2\leq j\leq n$}.

It leads to the first result. By taking vn=tnv^{n}=t_{n} in (1.8), one has τtn=1\partial_{\tau}t_{n}=1 and D2tn=1D_{2}t_{n}=1 for n2n\geq 2. Thus applying the equality (1) with k=2\mathrm{k}=2 yields

j=2nϑnj(2,n)=\displaystyle\sum_{j=2}^{n}\vartheta_{n-j}^{(2,n)}= j=2nϑnj(2,n)D2tj=(τt1)i=2nϑni(2,n)di1(2,i)+τtn\displaystyle\,\sum_{j=2}^{n}\vartheta_{n-j}^{(2,n)}D_{2}t_{j}=(\partial_{\tau}t_{1})\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}d^{(2,i)}_{i-1}+\partial_{\tau}t_{n}
=\displaystyle=  1+ϑn2(2,n)d1(2,2)=1i=2nri1+2ri<1for n1,\displaystyle\,1+\vartheta_{n-2}^{(2,n)}d^{(2,2)}_{1}=1-\prod_{i=2}^{n}\frac{r_{i}}{1+2r_{i}}<1\quad\text{for $n\geq 1$,} (2.1)

where the explicit formula of ϑn2(2,n)\vartheta_{n-2}^{(2,n)} and the definition (1.9) were used. It completes the proof. ∎

Theorem 2.1.

If τ1/(4Lf)\tau\leq 1/(4L_{f}), the BDF2 solution of (1.7) satisfies

|v~n|2exp(4Lftn1)(|v~1|+2τ|τv~1|+2tnmax2in|εi|)for 2nN.\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq 2\exp(4L_{f}t_{n-1})\Big{(}\big{|}\tilde{v}^{1}\big{|}+2\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+2t_{n}\max_{2\leq i\leq n}\big{|}\varepsilon^{i}\big{|}\Big{)}\quad\text{for $2\leq n\leq N$.}

Thus the variable-step BDF2 scheme is stable on arbitrary time meshes.

Proof.

Take k=2\mathrm{k}=2. Multiplying both sides of (1.16) with 2τnv~n2\tau_{n}\tilde{v}^{n}, one applies Lemma 2.1 (the positivity and boundedness of DOC kernels) to obtain that

|v~n|2\displaystyle\big{|}\tilde{v}^{n}\big{|}^{2}- |v~n1|2+|τnτv~n|2=2τnv~n2n[v~]+2τnv~ni=2nϑni(2,n)[f(ti,v¯i)f(ti,vi)+εi]\displaystyle\,\big{|}\tilde{v}^{n-1}\big{|}^{2}+\big{|}\tau_{n}\partial_{\tau}\tilde{v}^{n}\big{|}^{2}=-2\tau_{n}\tilde{v}^{n}\mathcal{I}_{2}^{n}[\tilde{v}]+2\tau_{n}\tilde{v}^{n}\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}\left[f(t_{i},\bar{v}^{i})-f(t_{i},v^{i})+\varepsilon^{i}\right]
\displaystyle\leq  2τn|v~n||2n[v~]|+2τn|v~n|i=2nϑni(2,n)|f(ti,v¯i)f(ti,vi)|+2τn|v~n|max1in|εi|i=2nϑni(2,n)\displaystyle\,2\tau_{n}\big{|}\tilde{v}^{n}\big{|}\big{|}\mathcal{I}_{2}^{n}[\tilde{v}]\big{|}+2\tau_{n}\big{|}\tilde{v}^{n}\big{|}\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}\big{|}f(t_{i},\bar{v}^{i})-f(t_{i},v^{i})\big{|}+2\tau_{n}\big{|}\tilde{v}^{n}\big{|}\max_{1\leq i\leq n}\big{|}\varepsilon^{i}\big{|}\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}
\displaystyle\leq  2τn|v~n||2n[v~]|+2Lfτn|v~n|i=2nϑni(2,n)|v~i|+2|v~n|τnmax1in|εi|,\displaystyle\,2\tau_{n}\big{|}\tilde{v}^{n}\big{|}\big{|}\mathcal{I}_{2}^{n}[\tilde{v}]\big{|}+2L_{f}\tau_{n}\big{|}\tilde{v}^{n}\big{|}\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}\big{|}\tilde{v}^{i}\big{|}+2\big{|}\tilde{v}^{n}\big{|}\tau_{n}\max_{1\leq i\leq n}\big{|}\varepsilon^{i}\big{|},

where the Lipschitz continuous property of the nonlinear function ff has been used. By dropping the nonnegative term at the left side and summing nn from n=2n=2 to mm, we have

|v~m|2\displaystyle\big{|}\tilde{v}^{m}\big{|}^{2}\leq |v~1|2+2j=2mτj|v~j||2j[v~]|+2Lfj=2mτj|v~j|i=2jϑji(2,j)|v~i|+2j=2mτj|v~j|max1ij|εi|.\displaystyle\,\big{|}\tilde{v}^{1}\big{|}^{2}+2\sum_{j=2}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}\big{|}\mathcal{I}_{2}^{j}[\tilde{v}]\big{|}+2L_{f}\sum_{j=2}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}\sum_{i=2}^{j}\vartheta_{j-i}^{(2,j)}\big{|}\tilde{v}^{i}\big{|}+2\sum_{j=2}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}\max_{1\leq i\leq j}\big{|}\varepsilon^{i}\big{|}.

Let m0m_{0} (1m0m1\leq m_{0}\leq m) be an integer such that |v~m0|=max1km|v~k|\big{|}\tilde{v}^{m_{0}}\big{|}=\max_{1\leq k\leq m}\big{|}\tilde{v}^{k}\big{|}. Now we take m:=m0m:=m_{0} in the above inequality and get

|v~m0|2\displaystyle\big{|}\tilde{v}^{m_{0}}\big{|}^{2}\leq |v~1||v~m0|+2|v~m0|j=2m0τj|2j[v~]|+2Lf|v~m0|j=2m0τj|v~j|+2|v~m0|max2im0|εi|j=2m0τj.\displaystyle\,\big{|}\tilde{v}^{1}\big{|}\big{|}\tilde{v}^{m_{0}}\big{|}+2\big{|}\tilde{v}^{m_{0}}\big{|}\sum_{j=2}^{m_{0}}\tau_{j}\big{|}\mathcal{I}_{2}^{j}[\tilde{v}]\big{|}+2L_{f}\big{|}\tilde{v}^{m_{0}}\big{|}\sum_{j=2}^{m_{0}}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+2\big{|}\tilde{v}^{m_{0}}\big{|}\max_{2\leq i\leq m_{0}}\big{|}\varepsilon^{i}\big{|}\sum_{j=2}^{m_{0}}\tau_{j}.

It leads to

|v~m||v~m0||v~1|+2j=2mτj|2j[v~]|+2Lfj=2mτj|v~j|+2tmmax2im|εi|for 2mN.\displaystyle\big{|}\tilde{v}^{m}\big{|}\leq\big{|}\tilde{v}^{m_{0}}\big{|}\leq\big{|}\tilde{v}^{1}\big{|}+2\sum_{j=2}^{m}\tau_{j}\big{|}\mathcal{I}_{2}^{j}[\tilde{v}]\big{|}+2L_{f}\sum_{j=2}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+2t_{m}\max_{2\leq i\leq m}\big{|}\varepsilon^{i}\big{|}\quad\text{for $2\leq m\leq N$.} (2.2)

By using Lemma 2.1, the starting error 2n[v~]\mathcal{I}_{2}^{n}[\tilde{v}] in (1.15) reads

2n[v~]=τv~1i=2nϑni(2,n)di1(2,i)=ϑn2(2,n)d1(2,2)(τv~1)=(τv~1)i=2nri1+2rifor n2\displaystyle\mathcal{I}_{2}^{n}[\tilde{v}]=\partial_{\tau}\tilde{v}^{1}\sum_{i=2}^{n}\vartheta_{n-i}^{(2,n)}d^{(2,i)}_{i-1}=\vartheta_{n-2}^{(2,n)}d^{(2,2)}_{1}(\partial_{\tau}\tilde{v}^{1})=-(\partial_{\tau}\tilde{v}^{1})\prod_{i=2}^{n}\frac{r_{i}}{1+2r_{i}}\quad\text{for $n\geq 2$}

such that

j=2mτj|2j[v~]|τ|τv~1|j=2m12j1τ|τv~1|for m2.\displaystyle\sum_{j=2}^{m}\tau_{j}\big{|}\mathcal{I}_{2}^{j}[\tilde{v}]\big{|}\leq\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}\sum_{j=2}^{m}\frac{1}{2^{j-1}}\leq\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}\quad\text{for $m\geq 2$.}

It follows from (2.2) that

|v~n||v~1|+2τ|τv~1|+2Lfj=2nτj|v~j|+2tnmax2in|εi|for 2nN.\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq\big{|}\tilde{v}^{1}\big{|}+2\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+2L_{f}\sum_{j=2}^{n}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+2t_{n}\max_{2\leq i\leq n}\big{|}\varepsilon^{i}\big{|}\quad\text{for $2\leq n\leq N$.}

