An interpolation of discrete rough differential equations and its applications to analysis of error distributions


Abstract

We consider the solution \(Y_t\) \((0\le t\le 1)\) and several approximate solutions \(\hat{Y}^m_t\) of a rough differential equation driven by a fractional Brownian motion \(B_t\) with the Hurst parameter \(1/3<H\leq 1/2\) associated with a dyadic partition of \([0,1]\). We are interested in analysis of asymptotic error distribution of \(\hat{Y}^m_t-Y_t\) as \(m\to\infty\). In the preceding results, it was proved that the weak limit of \(\{(2^m)^{2H-1/2}(\hat{Y}^m_t-Y_t)\}_{0\le t\le 1}\) coincides with the weak limit of \(\{(2^m)^{2H-1/2}J_tI^m_t\}_{0\le t\le 1}\), where \(J_t\) is the Jacobian process of \(Y_t\) and \(I^m_t\) is a certain weighted sum process of Wiener chaos of order \(2\) defined by \(B_t\). However, it is non-trivial to reduce a problem about \(\hat{Y}^m_t-Y_t\) to one about \(J_t\) and \(I^m_t\). In this paper, we introduce an interpolation process between \(Y_t\) and \(\hat{Y}^m_t\), and give several estimates of the interpolation process itself and its associated processes. The analysis provides a framework to deal with the reduction problem and provides a stronger result that the difference \(R^m_t=\hat{Y}^m_t-Y_t-J_tI^m_t\) is really small compared to the main term \(J_tI^m_t\). More precisely, we show that \((2^m)^{2H-1/2+\varepsilon}\sup_{0\le t\le 1}|R^m_t|\to 0\) almost surely and in \(L^p\) (for all \(p>1\)) for certain explicit positive number \(\varepsilon>0\). As a consequence, we obtain an estimate of the convergence rate of \(\sup_{0\le t\le 1}|\hat{Y}^{m}_t-Y_t|\to 0\) in \(L^p\) also.

Keywords: Rough differential equation; Error distribution; Fractional Brownian motion

MSC2020 subject classifications: 60F05; 60H35; 60G15.

1 Introduction↩︎

In this paper, we study asymptotic error distributions for several approximation schemes of rough differential equations(=RDEs). Typical driving processes of RDEs are long-range correlated Gaussian processes and we cannot use several important tools in the study of stochastic differential equations driven by standard Brownian motions. For example, martingale central limit theorems cannot be applied to the study of asymptotic error distributions. However, the fourth moment theorem can be applicable for the study of long-range correlated Gaussian processes and several limit theorems of weighted sum processes of Wiener chaos have been established ([1][3] and references therein). Furthermore, these limit theorems are important in the study of asymptotic error distributions of RDEs ([4][9]). However, it is not trivial to reduce the problem of asymptotic error distributions of solutions of RDEs to that of weighted sum processes of Wiener chaos. We study this problem by introducing certain interpolation processes between the solution and the approximate solutions of RDEs.

More precisely, we explain our main results and the relation with previously known results. We consider a solution \(Y_t\) of a multidimensional RDE driven by fractional Brownian motion(=fBm) \(B_t\) with the Hurst parameter \(\frac{1}{3}<H\le\frac{1}{2}\), \[\begin{align} Y_t=\xi+\int_0^t\sigma(Y_s)d\boldsymbol{B}_s+\int_0^tb(Y_s)ds,\quad\quad 0\le t\le 1, \end{align}\] where \(\boldsymbol{B}_t\) is a naturally lifted geometric rough path of \(B_t\). The precise meanings of rough paths and RDEs will be given in Section 2. Let \(\hat{Y}^m_t\) be an approximate solution associated with the dyadic partition \(D_m=\{\tau^m_k\}_{k=0}^{2^m}\), where \(\tau^m_k=k2^{-m}\). Actually there are many approximation schemes, e.g., the implementable Milstein, Crank-Nicolson, Milstein and first-order Euler schemes of RDEs. The first-order Euler scheme was introduced by Hu-Liu-Nualart [5] and further studied by Liu-Tindel [7]. Among them, we explain the result in Liu and Tindel [7] which is closely related to our main results. For the first-order Euler approximate solution \(\hat{Y}^{m}_t\), they proved that \(\{(2^m)^{2H-\frac{1}{2}}(\hat{Y}^{m}_t-Y_t)\}_{0\le t\le 1}\) weakly converges to the weak limit of \(\{(2^m)^{2H-\frac{1}{2}}J_tI^m_t\}_{0\le t\le 1}\) as \(m\to\infty\) in \(D([0,1])\) with respect to the Skorokhod topology. Here \(J_t(=\partial_{\xi}Y_t(\xi))\) is the Jacobian (derivative) process of \(Y_t\) and \(I^m_t\) is a certain weighted sum process of Wiener chaos of order 2 defined by fBm \(B_t\). Note that the weak convergence of \(\{(2^m)^{2H-\frac{1}{2}}I^m_t\}\) can be proved by using the fourth moment theorem. Their limit theorem of the error \(\hat{Y}^{m}_t-Y_t\) is the first result for solutions of multidimensional RDEs with the Hurst parameter \(\frac{1}{3}<H<\frac{1}{2}\). We are interested in the difference \(R^m_t=\hat{Y}^{m}_t-Y_t-J_tI^m_t\). The convergence results of \(\{(2^m)^{2H-\frac{1}{2}}(\hat{Y}^{m}_t-Y_t)\}\) and \(\{(2^m)^{2H-\frac{1}{2}}I^m_t\}\) suggests that \(R^m_t\) might be a small term in a certain sense as \(m\to\infty\). Conversely, if one can prove \(\lim_{m\to\infty}E[(2^m)^{2H-\frac{1}{2}}\sup_{0\leq t\leq 1}|R^m_t|]=0\), then the weak convergence of \(\{(2^m)^{2H-\frac{1}{2}}J_tI^m_t\}\) immediately implies the weak convergence of \(\{(2^m)^{2H-\frac{1}{2}}(\hat{Y}^{m}_t-Y_t)\}\) to the same limit distribution.