Assuming that τ1/(4Lf)\tau\leq 1/(4L_{f}), one has

|v~n|2|v~1|+4τ|τv~1|+4Lfj=2n1τj|v~j|+4tnmax2in|εi|for 2nN.\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq 2\big{|}\tilde{v}^{1}\big{|}+4\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+4L_{f}\sum_{j=2}^{n-1}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+4t_{n}\max_{2\leq i\leq n}\big{|}\varepsilon^{i}\big{|}\quad\text{for $2\leq n\leq N$.}

The standard discrete Grönwall inequality, e.g. [14, Lemma 3.1], completes the proof. ∎

3 Stability analysis of BDF3 method

3.1 Decaying of DOC kernels

Note that, the variable-step BDF3 method (1.8) of k=3\mathrm{k}=3 is exact for the linear polynomial v=tv=t. Taking vn=tnv^{n}=t_{n} in (1.8), one can find that D3tn=1D_{3}t_{n}=1 for n3n\geq 3. As done in (2), we can apply the discrete equality (1) with (1.15) to get

j=3nϑnj(3,n)=\displaystyle\sum_{j=3}^{n}\vartheta_{n-j}^{(3,n)}= j=12i=3nϑni(3,n)dij(3,i)+1for n3.\displaystyle\,\sum_{j=1}^{2}\sum_{i=3}^{n}\vartheta_{n-i}^{(3,n)}d^{(3,i)}_{i-j}+1\quad\text{for $n\geq 3$.}

However, it does not provide enough information for the subsequent stability analysis because no explicit formulas of DOC kernels ϑnj(3,n)\vartheta_{n-j}^{(3,n)} are available. Furthermore, the DOC kernels are not always positive, see Remark 4 below, we turn to bound j=3n|ϑnj(3,n)|\sum_{j=3}^{n}\big{|}\vartheta_{n-j}^{(3,n)}\big{|} for any time-levels tnt_{n} by imposing certain step-ratio constraint.

To do so, introduce the modified DOC kernels

ϑ^ij(3,i):=ϑij(3,i)d0(3,j)for ij3.\displaystyle\widehat{\vartheta}_{i-j}^{(3,i)}:=\vartheta_{i-j}^{(3,i)}d^{(3,j)}_{0}\quad\text{for $i\geq j\geq 3$.} (3.1)

The discrete identity (1.13) gives

i=jndni(3,n)ϑ^ij(3,i)δnjd0(3,j)for nj3.\displaystyle\sum_{i=j}^{n}d_{n-i}^{(3,n)}\widehat{\vartheta}^{(3,i)}_{i-j}\equiv\delta_{nj}d^{(3,j)}_{0}\quad\text{for $n\geq j\geq 3$.} (3.2)

For any fixed index j3j\geq 3, by taking n=jn=j and n=j+1n=j+1 in the identity (3.2), respectively, one can derive that

ϑ^0(3,j)=1andϑ^1(3,j+1)=d1(3,j+1)d0(3,j+1)for j3.\displaystyle\widehat{\vartheta}^{(3,j)}_{0}=1\quad\text{and}\quad\widehat{\vartheta}^{(3,j+1)}_{1}=-\frac{d_{1}^{(3,j+1)}}{d_{0}^{(3,j+1)}}\quad\text{for $j\geq 3$.} (3.3)

According to the definition (1.10), one has dni(3,n)=0d^{(3,n)}_{n-i}=0 for ni+3n\geq i+3. The identity (1.13) gives

ϑ^mj(3,m)+d1(3,m)d0(3,m)ϑ^m1j(3,m1)+d2(3,m)d0(3,m)ϑ^m2j(3,m2)=0for mj+25,\displaystyle\widehat{\vartheta}_{m-j}^{(3,m)}+\frac{d^{(3,{m})}_{1}}{d^{(3,m)}_{0}}\widehat{\vartheta}_{m-1-j}^{(3,m-1)}+\frac{d^{(3,{m})}_{2}}{d^{(3,m)}_{0}}\widehat{\vartheta}_{m-2-j}^{(3,m-2)}=0\quad\text{for $m\geq j+2\geq 5$,}

or the difference equation

ϑ^mj(3,m)αmϑ^m1j(3,m1)+βmϑ^m2j(3,m2)=0for mj+25,\displaystyle\widehat{\vartheta}_{m-j}^{(3,m)}-\alpha_{m}\widehat{\vartheta}_{m-1-j}^{(3,m-1)}+\beta_{m}\widehat{\vartheta}_{m-2-j}^{(3,m-2)}=0\quad\text{for $m\geq j+2\geq 5$,} (3.4)

where, by using (1.1)-(1.3) and (A.2)-(A.3), the coefficients αm\alpha_{m} and βm\beta_{m} are defined by

αm:=d1(3,m)d0(3,m)=α(rm,rm1)andβm:=d2(3,m)d0(3,m)=β(rm,rm1)for m3.\displaystyle\alpha_{m}:=-\frac{d^{(3,{m})}_{1}}{d^{(3,m)}_{0}}=\alpha(r_{m},r_{m-1})\quad\text{and}\quad\beta_{m}:=\frac{d^{(3,{m})}_{2}}{d^{(3,m)}_{0}}=\beta(r_{m},r_{m-1})\quad\text{for $m\geq 3$.} (3.5)

To evaluate the asymptotic behaviors of the DOC kernels ϑ^mj(3,m)\widehat{\vartheta}_{m-j}^{(3,m)}, introduce the following vector and companion matrix

um=(ϑ^mj(3,m)ϑ^m1j(3,m1))andAm:=(αmβm10).\displaystyle\vec{u}_{m}=\left(\begin{array}[]{c}\widehat{\vartheta}_{m-j}^{(3,m)}\\[8.61108pt] \widehat{\vartheta}_{m-1-j}^{(3,m-1)}\\ \end{array}\right)\quad\text{and}\quad A_{m}:=\left(\begin{array}[]{cc}\alpha_{m}&-\beta_{m}\\[8.61108pt] 1&0\\ \end{array}\right). (3.10)

For any fixed index j1j\geq 1, the difference equation (3.4) reads

um=Amum1for mj+25,\displaystyle\vec{u}_{m}=A_{m}\vec{u}_{m-1}\quad\text{for $m\geq j+2\geq 5$,}

or

un=Anun1=AnAn1un2==i=j+2nAiuj+1for nj+25.\displaystyle\vec{u}_{n}=A_{n}\vec{u}_{n-1}=A_{n}A_{n-1}\vec{u}_{n-2}=\cdots=\prod_{i=j+2}^{n}A_{i}\vec{u}_{j+1}\quad\text{for $n\geq j+2\geq 5$.} (3.11)

The associated initial vector uj+1=(αj+1,1)T\vec{u}_{j+1}=(\alpha_{j+1},1)^{T} according to (3.3).

If all of step-ratios 0<rk<R^33.44050<r_{k}<\hat{R}_{3}\approx 3.4405 (k2k\geq 2), Lemmas A.1-A.2 show that

βi<αi<1+βi<2for i3.\displaystyle\beta_{i}<\alpha_{i}<1+\beta_{i}<2\quad\text{for $i\geq 3$.} (3.12)

They imply that the roots of λ2αiλ+βi=0\lambda^{2}-\alpha_{i}\lambda+\beta_{i}=0 satisfy |λ|<1\left|\lambda\right|<1. Thus the eigenvalues of AiA_{i} have the module less than 1 and the spectral radius ρ(Ai)<1\rho(A_{i})<1 for any i3i\geq 3. In general, it is not sufficient to ensure the global decaying of DOC kernels ϑ^nj(3,n)\widehat{\vartheta}_{n-j}^{(3,n)}.

It needs to examine the joint companion matrix i=j+2nAi\prod_{i=j+2}^{n}A_{i} for nj+25n\geq j+2\geq 5. Let \left\|\cdot\right\|_{\infty} be the maximum norm on the space 2\mathbb{C}^{2} and let the same symbol \left\|\cdot\right\|_{\infty} denote also the corresponding induced matrix norm. We adopt the technique of Calvo et al [2] by considering a complex constant μ\mu with Im(μ)0\mathrm{Im}(\mu)\neq 0 and the following elliptic type norm

AiH,:=H1AiHwithH:=(μμ¯11).\displaystyle\left\|A_{i}\right\|_{H,\infty}:=\left\|H^{-1}A_{i}H\right\|_{\infty}\quad\text{with}\quad H:=\left(\begin{array}[]{cc}\mu&\bar{\mu}\\[8.61108pt] 1&1\\ \end{array}\right). (3.15)

Then one can bound the DOC kernels via

unHi=j+2nAiH,H1uj+1for nj+25.\displaystyle\big{\|}\vec{u}_{n}\big{\|}_{\infty}\leq\big{\|}H\big{\|}_{\infty}\prod_{i=j+2}^{n}\big{\|}A_{i}\big{\|}_{H,\infty}\big{\|}H^{-1}\big{\|}_{\infty}\big{\|}\vec{u}_{j+1}\big{\|}_{\infty}\quad\text{for $n\geq j+2\geq 5$.} (3.16)

To process the analysis, it is to determine a fixed parameter μ\mu. We will choose a proper constant R~3<R^33.4405\tilde{R}_{3}<\hat{R}_{3}\approx 3.4405 to ensure that

AiH,<1for i3,\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}<1\quad\text{for $i\geq 3$,} (3.17)

if all of step-ratios rk<R~3r_{k}<\tilde{R}_{3} (k2k\geq 2). It is easy to derive that