In this paper, in the case of fBm, for the four schemes mentioned above, we prove that
\((2^m)^{2H-\frac{1}{2}+\varepsilon}\sup_{0\leq t\leq 1}|R^m_t|\) converges to 0 almost surely and in \(L^p\) for all \(p\ge 1\). Here \(0<\varepsilon<3H-1\) is an arbitrary constant. This is one of our main theorems (Theorem 15). Our proof of this result does not rely on the weak convergence of \(\{(2^m)^{2H-\frac{1}{2}}I^m_t\}\) but the uniform \(L^p\) estimate of the Hölder norm of \(\{(2^m)^{2H-\frac{1}{2}}I^m_t\}\) independent of \(m\). Our result shows that the remainder term \(R^m_t\) is really small compared to the term \(J_tI^m_t\) and that it suffices to establish the limit theorem of weighted sum process of Wiener chaos to obtain a limit theorem of the error of \(\hat{Y}^{m}_t-Y_t\) in certain cases. In addition, we can give an estimate of the convergence rate of \(\sup_{0\le t\le 1}|\hat{Y}^{m}_t-Y_t|\to 0\) in \(L^p\) sense (see Remark 16). To the best of the authors’ knowledge, \(L^p\) convergence rate does not appear in the literature concerning fBm with the Hurst parameter \(\frac{1}{3}<H<\frac{1}{2}\).

Our idea to obtain the estimate of \(R^m_t\) is as follows. The approximate solutions considered in this paper are essentially defined at the discrete times \(D_m\). We denote the solution and approximate solution at the discrete times \(D_m\) by \(\{Y_t\}_{t\in D_m}\) and \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) respectively. We note that all four schemes are given by similar recurrence relations. More precisely, the recurrence relations of three schemes, implementable Milstein, Crank-Nicolson and first-order Euler schemes, can be obtained by adding extra two terms containing \(d^m\) and \(\hat{\epsilon}^{m}\) to the recurrence relation of the Milstein scheme as we will see in 21 . Based on this observation, we introduce an interpolation process \(\{Y^{m,\rho}_t\}_{t\in D_m}\) which is parameterized by \(\rho\in [0,1]\) and satisfies \(Y^{m,0}_t=Y_t\) and \(Y^{m,1}_t=\hat{Y}^{m}_t\) for all \(t\in D_m\). Note that \(Y^{m,\rho}_t\) is different from the standard linear interpolation \((1-\rho)Y_t+\rho\hat{Y}^{m}_t\). We define \(\{Y^{m,\rho}_t\}_{t\in D_m}\) by 46 . Let \(Z^{m,\rho}_t=\partial_{\rho}Y^{m,\rho}_t\). We can represent the process \(\{Z^{m,\rho}_t\}_{t\in D_m}\) by a constant variation method by using a certain matrix valued process \(\{\tilde{J}^{m,\rho}_t\}_{t\in D_m}\) which approximates the derivative process \(J_t\). The important point is that all processes \(\{(Y^{m,\rho}_t, Z^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1})\}_{t\in D_m}\) are solutions of certain discrete RDEs and we can get good estimates of them. We study the error process by the estimates and the expression \(\hat{Y}^{m}_t-Y_t=\int_0^1Z^{m,\rho}_t d\rho\). More precisely, we show that the main part of the right-hand side of this identity is given by \(J_tI^m_t\) and prove our main theorems.

We revisit Liu-Tindel’s result [7]. They also obtained an expression of \(\hat{Y}^{m}_t-Y_t\) by using the process \(\Phi^m_t\) which also approximates \(J_t\). See Lemma 6.4 in [7]. Their proof for the convergence of \(\{(2^m)^{2H-\frac{1}{2}}(\hat{Y}^{m}_t-Y_t)\}\) is based on the expression. The process \(\Phi^m_t\) is defined by using the standard linear interpolation process \((1-\rho)Y_t+\rho\hat{Y}^{m}_t\) and \(\Phi^m_t\) is different from our \(\tilde{J}^{m,\rho}_t\). For the sake of conciseness of the paper, they did not get into the detailed study of the integrability of \(\Phi^m_t\) but they believed the integrability of it and its inverse. Hence they could provide only the almost sure convergence rate of \(\sup_{0\le t\le 1}|\hat{Y}^{m}_t-Y_t|\to 0\), but not the \(L^p\) convergence rate. One may prove the integrabilities, but, we introduce different kind of interpolation process \(Y^{m,\rho}_t\) and prove the integrability of \(\tilde{J}^{m,\rho}\) to obtain our main results including the \(L^p\) convergence rate.

We now explain how to implement our idea mentioned above. In fact, Theorem 15 is deduced from more general results (Theorem 12 and Corollary 1). As we already explained, the recurrence relations of the three schemes contain extra terms containing \(d^m\) and \(\hat{\epsilon}^{m}\), which are not contained in the recurrence relation of the Milstein scheme. Recall that the Milstein approximation solution converges to the solution in pathwise sense in [10], [11]. Hence we expect that if these extra terms are sufficiently small in a certain sense then the approximate solutions converge to the solution, not to mention the case of the four schemes. In Theorem 12, we are concerned with such more general approximate solutions and general driving Gaussian processes and provide estimates of the errors at discrete times \(D_m\). More precisely, in such a setting, we give the estimate of the remainder term \(R^m_t\) \((t\in D_m)\) under Conditions 3 and 8\(\sim\)11. Condition 3 is a natural condition on the covariance of the driving Gaussian process \(B\) which ensures that \(B\) can be lifted to a geometric rough path. The other conditions are smallness conditions on \(d^m\) and \(\hat{\epsilon}^{m}\). The main non-trivial condition among them is Condition 10 on \(I^m\), that is, the uniform estimate of the \(L^p\) norm of the Hölder norm of \((2^m)^{2H-\frac{1}{2}}I^m\) independent of \(m\). In the case of the implementable Milstein, Milstein, and first-order Euler schemes whose driving process is an fBm, all conditions can be checked. Hence, after establishing the continuous time version of Theorem 12, in Corollary 1, Theorem 15 for the three schemes follows from these results. In the case of the Crank-Nicolson scheme, some of the conditions are not satisfied, so Theorem 15 requires additional arguments to be established. Here we mention how to show that Conditions 8\(\sim\)11 are satisfied. These conditions can be checked for the four schemes (as mentioned above, only partially, in the case of Crank-Nicolson scheme) whose driving process is an fBm by using the previously known results, e.g., in [7]. We can also prove that these conditions hold by a different idea based on the Malliavin calculus and estimates for multidimensional Young integrals although we need more smoothness assumption on \(\sigma\) and \(b\) to prove Condition 10 than the previous study in [7]. To make the paper reasonable size, we study these problems in a separate paper [12].