AiH,=|μ2αiμ+βi|+||μ|2αiμ+βi||μμ¯|for i3.\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}=\frac{\left|\mu^{2}-\alpha_{i}\mu+\beta_{i}\right|+\big{|}\left|\mu\right|^{2}-\alpha_{i}\mu+\beta_{i}\big{|}}{\left|\mu-\bar{\mu}\right|}\quad\text{for $i\geq 3$.} (3.18)

Noticing that the following fact

||μ|2αiμ+βi|2|μ2αiμ+βi|2=βi|μμ¯|2,\displaystyle\big{|}\left|\mu\right|^{2}-\alpha_{i}\mu+\beta_{i}\big{|}^{2}-\big{|}\mu^{2}-\alpha_{i}\mu+\beta_{i}\big{|}^{2}=\beta_{i}\big{|}\mu-\bar{\mu}\big{|}^{2},

one has

AiH,=|μ2αiμ+βi||μμ¯|+βi+|μ2αiμ+βi|2|μμ¯|2for i3.\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}=\frac{\left|\mu^{2}-\alpha_{i}\mu+\beta_{i}\right|}{\left|\mu-\bar{\mu}\right|}+\sqrt{\beta_{i}+\frac{\left|\mu^{2}-\alpha_{i}\mu+\beta_{i}\right|^{2}}{\left|\mu-\bar{\mu}\right|^{2}}}\quad\text{for $i\geq 3$.} (3.19)

Then the requirement (3.17) is equivalent to

|μ2αiμ+βi||μμ¯|<12(1βi)for i3.\displaystyle\frac{\big{|}\mu^{2}-\alpha_{i}\mu+\beta_{i}\big{|}}{\left|\mu-\bar{\mu}\right|}<\frac{1}{2}(1-\beta_{i})\quad\text{for $i\geq 3$}. (3.20)

Then, taking into account the coefficient relationship (3.12), we can find that the condition (3.20) can be written equivalently in the following form, also see [2, Eq. (14)],

|μ12αiı2(1+βi)2αi2|<12(1βi)for i3.\displaystyle\left|\mu-\frac{1}{2}\alpha_{i}-\frac{\imath}{2}\sqrt{(1+\beta_{i})^{2}-\alpha_{i}^{2}}\right|<\frac{1}{2}(1-\beta_{i})\quad\text{for $i\geq 3$.} (3.21)

Note that, the other branch requiring |μ12αi+ı2(1+βi)2αi2|<12(1βi)\big{|}\mu-\frac{1}{2}\alpha_{i}+\frac{\imath}{2}\sqrt{(1+\beta_{i})^{2}-\alpha_{i}^{2}}\big{|}<\frac{1}{2}(1-\beta_{i}) is omitted here since it is equivalent to (3.21) by replacing the undetermined parameter μ\mu with μ¯\bar{\mu}. Since the coefficients αi,βi\alpha_{i},\beta_{i} are dependent on the ratios rir_{i} and ri1r_{i-1}, The inequalities in (3.21) define a family of complex disks 𝔇(ri,ri1)\mathfrak{D}(r_{i},r_{i-1}) centered at (12αi,12(1+βi)2αi2)\big{(}\frac{1}{2}\alpha_{i},\frac{1}{2}\sqrt{(1+\beta_{i})^{2}-\alpha_{i}^{2}}\big{)} with the radius 12(1βi)\frac{1}{2}(1-\beta_{i}).

A heuristic analysis is considered firstly to obtain a rough estimate for the complex parameter μ\mu, while the mathematical proof is left to next subsection. To ensure (3.21), one should choose a fixed μ0\mu_{0} such that the intersection set

ri,ri1[0,R~3]𝔇(ri,ri1) is not empty.\bigcap_{r_{i},r_{i-1}\in[0,\tilde{R}_{3}]}\mathfrak{D}(r_{i},r_{i-1})\quad\text{ is not empty.}

Reminding the increasing functions (A.2)-(A.3), we see that the disk 𝔇(0,0)\mathfrak{D}(0,0) centered at (0,12)\big{(}0,\frac{1}{2}\big{)} has the maximum radius 12\frac{1}{2}, while the disk 𝔇(R~3,R~3)\mathfrak{D}(\tilde{R}_{3},\tilde{R}_{3}) centered at

(12α(R~3,R~3),12[1+β(R~3,R~3)]2α2(R~3,R~3))\left(\frac{1}{2}\alpha(\tilde{R}_{3},\tilde{R}_{3}),\frac{1}{2}\sqrt{\left[1+\beta(\tilde{R}_{3},\tilde{R}_{3})\right]^{2}-\alpha^{2}(\tilde{R}_{3},\tilde{R}_{3})}\right)

has the minimum radius 12(1β(R~3,R~3))\frac{1}{2}\big{(}1-\beta(\tilde{R}_{3},\tilde{R}_{3})\big{)}. Obviously, the largest value of R~3\tilde{R}_{3} may be determined from the fact that the smallest disk 𝔇(R~3,R~3)\mathfrak{D}(\tilde{R}_{3},\tilde{R}_{3}) is tangential to the largest one 𝔇(0,0)\mathfrak{D}(0,0) at the tangential point μ0\mu_{0}. This necessary condition holds if and only if

14α2(R~3,R~3)+(12[1+β(R~3,R~3)]2α2(R~3,R~3)12)2=14[2β(R~3,R~3)]2.\displaystyle\frac{1}{4}\alpha^{2}(\tilde{R}_{3},\tilde{R}_{3})+\left(\frac{1}{2}\sqrt{\left[1+\beta(\tilde{R}_{3},\tilde{R}_{3})\right]^{2}-\alpha^{2}(\tilde{R}_{3},\tilde{R}_{3})}-\frac{1}{2}\right)^{2}=\frac{1}{4}\big{[}2-\beta(\tilde{R}_{3},\tilde{R}_{3})\big{]}^{2}\,.

which leads to

α2(R~3,R~3)+8β2(R~3,R~3)8β(R~3,R~3)=0,\displaystyle\alpha^{2}(\tilde{R}_{3},\tilde{R}_{3})+8\beta^{2}(\tilde{R}_{3},\tilde{R}_{3})-8\beta(\tilde{R}_{3},\tilde{R}_{3})=0, (3.22)

or equivalently,

9R~362R~3535R~3442R~3322R~324R~3+1=0.\displaystyle 9\tilde{R}_{3}^{6}-2\tilde{R}_{3}^{5}-35\tilde{R}_{3}^{4}-42\tilde{R}_{3}^{3}-22\tilde{R}_{3}^{2}-4\tilde{R}_{3}+1=0.

This polynomial equation has two positive roots R~32.5808\tilde{R}_{3}\approx 2.5808 and R~30.1304\tilde{R}_{3}\approx 0.1304 (it is checked that this situation also occurs in [2]). We throw away the small root R~30.1304\tilde{R}_{3}\approx 0.1304 and choose the large one R~32.5808\tilde{R}_{3}\approx 2.5808 with the corresponding tangential point μ~0.4979+0.5454ı\tilde{\mu}_{*}\approx 0.4979+0.5454\imath.

Remark 1.

It is to mention that, by following the joint spectral radius approach in [8], one can obtain a slightly improved step-ratio constraint 0<rk<2.7050<r_{k}<2.705 to ensure the boundedness of DOC kernels; however, it seems difficult to obtain the desired decaying behavior of DOC kernels theoretically from the extremal norm estimate of the family of companion matrices {Ai|i3}\{A_{i}\,|\,i\geq 3\}.

Remark 2.

The condition (3.21) is sufficient for (3.17), or AiH,<1\left\|A_{i}\right\|_{H,\infty}<1; but different choices of the parameter μ\mu leads to different values of AiH,\left\|A_{i}\right\|_{H,\infty}. Consider the simple case with ri10r_{i-1}\equiv 0 for any i3i\geq 3, corresponding to the variable-step BDF2 scheme. By using the definition (3.5) together with (1.1)-(1.3), one has αi=ri1+2ri\alpha_{i}=\frac{r_{i}}{1+2r_{i}}, βi=0\beta_{i}=0 and the reduced class of complex discs

𝔇(ri,0)={μ||μ12α(ri,0)ı21α2(ri,0)|<12}for i3.\displaystyle\mathfrak{D}(r_{i},0)=\left\{\mu\,\left|\,\Big{|}\mu-\frac{1}{2}\alpha(r_{i},0)-\frac{\imath}{2}\sqrt{1-\alpha^{2}(r_{i},0)}\Big{|}<\frac{1}{2}\right.\right\}\quad\text{for $i\geq 3$.}

Recalling the fact (αi+12)(αi1)<0(\alpha_{i}+\frac{1}{2})(\alpha_{i}-1)<0 for any ri0r_{i}\geq 0, one can check that

|1412α(ri,0)+ı2(321α2(ri,0))|<12.\left|\frac{1}{4}-\frac{1}{2}\alpha(r_{i},0)+\frac{\imath}{2}\Big{(}\frac{\sqrt{3}}{2}-\sqrt{1-\alpha^{2}(r_{i},0)}\Big{)}\right|<\frac{1}{2}.

That is, the distance between the center (14,34)\big{(}\frac{1}{4},\frac{\sqrt{3}}{4}\big{)} of 𝔇(,0)\mathfrak{D}(\infty,0) and the center (12αi,121αi2)\big{(}\frac{1}{2}\alpha_{i},\frac{1}{2}\sqrt{1-\alpha_{i}^{2}}\big{)} of 𝔇(ri,0)\mathfrak{D}(r_{i},0) is always less than 12\frac{1}{2}. Thus these complex discs are always have a common region for any ri>0r_{i}>0, that is, ri>0𝔇(ri,0)\cap_{r_{i}>0}\mathfrak{D}(r_{i},0) is not empty. There exists a fixed parameter μ\mu such that AiH,<1\big{\|}A_{i}\big{\|}_{H,\infty}<1 for any ri>0r_{i}>0. By taking μ1=ı/2\mu_{1}=\imath/2 in (3.19), one has

AiH,=2|μ1||μ1αi||μ1μ¯1|=αi2+14for ri>0.\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}=\frac{2\left|\mu_{1}\right|\left|\mu_{1}-\alpha_{i}\right|}{\left|\mu_{1}-\bar{\mu}_{1}\right|}=\sqrt{\alpha_{i}^{2}+\frac{1}{4}}\quad\text{for $r_{i}>0$.}

Since αi(0,12)\alpha_{i}\in(0,\frac{1}{2}) for any ri>0r_{i}>0, one gets AiH,(12,12)\left\|A_{i}\right\|_{H,\infty}\in\big{(}\tfrac{1}{2},\tfrac{1}{\sqrt{2}}\big{)}. On the other hand, taking μ2=1/4+ı/4\mu_{2}=1/4+\imath/4 in (3.19) arrives at

AiH,=2|μ2||μ2αi||μ2μ¯2|=2(14αi)2+18for ri>0.\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}=\frac{2\left|\mu_{2}\right|\left|\mu_{2}-\alpha_{i}\right|}{\left|\mu_{2}-\bar{\mu}_{2}\right|}=\sqrt{2\big{(}\frac{1}{4}-\alpha_{i}\big{)}^{2}+\frac{1}{8}}\quad\text{for $r_{i}>0$.}

For any ri>0r_{i}>0, we have AiH,(24,12)\left\|A_{i}\right\|_{H,\infty}\in\big{(}\frac{\sqrt{2}}{4},\frac{1}{2}\big{)}, which are less than 1/2. Always, the estimate (3.16) implies the globally asymptotic decaying of DOC kernels, as shown by Lemma 2.1.

3.2 Stability analysis

To avoid the undesired equation (3.22) which has two different positive roots, and to simplify the subsequent mathematical derivations, we fix the complex parameter

μ=μ:=1/2+ı/2,\mu=\mu_{*}:=1/2+\imath/2,

which is very close to the optimal tangential point μ~0.4979+0.5454ı\tilde{\mu}_{*}\approx 0.4979+0.5454\imath in the heuristic analysis. Reminding the condition (3.21), we determine the maximum step-ratio R3R_{3} by

|1α(R3,R3)+ı(1(1+β(R3,R3))2α(R3,R3)2)|=1β(R3,R3),\displaystyle\left|1-\alpha(R_{3},R_{3})+\imath\Big{(}1-\sqrt{(1+\beta(R_{3},R_{3}))^{2}-\alpha(R_{3},R_{3})^{2}}\Big{)}\right|=1-\beta(R_{3},R_{3}),

which leads to the polynomial equation

R36+R354R348R3310R326R32=0.\displaystyle R_{3}^{6}+R_{3}^{5}-4R_{3}^{4}-8R_{3}^{3}-10R_{3}^{2}-6R_{3}-2=0.

It has a unique positive root R32.553R_{3}\approx 2.553. Now we prove the following result.

Lemma 3.1.

If the step ratios rk<R3r_{k}<R_{3} (k2)(k\geq 2), there exists a constant C3>0C_{3}>0, independent of the time tnt_{n}, such that the DOC kernels ϑnj(3,n)\vartheta_{n-j}^{(3,n)} in (1.11) satisfy

j=3n|ϑnj(3,n)|C3andj=in|ϑji(3,j)|C3for n3(i3).\displaystyle\sum_{j=3}^{n}\big{|}\vartheta_{n-j}^{(3,n)}\big{|}\leq C_{3}\quad\text{and}\quad\sum_{j=i}^{n}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\leq C_{3}\quad\text{for $n\geq 3\;(i\geq 3)$.}
Proof.

The definitions (1.1) and (1.10) give d0(3,j)=d0(rj,rj1)1d^{(3,j)}_{0}=d_{0}(r_{j},r_{j-1})\geq 1 for j3j\geq 3. By using (3.1), (3.3) and the coefficient bound (3.12), one gets

ϑ0(3,j)=1d0(3,j)1andϑ1(3,j+1)=αj+1d0(3,j)2d0(3,j)2,for j3.\vartheta_{0}^{(3,j)}=\frac{1}{d^{(3,j)}_{0}}\leq 1\quad\text{and}\quad\vartheta_{1}^{(3,j+1)}=\frac{\alpha_{j+1}}{d^{(3,j)}_{0}}\leq\frac{2}{d^{(3,j)}_{0}}\leq 2,\quad\text{for $j\geq 3$.}

That is, the first two kernels are always positive and bounded so that

j=3n|ϑnj(3,n)|3andj=in|ϑji(3,j)|3for n=3,4(i3).\displaystyle\sum_{j=3}^{n}\big{|}\vartheta_{n-j}^{(3,n)}\big{|}\leq 3\quad\text{and}\quad\sum_{j=i}^{n}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\leq 3\quad\text{for $n=3,4\;(i\geq 3)$.}

Now consider the general case n5n\geq 5. Taking μ=μ\mu=\mu^{*} in the condition (3.20) yields

|2βiαi+ı(1αi)|<1βifor i3,\displaystyle\big{|}2\beta_{i}-\alpha_{i}+\imath(1-\alpha_{i})\big{|}<1-\beta_{i}\quad\text{for $i\geq 3$,}

or equivalently,

2αi2+3βi24αiβi2αi+2βi<0for i3.\displaystyle 2\alpha_{i}^{2}+3\beta_{i}^{2}-4\alpha_{i}\beta_{i}-2\alpha_{i}+2\beta_{i}<0\quad\text{for $i\geq 3$.}

Thanks to Lemma A.3, they are valid if all of step-ratios 0<rk<R30<r_{k}<R_{3} (k2)(k\geq 2). According to (3.10) and (3.19), the elliptic norm AiH,\big{\|}A_{i}\big{\|}_{H,\infty} is a continuous function with respect to ri,ri1[ϵ,R3ϵ]r_{i},r_{i-1}\in[\epsilon,R_{3}-\epsilon] for aribitrary small ϵ>0\epsilon>0. Thus there exists a constant δ(0,1)\delta\in(0,1) such that

AiH,δ<1for i3.\displaystyle\big{\|}A_{i}\big{\|}_{H,\infty}\leq\delta<1\quad\text{for $i\geq 3$.}

By taking the parameter μ=μ\mu=\mu^{*} in (3.15) and (3.19), one has H=2\big{\|}H\big{\|}_{\infty}=2 and H1=1+22.\big{\|}H^{-1}\big{\|}_{\infty}=1+\frac{\sqrt{2}}{2}. Moreover, by (A.4)-(A.5), one has uj+1max{α(R3,R3),1}=α(R3,R3)\big{\|}\vec{u}_{j+1}\big{\|}_{\infty}\leq\max\{\alpha(R_{3},R_{3}),1\}=\alpha(R_{3},R_{3}). Thus, the estimate (3.16) gives

|ϑnj(3,n)|d0(3,j)uncRδnj1for nj+25,\displaystyle\big{|}\vartheta_{n-j}^{(3,n)}\big{|}d^{(3,j)}_{0}\leq\big{\|}\vec{u}_{n}\big{\|}_{\infty}\leq c_{R}\delta^{n-j-1}\quad\text{for $n\geq j+2\geq 5$,}

where cR=(2+2)α(R3,R3)c_{R}=(2+\sqrt{2})\alpha(R_{3},R_{3}), and then

|ϑnj(3,n)|cRδnj1for nj+25.\displaystyle\big{|}\vartheta_{n-j}^{(3,n)}\big{|}\leq c_{R}\delta^{n-j-1}\quad\text{for $n\geq j+2\geq 5$.} (3.23)

Therefore it follows that

j=3n|ϑnj(3,n)|cRj=3nδnj1<cRδ(1δ)for n5,\displaystyle\sum_{j=3}^{n}\big{|}\vartheta_{n-j}^{(3,n)}\big{|}\leq c_{R}\sum_{j=3}^{n}\delta^{n-j-1}<\frac{c_{R}}{\delta(1-\delta)}\quad\text{for $n\geq 5$,}
j=in|ϑji(3,j)|cRj=inδji1<cRδ(1δ)for ni3.\displaystyle\sum_{j=i}^{n}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\leq c_{R}\sum_{j=i}^{n}\delta^{j-i-1}<\frac{c_{R}}{\delta(1-\delta)}\quad\text{for $n\geq i\geq 3$.}

We obtain the desired estimates by taking a constant C3:=cRδ(1δ),C_{3}:=\frac{c_{R}}{\delta(1-\delta)}, which is independent of the time-level index nn. This proof is complete. ∎

Remark 3.