This paper is organized as follows. In Section 2, we recall basic notions and estimates of rough path analysis and the definition of the typical four schemes. We next state our main theorems and make remarks on them. After that we prove Theorem 15 assuming Theorem 12 and Corollary 1. We close this section by introducing notion of small order nice discrete process which includes the process of \(d^m\) and \(\hat{\epsilon}^{m}\) as examples. The estimates of discrete Young integrals with respect to these processes play an important role in this study. In Section 3, we introduce processes \(\{(Y^{m,\rho}_t, Z^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1})\}\) and put the list of notations which we will use in this paper. In Section 4, we give estimates for \(\{(Y^{m,\rho}_t, Z^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1})\}\) by using Davie’s argument in [10]. We next give \(L^p\) estimates for \(\tilde{J}^{m,\rho}_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}\) by using the estimate of Cass-Litterer-Lyons [13]. Thanks to this integrability, we can obtain good enough estimates of several quantities to prove our main theorems. In Section 5, we give a more precise estimate of \(\{Z^{m,\rho}_t\}\). In the final part of this section, we give proofs of Theorem 12 and Corollary 1.

2 Main results, remarks, and preliminaries↩︎

This section begins with a collection of the notation that will be used later. Throughout this paper, \(m\) denotes a positive integer. Set \(\Delta_m=2^{-m}\) and \(\tau^m_k=k 2^{-m}\) (\(0\le k\le 2^m\)) and write \(D_m=\{\tau^m_k\}_{k=0}^{2^m}\) for the dyadic partition of \([0,1]\). We identify the set of partition points and the partition. The standard basis of \({\mathbb{R}}^d\) is denoted by \(\{e_{\alpha}\}_{\alpha=1}^d\) and \(\lfloor x \rfloor=\max\{n\in \mathbb{Z}~|~n\le x\}\) for \(x\geq 0\).

Let us consider a process \(F=\{F_t\}_{t\in I}\) for \(I=[0,1]\) or \(D_m\). We say that \(F\) is a discrete process if \(I=D_m\), namely \(F_t\) is evaluated at \(t\in D_m\). We write \(F_{s,t}=F_t-F_s\) for \(s<t\) and, for \(0<\theta<1\), define the (discrete) \(\theta\)-Hölder norm by \[\begin{align} \label{def32Holder32norm} \|F\|_\theta =\max_{s, t\in I, s<t}\frac{|F_{s,t}|}{|t-s|^\theta}. \end{align}\tag{1}\] For two-parameter functions \(F=\{F_{s,t}\}_{s<t}\), we define the \(\theta\)-Hölder norm in the same way. In addition, the Hölder norm of \(F\) on the interval \(J\subset I\) is denoted by \(\|F\|_{J,\theta}\).

When we are given a sequence of random variables \(\{\eta_{\tau^m_{i-1},\tau^m_i}\}_{i=1}^{2^m}\), we define a discrete stochastic process \(\{\eta_t\}_{t\in D_m}\) and its increment process \(\{\eta_{s,t}\}_{s\le t, s,t\in D_m}\) by \[\begin{align} \eta_t &= \sum_{i=1}^{2^mt} \eta_{\tau^m_{i-1},\tau^m_i},\quad & \eta_{s,t} &= \eta_t-\eta_s \label{eq89034802914832} \end{align}\tag{2}\] with the convention \(\eta_0=0\). In our study, such an \(\{\eta_{\tau^m_{i-1},\tau^m_i}\}\) arises as a small increment in the time interval \([\tau^m_{i-1},\tau^m_i]\).

The remainder of this section is structured as follows. In Section 2.1, we recall basic notion in rough path analysis and introduce a condition (Condition 3) on the covariance of the driving Gaussian process \(B\) under which \(B\) can be lifted to a rough path. We next introduce the small remainder term \(\epsilon^m_{\tau^m_{k-1},\tau^m_k}\) of the solution. In Section 2.2, we explain four approximation schemes of RDE and introduce two important quantities \(d^m_{\tau^m_{k-1},\tau^m_k}\) which belongs to Wiener chaos of order 2 and \(\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}\) which is defined as a small remainder term of approximate solution similarly to \(\epsilon^m_{\tau^m_{k-1},\tau^m_k}\). We next explain that the approximation equations can be written as common recurrence equations using \(d^m_{\tau^m_{k-1},\tau^m_k}\) and \(\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}\). This observation is important for our study. In Section 2.3, taking the common recurrence equations into account, we consider more general approximation equations. We next introduce Conditions 8\(\sim\)11 on \(d^m\), \(\hat{\epsilon}^{m}\) and iterated integrals of \(B\) and state our main theorems (Theorem 12, Corollary 1, and Theorem 15). In Section 2.4, we show Theorem 15 in the case of the implementable Milstein, Crank-Nicolson, Milstein and first-order Euler schemes, assuming Theorem 12 and Corollary 1. In Section 2.5, we define a class of discrete processes, small order nice discrete processes, which includes \(d^m, \epsilon^m, \hat{\epsilon}^{m}\).

2.1 Rough paths and solutions to RDEs↩︎

Here we recall some basic notions of rough path analysis. For details, see [11], [14], [15].

Let \(\frac{1}{3}<\theta\le \frac{1}{2}\). Let \(X=\{X_{s,t}\}_{0\leq s<t\leq 1}\) and \({\mathbb{X}}=\{{\mathbb{X}}_{s,t}\}_{0\leq s<t\leq 1}\) be two-parameter functions with values in \({\mathbb{R}}^d\) and \({\mathbb{R}}^d\otimes{\mathbb{R}}^d\), respectively.

Definition 1.

  1. We say that the pair \(\boldsymbol{X}=(X,{\mathbb{X}})\) is a \(\theta\)-Hölder rough path if \(\|X\|_{\theta}<\infty\), \(\|{\mathbb{X}}\|_{2\theta}<\infty\) and \(X_{s,t}=X_{s,u}+X_{u,t}\), \({\mathbb{X}}_{s,t}={\mathbb{X}}_{s,u}+{\mathbb{X}}_{u,t}+X_{s,u}\otimes X_{u,t}\) for \(0\leq s<u<t\leq 1\) (Chen’s identity).

  2. We say that a \(\theta\)-Hölder rough path \(\boldsymbol{X}=(X,{\mathbb{X}})\) is geometric if it satisfies the following: there exists a sequence of smooth paths \(X^m\) such that its natural lift \(\boldsymbol{X}^m=(X^m,{\mathbb{X}}^m)\), where \({\mathbb{X}}^m_{s,t}=\int_s^t X^m_{s,u}\otimes dX^m_{0,u}\), approximates \(\boldsymbol{X}=(X,{\mathbb{X}})\) in the rough path metric, that is, \[\begin{align} \lim_{m\to\infty} \{\|X-X^m\|_{\theta}+\|{\mathbb{X}}-{\mathbb{X}}^m\|_{2\theta}\} = 0. \end{align}\] We denote by \(\mathscr{C}^\theta_g\) the set of all \(\theta\)-Hölder geometric rough paths.