The elliptic type norm in (3.15) admits some other nonsingular matrices HH. Here consider a simple case with

H0:=(1011)such thatH01AiH0=(αiβiβi1αi+βiβi)H_{0}:=\Big{(}\begin{array}[]{cc}1&0\\ 1&1\\ \end{array}\Big{)}\quad\text{such that}\quad H_{0}^{-1}A_{i}H_{0}=\left(\begin{array}[]{cc}\alpha_{i}-\beta_{i}&-\beta_{i}\\ 1-\alpha_{i}+\beta_{i}&\beta_{i}\\ \end{array}\right)

Let R3,0:=13+1319+3333+131933331.839R_{3,0}:=\frac{1}{3}+\frac{1}{3}\sqrt[3]{19+3\sqrt{33}}+\frac{1}{3}\sqrt[3]{19-3\sqrt{33}}\approx 1.839 be the unique positive root of the cubic equation R3,03R3,02R3,01=0R_{3,0}^{3}-R_{3,0}^{2}-R_{3,0}-1=0. If the step-ratios satisfy 0<rk<R3,00<r_{k}<R_{3,0} (it is also superior to the existing mesh conditions in the literature [2, 6, 8]), one can follow the proofs of Lemmas A.1-A.2 to check that

2βi<αi<1 for i3.2\beta_{i}<\alpha_{i}<1\quad\text{ for $i\geq 3$.}

It follows that

AiH0,=max{αi,1αi+2βi}<1for i3,\big{\|}A_{i}\big{\|}_{H_{0},\infty}=\max\left\{\alpha_{i},1-\alpha_{i}+2\beta_{i}\right\}<1\quad\text{for $i\geq 3$,}

and then the DOC kernels ϑnj(3,n)\vartheta_{n-j}^{(3,n)} are globally decaying. Although the proof of Lemma 3.1 is technically complex, the matrix HH in (3.15) would be optimal in the sense that the companion matrix AiA_{i} always has a pair of conjugate complex eigenvalues. Actually, the inequality αi2<4βi\alpha_{i}^{2}<4\beta_{i} holds when the adjacent step ratios ri,ri1r_{i},r_{i-1} are close to 1, cf. Remark 4.

Remark 4.

Consider the uniform time-step τk=τ\tau_{k}=\tau, the BDF3 kernels (1.10) give

d0(3,n)=116,d1(3,n)=76,d2(3,n)=13anddj(3,n)=0for j3.d_{0}^{(3,n)}=\frac{11}{6},\quad d_{1}^{(3,n)}=-\frac{7}{6},\quad d_{2}^{(3,n)}=\frac{1}{3}\quad\text{and}\quad d_{j}^{(3,n)}=0\quad\text{for $j\geq 3$}.

The difference equation (3.4) becomes

ϑ^nj(3,n)711ϑ^n1j(3,n1)+211ϑ^n2j(3,n2)=0for nj+25.\displaystyle\widehat{\vartheta}_{n-j}^{(3,n)}-\frac{7}{11}\widehat{\vartheta}_{n-1-j}^{(3,n-1)}+\frac{2}{11}\widehat{\vartheta}_{n-2-j}^{(3,n-2)}=0\quad\text{for $n\geq j+2\geq 5$.}

Since the associated characteristic equation λ2711λ+211=0\lambda^{2}-\frac{7}{11}\lambda+\frac{2}{11}=0 has two roots λ1,2=(7±ı39)/22\lambda_{1,2}=(7\pm\imath\sqrt{39})/22, we have the following explicit formula of DOC kernels

ϑnj(3,n)=11ıd0(3,j)39[(7ı3922)nj+1(7+ı3922)nj+1]for nj+25.\displaystyle\vartheta_{n-j}^{(3,n)}=\frac{11\,\imath}{d_{0}^{(3,j)}\sqrt{39}}\left[\Big{(}\frac{7-\imath\sqrt{39}}{22}\Big{)}^{n-j+1}-\Big{(}\frac{7+\imath\sqrt{39}}{22}\Big{)}^{n-j+1}\right]\quad\text{for $n\geq j+2\geq 5$.}
Refer to caption
Figure 1: The DOC kernels ϑj(3,n)(j2)\vartheta_{j}^{(3,n)}\,(j\geq 2) for different step-ratio patterns.

In Figure 1, we fix n=30n=30 and list the values of DOC kernels ϑj(3,n)\vartheta_{j}^{(3,n)} for different choices (let 0<ϵk<10<\epsilon_{k}<1 be the uniformly distributed random number):

  • (a)

    the uniform mesh with rk=1r_{k}=1 for 2k302\leq k\leq 30;

  • (b)

    the random mesh with rk:=R3ϵkr_{k}:=R_{3}\epsilon_{k} for 2k302\leq k\leq 30;

  • (c)

    the random mesh with rk:=3ϵkr_{k}:=3\epsilon_{k} for 2k302\leq k\leq 30.

As observed, the DOC kernels ϑnj(3,n)\vartheta_{n-j}^{(3,n)} are not always positive, but they always decay fast, as predicted by (3.23). It is interesting to find that, from the case (c), the DOC kernels maintain the fast damping property as a whole although there are many of step ratios greater than R3R_{3} (about 15%15\% in the above test).

Theorem 3.1.

If τ1/(4C3Lf)\tau\leq 1/(4C_{3}L_{f}), the BDF3 solution of (1.7) satisfies

|v~n|2exp(4C3Lftn1)(|v~2|+5C3τ|τv~2|+2C3τ|τv~1|+2C3tnmax3in|εi|)for 3nN.\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq 2\exp(4C_{3}L_{f}t_{n-1})\Big{(}\big{|}\tilde{v}^{2}\big{|}+5C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}+2C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+2C_{3}t_{n}\max_{3\leq i\leq n}\big{|}\varepsilon^{i}\big{|}\Big{)}\quad\text{for $3\leq n\leq N$.}

Thus the variable-step BDF3 scheme is stable if 0<rk<R30<r_{k}<R_{3} for k2k\geq 2.

Proof.

Multiplying both sides of (1.16) with 2τnv~n2\tau_{n}\tilde{v}^{n}, one applies Lemma 3.1 to obtain that

|v~n|2|v~n1|2\displaystyle\big{|}\tilde{v}^{n}\big{|}^{2}-\big{|}\tilde{v}^{n-1}\big{|}^{2}\leq 2τnv~n3n[v~]+2τnv~ni=3nϑni(3,n)[f(ti,v¯i)f(ti,vi)+εi]\displaystyle\,-2\tau_{n}\tilde{v}^{n}\mathcal{I}_{3}^{n}[\tilde{v}]+2\tau_{n}\tilde{v}^{n}\sum_{i=3}^{n}\vartheta_{n-i}^{(3,n)}\left[f(t_{i},\bar{v}^{i})-f(t_{i},v^{i})+\varepsilon^{i}\right]
\displaystyle\leq  2τn|v~n||3n[v~]|+2Lfτn|v~n|i=3n|ϑni(3,n)||v~i|+2C3|v~n|τnmax3in|εi|.\displaystyle\,2\tau_{n}\big{|}\tilde{v}^{n}\big{|}\big{|}\mathcal{I}_{3}^{n}[\tilde{v}]\big{|}+2L_{f}\tau_{n}\big{|}\tilde{v}^{n}\big{|}\sum_{i=3}^{n}\big{|}\vartheta_{n-i}^{(3,n)}\big{|}\big{|}\tilde{v}^{i}\big{|}+2C_{3}\big{|}\tilde{v}^{n}\big{|}\tau_{n}\max_{3\leq i\leq n}\big{|}\varepsilon^{i}\big{|}.

By summing the above inequality for nn from n=3n=3 to mm, we have

|v~m|2\displaystyle\big{|}\tilde{v}^{m}\big{|}^{2}\leq |v~2|2+2j=3mτj|v~j||3j[v~]|+2Lfj=3mτj|v~j|i=3j|ϑji(3,j)||v~i|+2C3j=3m|v~j|τjmax3ij|εi|.\displaystyle\,\big{|}\tilde{v}^{2}\big{|}^{2}+2\sum_{j=3}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}\big{|}\mathcal{I}_{3}^{j}[\tilde{v}]\big{|}+2L_{f}\sum_{j=3}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}\sum_{i=3}^{j}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\big{|}\tilde{v}^{i}\big{|}+2C_{3}\sum_{j=3}^{m}\big{|}\tilde{v}^{j}\big{|}\tau_{j}\max_{3\leq i\leq j}\big{|}\varepsilon^{i}\big{|}.

Taking some integer m1m_{1} (2m1m2\leq m_{1}\leq m) such that |v~m1|=max2km|v~k|\big{|}\tilde{v}^{m_{1}}\big{|}=\max_{2\leq k\leq m}\big{|}\tilde{v}^{k}\big{|}. Now we take m:=m1m:=m_{1} in the above inequality and get

|v~m||v~m1||v~2|+2τj=3m|3j[v~]|+2C3Lfj=3mτj|v~j|+2C3tmmax3im|εi|\displaystyle\big{|}\tilde{v}^{m}\big{|}\leq\big{|}\tilde{v}^{m_{1}}\big{|}\leq\big{|}\tilde{v}^{2}\big{|}+2\tau\sum_{j=3}^{m}\big{|}\mathcal{I}_{3}^{j}[\tilde{v}]\big{|}+2C_{3}L_{f}\sum_{j=3}^{m}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+2C_{3}t_{m}\max_{3\leq i\leq m}\big{|}\varepsilon^{i}\big{|} (3.24)

for 3mN3\leq m\leq N. It remains to evaluate the error j=3m|3j[v~]|\sum_{j=3}^{m}\big{|}\mathcal{I}_{3}^{j}[\tilde{v}]\big{|} from the starting values. Taking the index k=3\mathrm{k}=3 in (1.15) gives

3n[v~]=\displaystyle\mathcal{I}_{3}^{n}[\tilde{v}]= τv~2i=3nϑni(3,n)di2(3,i)+τv~1i=3nϑni(3,n)di1(3,i)for n3.\displaystyle\,\partial_{\tau}\tilde{v}^{2}\sum_{i=3}^{n}\vartheta_{n-i}^{(3,n)}d^{(3,i)}_{i-2}+\partial_{\tau}\tilde{v}^{1}\sum_{i=3}^{n}\vartheta_{n-i}^{(3,n)}d^{(3,i)}_{i-1}\quad\text{for $n\geq 3$.}

Recalling the definition (1.10) of BDF3 kernels with the increasing property (A.1), it is easy to check that d2(3,3)d2(R3,R3)2d^{(3,3)}_{2}\leq d_{2}(R_{3},R_{3})\leq 2 and |d1(3,3)|+d2(3,4)d1(R3,R3)+d2(R3,R3)5\big{|}d^{(3,3)}_{1}\big{|}+d^{(3,4)}_{2}\leq-d_{1}(R_{3},R_{3})+d_{2}(R_{3},R_{3})\leq 5. Thus we apply Lemma 3.1 to get

j=3m|3j[v~]|\displaystyle\sum_{j=3}^{m}\big{|}\mathcal{I}_{3}^{j}[\tilde{v}]\big{|}\leq |τv~2|j=3mi=3j|ϑji(3,j)||di2(3,i)|+|τv~1|j=3mi=3j|ϑji(3,j)||di1(3,i)|\displaystyle\,\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}\sum_{j=3}^{m}\sum_{i=3}^{j}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\big{|}d^{(3,i)}_{i-2}\big{|}+\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}\sum_{j=3}^{m}\sum_{i=3}^{j}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}\big{|}d^{(3,i)}_{i-1}\big{|}
=\displaystyle= |τv~2|i=3m|di2(3,i)|j=im|ϑji(3,j)|+|τv~1|i=3m|di1(3,i)|j=im|ϑji(3,j)|\displaystyle\,\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}\sum_{i=3}^{m}\big{|}d^{(3,i)}_{i-2}\big{|}\sum_{j=i}^{m}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}+\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}\sum_{i=3}^{m}\big{|}d^{(3,i)}_{i-1}\big{|}\sum_{j=i}^{m}\big{|}\vartheta_{j-i}^{(3,j)}\big{|}
\displaystyle\leq C3|τv~2|(|d1(3,3)|+d2(3,4))+C3|τv~1|d2(3,3)\displaystyle\,C_{3}\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}\big{(}\big{|}d^{(3,3)}_{1}\big{|}+d^{(3,4)}_{2}\big{)}+C_{3}\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}d^{(3,3)}_{2}
\displaystyle\leq  5C3|τv~2|+2C3|τv~1|for m3.\displaystyle\,5C_{3}\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}+2C_{3}\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}\qquad\text{for $m\geq 3$.}

It follows from (3.24) that

|v~n||v~2|+5C3τ|τv~2|+2C3τ|τv~1|+2C3Lfj=3nτj|v~j|+2C3tnmax3in|εi|.\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq\big{|}\tilde{v}^{2}\big{|}+5C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}+2C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+2C_{3}L_{f}\sum_{j=3}^{n}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+2C_{3}t_{n}\max_{3\leq i\leq n}\big{|}\varepsilon^{i}\big{|}.

Assuming that τ1/(4C3Lf)\tau\leq 1/(4C_{3}L_{f}), one has

|v~n|2|v~2|+10C3τ|τv~2|+4C3τ|τv~1|+4C3Lfj=3n1τj|v~j|+4C3tnmax3in|εi|\displaystyle\big{|}\tilde{v}^{n}\big{|}\leq 2\big{|}\tilde{v}^{2}\big{|}+10C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{2}\big{|}+4C_{3}\tau\big{|}\partial_{\tau}\tilde{v}^{1}\big{|}+4C_{3}L_{f}\sum_{j=3}^{n-1}\tau_{j}\big{|}\tilde{v}^{j}\big{|}+4C_{3}t_{n}\max_{3\leq i\leq n}\big{|}\varepsilon^{i}\big{|}

for 3nN3\leq n\leq N. The standard Grönwall inequality completes the proof. ∎

From the proof of Lemma 3.1, the uniform boundedness of j=1n|ϑnj(3,n)|\sum_{j=1}^{n}\big{|}\vartheta_{n-j}^{(3,n)}\big{|} and j=in|ϑji(3,j)|\sum_{j=i}^{n}\big{|}\vartheta_{j-i}^{(3,j)}\big{|} does not require all DOC kernels to decay rapidly. It always allows a finite number of bounded DOC kernels. As seen from the numerical tests in the next section, the imposed step-ratio condition 0<rk<R30<r_{k}<R_{3} in Theorem 3.1 is sufficient, but far from necessary. A practical step-ratio constraint for stability is that

almost all step ratios satisfy 0<rk<R30<r_{k}<R_{3} for 2kN2\leq k\leq N.

4 Numerical tests

Consider an ODE model v=2v3etv^{\prime}=2v-3e^{-t} for 0<t10<t\leq 1 with a smooth solution v=exp(t)v=\exp(-t) in our numerical tests. We run the variable-step BDF2 and BDF3 schemes in two scenarios:

  1. (a)

    The graded meshes tk=(k/N)γt_{k}=(k/N)^{\gamma} for 0kN0\leq k\leq N with grading parameters γ>1\gamma>1. The maximum step-ratio is rmax=r2=2γ1r_{\max}=r_{2}=2^{\gamma}-1 and τ/τ1=τN/τ1=Nγ(N1)γγNγ1\tau/\tau_{1}=\tau_{N}/\tau_{1}=N^{\gamma}-(N-1)^{\gamma}\approx\gamma N^{\gamma-1}.

  2. (b)

    The random meshes with time-steps τk=ϵk/k=1Nϵk\tau_{k}=\epsilon_{k}/\sum_{k=1}^{N}\epsilon_{k}, where ϵk(0,1)\epsilon_{k}\in(0,1) are uniformly distributed random numbers.

To start the multi-step schemes, a two-stage third-order singly diagonally implicit Runge-Kutta method is employed. We record the maximum error e(N):=max1nN|v(tn)vn|e(N):=\max_{1\leq{n}\leq{N}}|v(t_{n})-v^{n}| in each run and compute the convergence order by

Orderlog(e(N)/e(2N))log(τ(N)/τ(2N))\text{Order}\approx\frac{\log\left(e(N)/e(2N)\right)}{\log\left(\tau(N)/\tau(2N)\right)}

where τ(N)\tau(N) denotes the maximum time-step size for total NN subintervals.

Table 1: Numerical results of BDF2 method on graded time meshes

  NN ττ1\dfrac{\tau}{\tau_{1}} γ=2,rmax=3\gamma=2,\,\,r_{\max}=3 ττ1\dfrac{\tau}{\tau_{1}} γ=3,rmax=7\gamma=3,\,\,r_{\max}=7 ττ1\dfrac{\tau}{\tau_{1}} γ=4,rmax=15\gamma=4,\,\,r_{\max}=15 e(N)e(N) Order e(N)e(N) Order e(N)e(N) Order 40 7.9e+01 5.28e-04 - 4.7e+03 8.77e-04 - 2.5e+05 1.35e-03 - 80 1.6e+02 1.34e-04 1.97 1.9e+04 2.25e-04 1.96 2.0e+06 3.49e-04 1.95 160 3.2e+02 3.39e-05 1.99 7.6e+04 5.72e-05 1.98 1.6e+07 8.91e-05 1.97 320 6.4e+02 8.52e-06 1.99 3.1e+05 1.44e-05 1.99 1.3e+08 2.25e-05 1.98 640 1.3e+03 2.14e-06 2.00 1.2e+06 3.61e-06 1.99 1.1e+09 5.66e-06 1.99 1280 2.6e+03 5.34e-07 2.00 4.9e+06 9.06e-07 2.00 8.4e+09 1.42e-06 2.00  

Table 2: Numerical results of BDF3 method on graded time meshes

  NN ττ1\dfrac{\tau}{\tau_{1}} γ=2,rmax=3\gamma=2,\,\,r_{\max}=3 ττ1\dfrac{\tau}{\tau_{1}} γ=3,rmax=7\gamma=3,\,\,r_{\max}=7 ττ1\dfrac{\tau}{\tau_{1}} γ=4,rmax=15\gamma=4,\,\,r_{\max}=15 e(N)e(N) Order e(N)e(N) Order e(N)e(N) Order 40 7.9e+01 1.27e-05 - 4.7e+03 2.94e-05 - 2.5e+05 5.73e-05 - 80 1.6e+02 1.65e-06 2.95 1.9e+04 3.91e-06 2.91 2.0e+06 7.85e-06 2.87 160 3.2e+02 2.10e-07 2.97 7.6e+04 5.05e-07 2.96 1.6e+07 1.03e-06 2.94 320 6.4e+02 2.65e-08 2.99 3.1e+05 6.41e-08 2.98 1.3e+08 1.31e-07 2.97 640 1.3e+03 3.32e-09 2.99 1.2e+06 8.07e-09 2.99 1.1e+09 1.66e-08 2.98 1280 2.6e+03 4.16e-10 3.00 4.9e+06 1.01e-09 2.99 8.4e+09 2.08e-09 2.99  

Tables 1-2 list the numerical results of the BDF2 and BDF3 methods on graded meshes with three grading parameters, respectively. We see that, although the ratio τ/τ1\tau/\tau_{1} increases as fast as O(Nγ1)O(N^{\gamma-1}), the numerical solutions remains stable and convergent with full accuracy.