We denote by \(X^\alpha_{s,t}\) the \(e_\alpha\)-component of \(X_{s,t}\) and by \(X^{\alpha,\beta}_{s,t}\) the \(e_\alpha\otimes e_\beta\)-component of \({\mathbb{X}}_{s,t}\). Namely we write \(X_{s,t}=\sum_{\alpha=1}^d X^\alpha_{s,t}e_{\alpha}\) and \({\mathbb{X}}_{s,t}=\sum_{1\le \alpha,\beta\le d}X^{\alpha,\beta}_{s,t}e_{\alpha} \otimes e_{\beta}\). Recall that we can construct the third level rough paths from the first and second level rough paths. The \(e_{\alpha}\otimes e_{\beta}\otimes e_{\gamma}\)-component of the third level rough paths will be denoted by \(X^{\alpha,\beta,\gamma}_{s,t}\).

Next we introduce the notion of controlled paths and integration of controlled paths.

Definition 2. Let \(X=\{X_t\}_{0\leq t\leq 1}\) be a \(\theta\)-Hölder function with values in \({\mathbb{R}}^d\). A \(\theta\)-Hölder function \(Z=\{Z_t\}_{0\leq t\leq 1}\) with values in \({\mathbb{R}}^K\) is said to be a path controlled by \(X\) if there exist a \(\theta\)-Hölder function \(Z'=\{Z'_t\}_{0\leq t\leq 1}\) with valued in \(\mathcal{L}({\mathbb{R}}^d, {\mathbb{R}}^K)\) and a \((2\theta)\)-Hölder function \(R=\{R_{s,t}\}_{0\leq s<t\leq 1}\) satisfying \(Z_t-Z_s = Z'_s (X_t-X_s)+R_{s,t}\) \((0\leq s<t\leq 1)\). The set of all pairs \((Z,Z')\) is denoted by \(\mathscr{D}_X^{2\theta}([0,1],{\mathbb{R}}^K)\).

Let \(\boldsymbol{X}=(X,{\mathbb{X}})\) be a geometric \(\theta\)-Hölder rough path and identify \(X\) with a one-parameter function by \(X_t=X_{0,t}\). We can define an integration of a path \((Z,Z')\) controlled by \(X\) against \(\boldsymbol{X}=(X,{\mathbb{X}})\) as follows.

Theorem 1 ([14]). Let \((Z,Z')\in \mathscr{D}_X^{2\theta}([0,1],\mathcal{L}({\mathbb{R}}^d, {\mathbb{R}}^K))\). We can define an integration of \((Z,Z')\) along \(\boldsymbol{X}=(X,{\mathbb{X}})\) by \[\begin{align} \int_s^t Z_u d\boldsymbol{X}_u = \lim_{|\mathcal{P}|\to 0} \sum_{i=1}^M \{ Z_{t_{i-1}} X_{t_{i-1},t_i} +Z'_{t_{i-1}} {\mathbb{X}}_{t_{i-1},t_i} \}. \end{align}\] Here \(\mathcal{P}=\{t_i\}_{i=0}^M\) denotes a partition of the interval \([s,t]\) and \(|\mathcal{P}|=\max\{t_i-t_{i-1}|1\leq i\leq M\}\). We call the left-hand side a rough integral.

Let \(P_{\alpha}\) be the projection operator on \({\mathbb{R}}^d\) onto the subspace spanned by \(e_{\alpha}\). Then \(\int_0^tZ_ud\boldsymbol{X}_u=\sum_{\alpha=1}^d\int_0^tZ_uP_{\alpha}d\boldsymbol{X}_u\) holds. We may write \(\int_0^tZ_uP_{\alpha}d\boldsymbol{X}_u=\int_0^tZ_ue_{\alpha}dX^{\alpha}_u\). Actually, the rough integral \(\int_0^t \tilde{Z}_ud\bar{Z}_u\) can be defined for any paths \(\tilde{Z}_t, \bar{Z}_t\) controlled by \(X\) (see [14]). Also note that \(Z_te_{\alpha}\) and \(X^{\alpha}_t\) are \(\theta\)-Hölder paths controlled by \(X\). It is easy to check that \(\int_0^tZ_ue_{\alpha}dX^{\alpha}_u\) coincide with the rough integral in that sense. Note that the process \(\big\{ \big( \int_0^t Z_u d\boldsymbol{X}_u, Z_t \big) \big\}_{0\leq t\leq 1}\) is also a path controlled by \(X\) and we can define iterated integrals in the sense of rough integrals. Furthermore, we have the following formula: for any \(f\in C^3_b({\mathbb{R}}^K,{\mathbb{R}}^L)\), \[\begin{align} \label{eq483290842} f(Z_t)-f(Z_s) = \int_0^t (Df)(Z_u)Z'_ud\boldsymbol{X}_u +\int_0^t (Df)(Z_u)d\Gamma_u \end{align}\tag{3}\] if \((Z,Z')\in \mathscr{D}_X^{2\theta}([0,1],{\mathbb{R}}^K)\) satisfies \(Z_t=Z_0+\int_s^t Z'_u d\boldsymbol{X}_u+\Gamma_t\) for some \((Z',Z'')\in \mathscr{D}_X^{2\theta}([0,1],\mathcal{L}({\mathbb{R}}^d, {\mathbb{R}}^K))\) and smooth function \(\Gamma\) with values in \({\mathbb{R}}^K\). For detail, see [14].

Next we introduce the notion of solutions to RDEs. Let \(\xi\in{\mathbb{R}}^n\), \(\sigma\in C^4_b({\mathbb{R}}^{n}, \mathcal{L}({\mathbb{R}}^d, {\mathbb{R}}^{n}))\), \(b\in C^2_b({\mathbb{R}}^{n},{\mathbb{R}}^{n})\) and consider an RDE driven by \(X\) on \({\mathbb{R}}^{n}\), \[\begin{align} Y_t = \xi +\int_0^t\sigma(Y_s)d\boldsymbol{X}_s +\int_0^t b(Y_s)ds, \qquad 0\le t\le 1. \label{generalRDE} \end{align}\tag{4}\] Here the first integral should be understood as a rough integral. We also write \(Y_t(\xi,X)=Y_t\) if the solution \(Y_t\) exists. We have several notion of solution, which are equivalent. To state them, we set \[\begin{align} ((D\sigma)[\sigma])(y)[v\otimes w] = D\sigma(y)[\sigma(y)v]w, \qquad y\in{\mathbb{R}}^n, v,w\in{\mathbb{R}}^d. \label{eqSimplifiedNotationDef} \end{align}\tag{5}\] In this notation, we have \[\begin{align} \label{eqSimplifiedNotation02} ((D\sigma)[\sigma])(y){\mathbb{X}}_{s,t} = \sum_{\alpha,\beta=1}^d (D\sigma)(y)[\sigma(y)e_{\alpha}]e_{\beta} X^{\alpha,\beta}_{s,t}. \end{align}\tag{6}\]

Theorem 2 ([14]). The following are equivalent and both are valid.