Table 3: Numerical results of BDF2 method on random time meshes

  NN τ(N)\tau(N) e(N)e(N) Order rmaxr_{\max} τ/τ1\tau/\tau_{1} 40 4.49e-02 5.86e-04 28.30 1.22 80 2.40e-02 1.43e-04 2.03 91.41 1.06 160 1.18e-02 4.09e-05 1.81 32.54 19.88 320 6.18e-03 1.03e-05 1.99 418.41 1.21 640 3.01e-03 2.51e-06 2.04 604.02 2.66 1280 1.55e-03 6.63e-07 1.92 1963.80 1.02  

Table 4: Numerical results of BDF3 method on random time meshes

  NN τ(N)\tau(N) e(N)e(N) Order rmaxr_{\max} N1N_{1} τ/τ1\tau/\tau_{1} 40 4.60e-02 8.20e-06 7.58 9 1.57 80 2.46e-02 1.18e-06 2.79 361.49 18 1.93 160 1.22e-02 1.83e-07 2.69 1682.18 35 4.51 320 6.11e-03 2.40e-08 2.93 79.90 60 2.91 640 3.27e-03 3.40e-09 2.82 5765.00 146 1.20 1280 1.56e-03 4.16e-10 3.03 9677.92 250 6.64  

Tables 3-4 record the numerical results on random time meshes. Table 3 suggests that the variable-step BDF2 method is robust with respect to the step-ratios rkr_{k}, and supports the theoretical findings in Theorem 2.1. Reminding the step-ratio restriction 0<rk<2.5330<r_{k}<2.533 in Theorem 3.1 for the BDF3 method, we also record the number (denoted by N1N_{1} in Table 4) of time levels with rk2.533r_{k}\geq 2.533. It is seen that the variable-step BDF3 method is stable and third-order convergent on random meshes, even if there are many of (about 20%20\% in our tests) step-ratios do not meet our theoretical condition. Nonetheless, it remains mysterious to us.

5 Concluding remarks

The stability of BDF2 and BDF3 methods with unequal time-steps is investigated by a new theoretical framework using the discrete orthogonal convolution kernels. Thanks to the global analysis, that is, the present technique formulates the current BDF solution as a convolution summation of all previous information with DOC kernels as the convolutional weights, see the form (1.16), we improved the setp-ratio constraints for the stability. It is to mention that, the global decaying estimates of DOC kernels are also critical in the numerical analysis of BDF methods for parabolic equations, cf. [11, 12, 14]. The stability and convergence theory of variable-step BDF3 scheme for these stiff problems will be addressed in forthcoming reports.

For the high-order BDF-k\mathrm{k} (k=4,5,6\mathrm{k}=4,5,6) time-stepping methods, we can also define the associated DOC kernels ϑnj(k,n)\vartheta_{n-j}^{(\mathrm{k},n)} via the recursive procedure (1.11). From the proof of Theorem 3.1, one needs a result similar to Lemma 3.1 under certain step-ratio condition. Actually, the uniform boundedness (there exists a constant CkC_{\mathrm{k}} independent of the time-level indexes nn) of the absolute summations of DOC kernels

j=kn|ϑnj(k,n)|Ckandj=in|ϑji(k,j)|Ckfor any nik,\displaystyle\sum_{j=\mathrm{k}}^{n}\big{|}\vartheta_{n-j}^{(\mathrm{k},n)}\big{|}\leq C_{\mathrm{k}}\quad\text{and}\quad\sum_{j=i}^{n}\big{|}\vartheta_{j-i}^{(\mathrm{k},j)}\big{|}\leq C_{\mathrm{k}}\quad\text{for any $n\geq i\geq\mathrm{k}$,}

is sufficient to the stability of BDF-k\mathrm{k} schemes for initial value problems. However, this issue seems to be theoretically challenging for k4\mathrm{k}\geq 4 and remains open to us.

Acknowledgements

The authors would like to thank Dr. Ji Bingquan for his valuable discussions.

Appendix A Two auxiliary functions for the BDF3 method

By using the definitions (1.1)-(1.3), it is easy to check that

x|dν(x,y)|>0andy|dν(x,y)|>0for ν=0,1,2 and x,y>0.\displaystyle\frac{\partial}{\partial x}\left|d_{\nu}(x,y)\right|>0\quad\text{and}\quad\frac{\partial}{\partial y}\left|d_{\nu}(x,y)\right|>0\quad\text{for $\nu=0,1,2$ and $x,y>0.$} (A.1)

That is, the functions d0(x,y)d_{0}(x,y), d1(x,y)-d_{1}(x,y) and d2(x,y)d_{2}(x,y) are increasing with respect to x,y>0x,y>0. We consider two auxiliary functions for the analysis of variable-step BDF3 method,

α(x,y):=\displaystyle\alpha(x,y):= d1(x,y)d0(x,y)=x(x2y2+4xy2+3y2+2xy+3y+1)(y+1)(3x2y+4xy+2x+y+1),\displaystyle\,-\frac{d_{1}(x,y)}{d_{0}(x,y)}=\frac{x\left(x^{2}y^{2}+4xy^{2}+3y^{2}+2xy+3y+1\right)}{(y+1)\left(3x^{2}y+4xy+2x+y+1\right)}, (A.2)
β(x,y):=\displaystyle\beta(x,y):= d2(x,y)d0(x,y)=x(x+1)2y2(y+1)(3x2y+4xy+2x+y+1).\displaystyle\,\frac{d_{2}(x,y)}{d_{0}(x,y)}=\frac{x(x+1)^{2}y^{2}}{(y+1)\left(3x^{2}y+4xy+2x+y+1\right)}. (A.3)

Their properties will be examined with respect to two independent variables x,y>0x,y>0 due to the facts αm=α(rm,rm1)\alpha_{m}=\alpha(r_{m},r_{m-1}) and βm=β(rm,rm1)\beta_{m}=\beta(r_{m},r_{m-1}) for m2m\geq 2 according to (3.5).

The above functions α\alpha and β\beta are strictly increasing with respect to x,y>0x,y>0. Actually, by simple but tedious calculations for (A.2) -(A.3), it is not difficult to check that

αx=(x+1)2(3x2+2x+3)y3+2(2x3+5x2+6x+3)y2+(x+2)2y+1(y+1)(3x2y+4xy+2x+y+1)2,\displaystyle\frac{\partial\alpha}{\partial x}=\frac{(x+1)^{2}\left(3x^{2}+2x+3\right)y^{3}+2\left(2x^{3}+5x^{2}+6x+3\right)y^{2}+(x+2)^{2}y+1}{(y+1)\left(3x^{2}y+4xy+2x+y+1\right)^{2}}, (A.4)
αy=x(x+1)2(xy+y+1)(3xy+3y+1)(y+1)2(3x2y+4xy+2x+y+1)2,\displaystyle\frac{\partial\alpha}{\partial y}=\frac{x(x+1)^{2}(xy+y+1)(3xy+3y+1)}{(y+1)^{2}\left(3x^{2}y+4xy+2x+y+1\right)^{2}}, (A.5)

and

βx=(x+1)y2(3x3y+5x2y+4x2+3xy+3x+y+1)(y+1)(3x2y+4xy+2x+y+1)2,\displaystyle\frac{\partial\beta}{\partial x}=\frac{(x+1)y^{2}\left(3x^{3}y+5x^{2}y+4x^{2}+3xy+3x+y+1\right)}{(y+1)\left(3x^{2}y+4xy+2x+y+1\right)^{2}}, (A.6)
βy=x(x+1)2y(3x2y+6xy+4x+2y+2)(y+1)2(3x2y+4xy+2x+y+1)2for x,y>0.\displaystyle\frac{\partial\beta}{\partial y}=\frac{x(x+1)^{2}y\left(3x^{2}y+6xy+4x+2y+2\right)}{(y+1)^{2}\left(3x^{2}y+4xy+2x+y+1\right)^{2}}\quad\text{for $x,y>0$.} (A.7)

We have the following results. Some of proofs are technically complex and the mathematical derivations have been checked carefully by a symbolic calculation software.

Lemma A.1.

It holds that

0<β(x,y)<α(x,y)<1+β(x,y)for x,y>0.0<\beta(x,y)<\alpha(x,y)<1+\beta(x,y)\quad\text{for $x,y>0$}.
Proof.

Simple calculations lead to

α(x,y)β(x,y)=\displaystyle\alpha(x,y)-\beta(x,y)= x(2xy+2y+1)3x2y+4xy+2x+y+1,\displaystyle\,\frac{x(2xy+2y+1)}{3x^{2}y+4xy+2x+y+1},
1α(x,y)+β(x,y)=\displaystyle 1-\alpha(x,y)+\beta(x,y)= (x+1)(xy+y+1)3x2y+4xy+2x+y+1.\displaystyle\,\frac{(x+1)(xy+y+1)}{3x^{2}y+4xy+2x+y+1}.