  1. There exists a unique \((Y,Y')\in \mathscr{D}_X^{2\theta}([0,1],{\mathbb{R}}^n)\) satisfying 4 with \(Y'=\sigma(Y)\).

  2. There exists a unique process \(Y\colon[0,1]\to{\mathbb{R}}^n\) satisfying \[\begin{align} |Y_t-Y_s-\sigma(Y_s)X_{s,t}-((D\sigma)[\sigma])(Y_s){\mathbb{X}}_{s,t}-b(Y_s)(t-s)| \leq C(t-s)^{3\theta} \label{definition32of32davie} \end{align}\tag{7}\] for \(0\leq s<t\leq 1\). Here \(C\) can be estimated by a polynomial function of \(\|X\|_{[0,1],\theta}\) and \(\|{\mathbb{X}}\|_{[0,1],2\theta}\). This is called a solution in the sense of Davie [10].

Note that we can choose \(C\) in 7 so that it can be estimated by a polynomial function of \(\|X\|_{[s,t],\theta}\) and \(\|{\mathbb{X}}\|_{[s,t],2\theta}\). We will record this estimate in Lemma 1 later. Although the estimate on \(C\) in (7 ) and the unique existence of solution hold under weaker assumption that \(\sigma\in C^3_b\) and \(b\in C^1_b\) (see [14]), we need to assume the above condition on \(\sigma\) and \(b\) in our study.

We now introduce a condition to construct a rough path associated to a Gaussian process under which we will work. Let \(\Omega=C_0([0,1],{\mathbb{R}}^d)\) be the set of \({\mathbb{R}}^d\)-valued continuous functions on \([0,1]\) starting at the origin, \(B\) be the canonical process on \(\Omega\), that is, \(B_t(\omega)=\omega(t)\) (\(\omega\in \Omega\)), and \(\mu\) be a centered Gaussian probability measure on \(\Omega\). Throughout this paper, we put the next condition on \(B\):

Condition 3. Let \(\frac{1}{3}<H\le \frac{1}{2}\). Let \(B^{\alpha}_t\) be the \(\alpha\)-th component of \(B_t\) \((1\le \alpha\le d)\). Then \(B^1_t,\ldots, B^d_t\) are independent centered continuous Gaussian processes. Let \(R^{\alpha}(s,t)=E[B^\alpha_sB^\alpha_t]\). Then \(V_{(2H)^{-1}}(R^\alpha ; [s,t]^2 )\le C_\alpha|t-s|^{2H}\) holds for all \(1\le \alpha\le d\) and \(0\le s<t\le 1\). Here \(V_p(R^{\alpha} ; [s,t]^2)\) denotes the \(p\)-variation norm of \(R^{\alpha}\) on \([s,t]^2\).

Note that Condition 3 holds for the fBm with the Hurst parameter \(\frac{1}{3}<H\le \frac{1}{2}\).

Remark 4. It is known that under Condition 3, \(B\) can be naturally lifted to \(\boldsymbol{B}=(B,{\mathbb{B}})\in \mathscr{C}^\theta_g\) for any \(\frac{1}{3}<\theta<H\). More precisely, we can prove the following property (Remark 10.7 in [14], Theorem 15.33 in [11]). We consider a sequence of smooth rough path \(\boldsymbol{B}^m(\omega)=(B^m(\omega),{\mathbb{B}}^m(\omega))\) defined by a piecewise linear approximation of \(B(\omega)\) such that \(\lim_{m\to\infty}\max_{0\le t\le 1}|B^m_t(\omega)-B_t(\omega)|=0\) for all \(\omega\in \Omega\). Then \(\boldsymbol{B}^m(\omega)=(B^m(\omega),{\mathbb{B}}^m(\omega))\in \mathscr{C}^\theta_g\) converges in probability in the \(\theta\)-Hölder rough path metric for any \(\frac{1}{3}<\theta<H\). This implies that there exists a subset \(\Omega_0\) with \(\mu(\Omega_0)=1\) such that, if necessary choosing a subsequence, the limit \(\boldsymbol{B}(\omega)=(B(\omega),{\mathbb{B}}(\omega))\) belongs to \(\mathscr{C}^\theta_g\) for any \(\omega\in \Omega_0\) and any \(\frac{1}{3}<\theta<H\). Of course, this rough path depends on the selected versions, but, note that any versions are almost surely identical. We consider solutions to RDEs driven by this rough path obtained by Gaussian process satisfying Condition 3.

Here we fix \(\frac{1}{3}<H^-<H\). For later use, we introduce a random variable \(C(B)\) by \[\begin{align} C(B) &= \max \left\{ \|B(\omega)\|_{H^-}, \|{\mathbb{B}}(\omega)\|_{2H^-} \right\}, \qquad \omega\in \Omega_0,\label{CB} \end{align}\tag{8}\] and a subset \(\Omega^{(m)}_0\) of \(\Omega_0\) by \[\begin{align} \Omega^{(m)}_0 = \bigg\{ \omega\in \Omega_0 ~\bigg|~ \sup_{|t-s|\le 2^{-m}} \left|\frac{B_{s,t}(\omega)}{(t-s)^{H^{-}}}\right| \le \frac{1}{2},\quad \sup_{|t-s|\le 2^{-m}} \left|\frac{{\mathbb{B}}_{s,t}(\omega)}{(t-s)^{2H^{-}}}\right| \le \frac{1}{2} \bigg\}. \end{align}\] Under Condition 3, \(C(B)\in \cap_{p\ge 1}L^p\) holds. We refer the readers for this to [11], [14], [16]. Therefore, under Condition 3, we see that \[\begin{align} \mu((\Omega^{(m)}_0)^\complement)\le C_{p} 2^{-mp}\quad\quad \text{for any p>1} \label{complement32of32Omega0} \end{align}\tag{9}\] which eventually implies that the complement set is negligible for our problem. Below, we actually consider analogous subset \(\Omega^{(m, d^m)}_0\) which will be introduced in Section 2.5. The proof of \((\ref{complement32of32Omega0})\) is as follows. Let \(\kappa>0\) be a positive number satisfying \(H^-+\kappa<H\). Let \(C(B)_{H^-+\kappa}\) denote the number obtained by replacing \(H^-\) by \(H^-+\kappa\) in the definition (8 ). Then we have \[\begin{align} \sup_{|t-s|\le 2^{-m}} \left|\frac{B_{s,t}(\omega)}{(t-s)^{H^{-}}}\right|+ \sup_{|t-s|\le 2^{-m}} \left|\frac{{\mathbb{B}}_{s,t}(\omega)}{(t-s)^{2H^{-}}}\right| &\le 2^{1-m\kappa}C(B)_{H^-+\kappa}. \end{align}\] Hence we obtain \(\liminf_{m\to\infty}\Omega^{(m)}_0=\Omega_0\) and \[\mu((\Omega^{(m)}_0)^\complement)\le \mu(C(B)_{H^-+\kappa}\ge 2^{m\kappa-2}) \le 2^{-p(m\kappa-2)}\|C(B)_{H^-+\kappa}\|_{L^p}^p,\] which is the desired result.