It completes the proof. ∎

Lemma A.2.

It holds that

0<β(x,y)<1for 0<x,y<R^33.4405,0<\beta(x,y)<1\quad\text{for $0<x,y<\hat{R}_{3}\approx 3.4405$,}

where R^3\hat{R}_{3} is the unique positive root of R^342R^334R^323R^31=0\hat{R}_{3}^{4}-2\hat{R}_{3}^{3}-4\hat{R}_{3}^{2}-3\hat{R}_{3}-1=0.

Proof.

According to (A.6)-(A.7), it holds that

β(x,y)1<β(R^3,R^3)1=R^342R^334R^323R^313R^33+4R^32+3R^3+1=0for 0<x,y<R^3.\displaystyle\beta(x,y)-1<\beta(\hat{R}_{3},\hat{R}_{3})-1=\frac{\hat{R}_{3}^{4}-2\hat{R}_{3}^{3}-4\hat{R}_{3}^{2}-3\hat{R}_{3}-1}{3\hat{R}_{3}^{3}+4\hat{R}_{3}^{2}+3\hat{R}_{3}+1}=0\quad\text{for $0<x,y<\hat{R}_{3}$.}

It completes the proof. ∎

Lemma A.3.

For 0<x,y<R32.5530<x,y<R_{3}\approx 2.553, it holds that

2α2(x,y)+3β2(x,y)4α(x,y)β(x,y)2α(x,y)+2β(x,y)<0.2\alpha^{2}(x,y)+3\beta^{2}(x,y)-4\alpha(x,y)\beta(x,y)-2\alpha(x,y)+2\beta(x,y)<0.
Proof.

Consider an auxiliary function, see Figure 2,

g(x,y):=\displaystyle g(x,y):= 1x(x+1)[2α(x,y)2+3β(x,y)24α(x,y)β(x,y)2α(x,y)+2β(x,y)]\displaystyle\,\frac{1}{x(x+1)}\Big{[}2\alpha(x,y)^{2}+3\beta(x,y)^{2}-4\alpha(x,y)\beta(x,y)-2\alpha(x,y)+2\beta(x,y)\Big{]}
=\displaystyle= (3x2y+4xy+2x+y+1)γ(x,y)[(x+1)2(x2+x4)y42(4x2+11x+7)y3.\displaystyle\,\frac{\left(3x^{2}y+4xy+2x+y+1\right)}{\gamma(x,y)}\Big{[}(x+1)^{2}\left(x^{2}+x-4\right)y^{4}-2\left(4x^{2}+11x+7\right)y^{3}\Big{.}
.2(2x2+10x+9)y22(3x+5)y2]for 0x,y<R3,\displaystyle\,\hskip 99.58464pt\Big{.}-2\left(2x^{2}+10x+9\right)y^{2}-2(3x+5)y-2\Big{]}\qquad\text{for $0\leq x,y<R_{3}$,}

where

γ(x,y):=(y+1)2(3x2y+4xy+2x+y+1)3.\gamma(x,y):=(y+1)^{2}\left(3x^{2}y+4xy+2x+y+1\right)^{3}.
Refer to caption
Figure 2: The surface of gg on [0,R3]2[0,R_{3}]^{2}.

Obviously, it is sufficient to show that g(x,y)<0g(x,y)<0 for 0<x,y<R30<x,y<R_{3}. Simple but tedious calculations yield

gx=\displaystyle\frac{\partial g}{\partial x}= 1γ(x,y)[2(27x2+88x+63)y2+2(12x3+90x2+158x+79)y3.\displaystyle\,\frac{1}{\gamma(x,y)}\Big{[}2\left(27x^{2}+88x+63\right)y^{2}+2\left(12x^{3}+90x^{2}+158x+79\right)y^{3}\Big{.}
.+(4x4+58x3+207x2+252x+99)y4+(25x)(x+1)3y5+(36x+50)y+8].\displaystyle\,\hskip 28.45274pt\Big{.}+\left(4x^{4}+58x^{3}+207x^{2}+252x+99\right)y^{4}+(25-x)(x+1)^{3}y^{5}+(36x+50)y+8\Big{]}.

Obviously, we have gx>0\frac{\partial g}{\partial x}>0 such that gg is increasing with respect to xx for 0<x,y<R30<x,y<R_{3}. Thus it has no extreme points over the open square (0,R3)2(0,R_{3})^{2}. It remains to consider the maximum value along the four sides of [0,R3]2[0,R_{3}]^{2}:

  • (i)

    Along the side y=0y=0, we have g(x,0)=2(2x+1)2<0g(x,0)=-\frac{2}{(2x+1)^{2}}<0 for 0<x<R30<x<R_{3}..

  • (ii)

    Along the side y=R3y=R_{3}, g(x,R3)<g(R3,R3)0.0000277<0g(x,R_{3})<g(R_{3},R_{3})\approx-0.0000277<0 for 0<x<R30<x<R_{3}.

  • (iii)

    Along the side x=0x=0, we have g(0,y)=4y+2y+1<0g(0,y)=-\frac{4y+2}{y+1}<0 for 0<y<R30<y<R_{3}..

  • (iv)

    Along the side x=R3x=R_{3}, one has

    g(R3,y)\displaystyle g(R_{3},y)\approx 0.06763y40.12922y30.100507y20.0267488y0.002113(y+1)2(y+0.19847)2for 0<y<R3.\displaystyle\,\frac{0.06763y^{4}-0.12922y^{3}-0.100507y^{2}-0.0267488y-0.002113}{(y+1)^{2}(y+0.19847)^{2}}\quad\text{for $0<y<R_{3}$.}

    It has a unique minimum point at y0.103652y\approx 0.103652, while

    g(R3,y)<max{g(R3,0),g(R3,R3)}=g(R3,R3)<0for 0<y<R3.g(R_{3},y)<\max\{g(R_{3},0),g(R_{3},R_{3})\}=g(R_{3},R_{3})<0\quad\text{for $0<y<R_{3}$.}

In summary, we have g(x,y)<0g(x,y)<0 for 0<x,y<R30<x,y<R_{3}. It completes the proof. ∎

References

  • [1] G. Albi, G. Dimarco and L. Pareschi, Implicit-explicit multistep methods for hyperbolic systems with multiscale relaxation, SIAM J. Sci. Comput., 42(4) (2020) A2402–A2435 .
  • [2] M. Calvo, T. Grande and R.D. Grigorieff, On the zero stability of the variable order variable stepsize BDF-formulas, Numer. Math., 57 (1990), 39–50.
  • [3] M. Crouzeix and F.J. Lisbona, The convergence of variable-stepsize, variable formula, multistep methods, SIAM J. Numer. Anal., 21 (1984), 512–534.
  • [4] V. DeCaria, A. Guzel, W. Layton and Y. Li, A variable stepsize, variable order family of low complexity, SIAM J. Sci. Comp., 43(3) (2021), A2130-A2160.
  • [5] G. Dimarco and L. Pareschi, Implicit-explicit linear multistep methods for stiff kinetic equations, SIAM J. Numer. Anal., 55 (2017), 664–690.
  • [6] R.D. Grigorieff, Stability of multistep-methods on variable grids, Numer. Math., 42 (1983), 359–377.
  • [7] R.D. Grigorieff and P.J. Paes-Leme, On the zero-stability of the 3-step BDF-formula on nonuniform grids, BIT (1984), 24, 85–91.
  • [8] N. Guglielmi and M. Zennaro, On the zero-stability of variable stepsize multistep methods: the spectral radius approach, Numer. Math., 88 (2001), 445–458.
  • [9] E. Hairer, S.P. Nørsett and G. Wanner, Solving Ordinary Differential Equations I: Nonstiff Problems, Volume 8 of Springer Series in Computational Mathematics, Second Edition, Springer-Verlag, 1992.
  • [10] E. Hairer and G. Wanner, Solving Ordinary Differential Equations II: Stiff and Differential-Algebraic Problems, Springer Series in Computational Mathematics Volume 14, Second Edition, Springer-Verlag, 2002.
  • [11] H.-L. Liao, B. Ji and L. Zhang, An adaptive BDF2 implicit time-stepping method for the phase field crystal model, IMA J. Numer. Anal., 2020, doi:10.1093/imanum/draa075.
  • [12] H.-L. Liao, B. Ji, L. Wang and Z. Zhang, Mesh-robustness of an energy stable BDF2 scheme with variable steps for the Cahn-Hilliard model, arXiv:2102.03731v1, 2021.
  • [13] H.-L. Liao, T. Tang and T. Zhou, On energy stable, maximum-bound preserving, second-order BDF scheme with variable steps for the Allen-Cahn equation, SIAM J. Numer. Anal., 58(4) (2020), 2294-2314.
  • [14] H.-L. Liao and Z. Zhang, Analysis of adaptive BDF2 scheme for diffusion equations, Math. Comp., 90 (2021), 1207–1226.
  • [15] D. Wang and S.T. Ruuth, Variable step-size implicit-explicit linear multistep methods for time-dependent partial differential equations, J. Comput. Math., 26(6) (2008), 838–855.