Remark 5 (About the constants in the estimates). When a positive constant \(C\) can be written as a polynomial function of the sup-norm of some functions \(\sigma, b, c\) and their derivatives, we may say \(C\) depends on \(\sigma, b, c\) polynomially. Similarly, when a constant \(C\) can be written as a polynomial of some positive random variable \(X\), the sup-norms of \(\sigma, b, c\) and their derivatives, we say that \(C\) depends on \(\sigma, b, c, X\) polynomially. Of course the coefficients of the polynomial should not depend on \(\omega\). When \(X=C(B)\), we may denote such a constant \(C\) by \(\tilde{C}(B)\).

Throughout this paper, we assume \(B\) satisfies Condition 3 and \(\boldsymbol{B}=(B,{\mathbb{B}})\) is the canonically defined rough path as explained above. Let \(Y_t=Y_t(\xi,B)\) be the solution to RDE on \({\mathbb{R}}^n\) driven by \(B\): \[\begin{align} Y_t(\xi,B) = \xi +\int_0^t\sigma(Y_s(\xi,B))d\boldsymbol{B}_s +\int_0^t b(Y_s(\xi,B))ds, \qquad 0\le t\le 1.\label{rde} \end{align}\tag{10}\] We may omit writing the starting point \(\xi\) and the driving process \(B\) in \(Y_t(\xi,B)\). Note that \(J_t=\partial_{\xi}Y_t(\xi)\in \mathcal{L}({\mathbb{R}}^n)\) and its inverse \(J^{-1}_t\) are the solutions to the following RDEs: \[\begin{align} \tag{11} J_t&=I+\int_0^t(D\sigma)(Y_u)[J_u]d\boldsymbol{B}_u+\int_0^t(Db)(Y_u)[J_u]du,\\ \tag{12} J^{-1}_t&=I-\int_0^tJ^{-1}_u(D\sigma)(Y_u)d\boldsymbol{B}_u-\int_0^tJ^{-1}_u(Db)(Y_u)du. \end{align}\]

We conclude this section by presenting a lemma and making a remark. For every \(1\leq k\leq 2^m\), define \(\epsilon^m_{\tau^m_{k-1},t}(\xi)\) \((\tau^m_{k-1}\le t\le \tau^m_k)\) by \[\begin{align} Y_t&=Y_{\tau^m_{k-1}}+\sigma(Y_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +((D\sigma)[\sigma])(Y_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},t} +b(Y_{\tau^m_{k-1}})(t-\tau^m_{k-1}) +\epsilon^m_{\tau^m_{k-1},t}(\xi). \label{definition32of32epm} \end{align}\tag{13}\] We may use the notation \(\epsilon^m_{\tau^m_{k-1},t}\) instead of \(\epsilon^m_{\tau^m_{k-1},t}(\xi)\) for simplicity. As we explained in the inequality (7 ), we have the following.

Lemma 1.

  1. There exists a constant \(C>0\) such that \[\begin{align} \label{eq459301490123} |\epsilon^m_{\tau^m_{k-1},t}| \le C (t-\tau^m_{k-1})^{3H^-} \qquad \text{for\,\,\,all} \qquad 1\le k\le 2^m,\quad \omega\in \Omega_0. \end{align}\tag{14}\] Here \(C\) depends on \(\|B\|_{[\tau^m_{k-1},t],H^-}\), \(\|{\mathbb{B}}\|_{[\tau^m_{k-1},t],2H^-}\), \(\sigma, b\) polynomially.

  2. There exists a constant \(C>0\) depending on \(\sigma,b\) polynomially and bounded Lipschitz continuous functions \(F_{\alpha,\beta,\gamma}\), \(F^1_{\alpha}\), \(F^{2}_{\alpha}\) from \({\mathbb{R}}^n\) to \({\mathbb{R}}^n\) such that for all \(1\le k\le 2^m\) and \(\tau^m_{k-1}\le t\le \tau^m_k\), \[\begin{gather} \left| \epsilon^m_{\tau^m_{k-1},t} - \sum_{\alpha,\beta,\gamma} F_{\alpha,\beta,\gamma}(Y_{\tau^m_{k-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{k-1},t} - \sum_{\alpha} F^1_{\alpha}(Y_{\tau^m_{k-1}}) B^{0,\alpha}_{\tau^m_{k-1},t} - \sum_{\alpha} F^2_{\alpha}(Y_{\tau^m_{k-1}}) B^{\alpha,0}_{\tau^m_{k-1},t} \right|\\ \le C (t-\tau^m_{k-1})^{4H^-}, \qquad \qquad \omega\in \Omega^{(m)}_0,\label{eq8940289042} \end{gather}\tag{15}\] where \[\begin{align} B^{0,\alpha}_{\tau^m_{k-1},t}=\int_{\tau^m_{k-1}}^t(s-\tau^m_{k-1})dB^{\alpha}_s,\quad\quad B^{\alpha,0}_{\tau^m_{k-1},t}=\int_{\tau^m_{k-1}}^tB^{\alpha}_{\tau^m_{k-1},s}ds.\label{B0alpha} \end{align}\tag{16}\]

Proof. We need only to prove 15 . First we give an expression of \(\epsilon^m_{\tau^m_{k-1},t}\). Note that the solution \(Y_t\) to 10 satisfies \((Y,\sigma(Y)) \in \mathscr{D}_X^{2\theta}([0,1],{\mathbb{R}}^n)\) and \((\sigma(Y),((D\sigma)[\sigma])(Y))\in \mathscr{D}_X^{2\theta}([0,1],\mathcal{L}({\mathbb{R}}^d, {\mathbb{R}}^n))\). Hence we can use 3 . Then by applying the formula to \(f(Y_t)-(Y_{\tau^m_{k-1}})\) for \(f\in C^3_b({\mathbb{R}}^n,{\mathbb{R}}^L)\) successively, we can decompose \(\epsilon^m_{\tau^m_{k-1},t}\) in the following way. This calculation is possible because \(\sigma\in C^4_b, b\in C^2_b\). We need the following functions to state it: \[\begin{align}F^0(y)&=(Db)(y)[b(y)], & F^1_{\alpha}(y)&=(Db)(y)[\sigma(y)e_{\alpha}], & F^2_{\alpha}(y)&=(D\sigma(y)e_{\alpha})[b(y)], \\ \begin{align} F_{\alpha,\beta,\gamma}(y) &= D\Bigl\{(D\sigma(y)e_{\gamma})[\sigma(y)e_{\beta}]\Bigr\} [\sigma(y)e_{\alpha}], & G_{\alpha,\beta}(y) &= D\Bigl\{(D\sigma(y)e_{\beta})[\sigma(y)e_{\alpha}]\Bigr\}[b(y)]. \end{align} \end{align}\] The decomposition formula is as follows, \[\begin{align} &\epsilon^m_{\tau^m_{k-1},t} = \sum_{\alpha,\beta,\gamma} \int_{\tau^m_{k-1}}^{t}\left\{\int_{\tau^m_{k-1}}^{s}\left(\int_{\tau^m_{k-1}}^{u} F_{\alpha,\beta,\gamma}(Y_v) dB^{\alpha}_v\right)dB^{\beta}_u\right\} dB^{\gamma}_s\nonumber\\ &\quad +\sum_{\alpha,\beta,\gamma} \int_{\tau^m_{k-1}}^{t}\left\{\int_{\tau^m_{k-1}}^{s}\left(\int_{\tau^m_{k-1}}^{u} G_{\alpha,\beta}(Y_v) dv\right)dB^{\alpha}_u\right\} dB^{\beta}_s +\int_{\tau^m_{k-1}}^{t}\left(\int_{\tau^m_{k-1}}^sF^0(Y_u)du\right)ds\nonumber\\ &\quad +\sum_{\alpha}\int_{\tau^m_{k-1}}^{t}\left( \int_{\tau^m_{k-1}}^sF^1_{\alpha}(Y_u)du\right)dB^{\alpha}_s +\sum_{\alpha}\int_{\tau^m_{k-1}}^{t}\left( \int_{\tau^m_{k-1}}^sF^2_{\alpha}(Y_u)dB^{\alpha}_u\right)ds\nonumber\\ &:=I_1+\cdots+I_5. \label{explicit32form32of32epm} \end{align}\tag{17}\]

By using estimates of rough integrals, we have the following estimates: for all \(\omega\in\Omega^{(m)}_0\), it holds that \[\begin{gather} \Bigl|I_1-\sum_{\alpha,\beta,\gamma}F_{\alpha,\beta,\gamma}(Y_{\tau^m_{k-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{k-1},t}\Bigr| \le C(t-\tau^m_{k-1})^{4H^-},\\ \left|I_2\right|\le C(t-\tau^m_{k-1})^{1+2H^-}, \qquad \qquad \quad \left|I_3\right|\le C(t-\tau^m_{k-1})^2, \\ \left|I_4-F^1_{\alpha}(Y_{\tau^m_{k-1}})B^{0,\alpha}_{\tau^m_{k-1},t}\right| +\left|I_5-F^2_{\alpha}(Y_{\tau^m_{k-1}})B^{\alpha,0}_{\tau^m_{k-1},t}\right|\le C(t-\tau^m_{k-1})^{1+2H^-},\label{I432I5} \end{gather}\tag{18}\] where \(C\) depends on \(\sigma\) and \(b\) polynomially. This completes the proof. ◻

Remark 6. For every \(s,t\in D_m\) with \(s\le t\), define \(\epsilon^m_t\) and \(\epsilon^m_{s,t}\) in the same way as 2 with \(\eta_{\tau^m_{i-1},\tau^m_i}=\epsilon^m_{\tau^m_{i-1},\tau^m_i}\). Note that the identity \(\epsilon^m_{s,t} = Y_t-Y_s-\sigma(Y_s)B_{s,t} -((D\sigma)[\sigma])(Y_s){\mathbb{B}}_{s,t} -b(Y_s)(t-s)\) does not hold for general \(s,t\in D_m\) with \(s\le t\).

2.2 Four approximation schemes↩︎

In this section, we introduce typical four approximation schemes. That is, we introduce the implementable Milstein approximate solution \(Y^{\mathrm{IM},m}_t\), the Milstein approximate solution \(Y^{\mathrm{M},m}_t\), the first-order Euler approximate solution \(Y^{\mathrm{FE},m}\), and the Crank-Nicolson approximate solution \(Y^{\mathrm{CN},m}_t\) associated to the dyadic partition \(D_m\). The first three schemes are explicit scheme and defined inductively as follows: \(Y^{\mathrm{IM},m}_0=Y^{\mathrm{M},m}_0=Y^{\mathrm{FE},m}_0=\xi\) and \[\begin{align} Y^{\mathrm{IM},m}_t &= Y^{\mathrm{IM},m}_{\tau^m_{k-1}} +\sigma(Y^{\mathrm{IM},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +((D\sigma)[\sigma])(Y^{\mathrm{IM},m}_{\tau^m_{k-1}}) \left[ \frac{1}{2} B_{\tau^m_{k-1},t} \otimes B_{\tau^m_{k-1},t} \right]\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad +b(Y^{\mathrm{IM},m}_{\tau^m_{k-1}})(t-\tau^m_{k-1}),\\ Y^{\mathrm{M},m}_t &= Y^{\mathrm{M},m}_{\tau^m_{k-1}} +\sigma(Y^{\mathrm{M},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +((D\sigma)[\sigma])(Y^{\mathrm{M},m}_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},t} +b(Y^{\mathrm{M},m}_{\tau^m_{k-1}})(t-\tau^m_{k-1}),\\ Y^{\mathrm{FE},m}_{t} &= Y^{\mathrm{FE},m}_{\tau^m_{k-1}} +\sigma(Y^{\mathrm{FE},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +((D\sigma)[\sigma])(Y^{\mathrm{FE},m}_{\tau^m_{k-1}}) \left[ \frac{1}{2} \sum_{\alpha=1}^d e_{\alpha} \otimes e_{\alpha} E[(B_{\tau^m_{k-1},t}^\alpha)^2] \right]\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad +b(Y^{\mathrm{FE},m}_{\tau^m_{k-1}})(t-\tau^m_{k-1}), \end{align}\] for every \(\tau^m_{k-1}< t\le\tau^m_k\) and \(1\le k\le 2^m\). In the above, we omit writing the initial value \(\xi\) for the solution. With the notation 5 , we have \[\begin{align} ((D\sigma)[\sigma])(y) \left[ \frac{1}{2} B_{s,t} \otimes B_{s,t} \right] &= \sum_{\alpha,\beta=1}^d \frac{1}{2}(D\sigma)(y)[\sigma(y)e_{\alpha}]e_{\beta} B^{\alpha}_{s,t} B^{\beta}_{s,t},\\ \label{eqSimplifiedNotation03} ((D\sigma)[\sigma])(y) \left[ \frac{1}{2} \sum_{\alpha=1}^d e_{\alpha} \otimes e_{\alpha} E[(B_{s,t}^\alpha)^2] \right] &= \sum_{\alpha=1}^d \frac{1}{2} (D\sigma)(y) [\sigma(y)e_{\alpha}]e_{\alpha}E[(B_{s,t}^\alpha)^2]. \end{align}\tag{19}\]

Next we introduce the Crank-Nicolson scheme. Since the Crank-Nicolson scheme is an implicit scheme and an equation stated later with respect to \(Y^{\mathrm{CN},m}_t\) must be solvable. For that purpose, we already introduced the set \(\Omega^{(m)}_0\). Since \(D\sigma\) and \(Db\) are bounded function, the mapping \[v \mapsto \eta +\frac{1}{2}\left(\sigma(\eta)+\sigma(v)\right)B_{\tau^m_{k-1},t} +\frac{1}{2}\left(b(\eta)+b(v)\right)(t-\tau^m_{k-1}), \qquad \tau^m_{k-1}\leq t\leq \tau^m_k,\] is a contraction mapping for any \(\eta\in{\mathbb{R}}^n\) and \(\omega\in \Omega^{(m)}_0\) for large \(m\). Therefore, for \(\omega\in \Omega^{(m)}_0\) for large \(m\), the Crank-Nicolson scheme \(Y^{\mathrm{CN},m}_t\) is uniquely defined as the following inductive equation: \(Y^{\mathrm{CN},m}_0=\xi\) and \[\begin{align} Y^{\mathrm{CN},m}_t &= Y^{\mathrm{CN},m}_{\tau^m_{k-1}} +\frac{1}{2}\left(\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})+\sigma(Y^{\mathrm{CN},m}_t)\right)B_{\tau^m_{k-1},t} \notag \\ &\qquad\qquad\qquad\qquad\qquad +\frac{1}{2}\left(b(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})+b(Y^{\mathrm{CN},m}_t)\right)(t-\tau^m_{k-1}) \label{CN32equation} \end{align}\tag{20}\] for every \(\tau^m_{k-1}< t\le\tau^m_k\) and \(1\le k\le 2^m\). For the completeness of definition, we set \(Y^{\mathrm{CN},m}_t\equiv \xi\) for \(\omega\in \Omega_0\setminus \Omega^{(m)}_0\).

In what follows, we discuss how to address the four schemes collectively. This is one of the key ingredients of this paper. We use the common notation \(\{\hat{Y}^m_t\}_{t\in [0,1]}\) to denote these four approximate solutions. The four approximate solutions \(\{\hat{Y}^m_t\}_{t\in [0,1]}\) also satisfy similar but a little bit different equations to (13 ). Indeed, by choosing a function \(c\in C^3_b({\mathbb{R}}^n, L({\mathbb{R}}^d\otimes {\mathbb{R}}^d,{\mathbb{R}}^n))\) and random variables \(d^m=\{d^m_{\tau^m_{k-1},t}\}_{1\leq k\leq 2^m,\tau^m_{k-1}<t\le \tau^m_k}\subset {\mathbb{R}}^d\otimes {\mathbb{R}}^d\) and \(\hat{\epsilon}^{m}(\xi)=\{\hat{\epsilon}^{m}_{\tau^m_{k-1},t}(\xi)\}_{1\leq k\leq 2^m,\tau^m_{k-1}<t\le \tau^m_k}\subset {\mathbb{R}}^n\) defined on \(\Omega_0\), these approximate equations can be written as the following common form on \(\Omega_0\): \(\hat{Y}^{m}_0=\xi\) and \[\begin{align} \hat{Y}^{m}_t &= \hat{Y}^{m}_{\tau^m_{k-1}}+\sigma(\hat{Y}^{m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +((D\sigma)[\sigma])(\hat{Y}^{m}_{\tau^m_k}){\mathbb{B}}_{\tau^m_{k-1},t} +b(\hat{Y}^{m}_{\tau^m_{k-1}})(t-\tau^m_{k-1})\nonumber\\ &\qquad\qquad\qquad\qquad +c(\hat{Y}^{m}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},t} +\hat{\epsilon}^m_{\tau^m_{k-1},t}(\xi), \quad \tau^m_{k-1}<t\le \tau^m_k.\label{approximate32solution} \end{align}\tag{21}\]

We explain more precisely what \(c, d^m, \hat{\epsilon}^{m}(\xi)\) are for all cases. In all cases, \(c\) is given by \[\begin{align} c(y)[v\otimes w] = ((D\sigma)[\sigma])(y)[v\otimes w] = D\sigma(y)[\sigma(y)v]w, \qquad y\in{\mathbb{R}}^n, v,w\in{\mathbb{R}}^d. \end{align}\] and \(d^m_{\tau^m_{k-1},t}\) arises from the difference between the second level rough paths and their approximations in each scheme. Furthermore, \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}(\xi)\) denotes a smaller term in each scheme. We may use the notation \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}\) for \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}(\xi)\) if there is no confusion. For \(Y^{\mathrm{IM},m}\), \(Y^{\mathrm{M},m}\) and \(Y^{\mathrm{FE},m}\), the pairs of \(d^m\) and \(\hat{\epsilon}^{m}\) are given by \[\begin{align} d^{\mathrm{IM},m}_{\tau^m_{k-1},t} &= \frac{1}{2} B_{\tau^m_{k-1},t} \otimes B_{\tau^m_{k-1},t} - {\mathbb{B}}_{\tau^m_{k-1},t}, & \hat{\epsilon}^{\mathrm{IM},m}_{\tau^m_{k-1},t} &= 0,\\ d^{\mathrm{M},m}_{\tau^m_{k-1},t} &= 0, & \hat{\epsilon}^{\mathrm{M},m}_{\tau^m_{k-1},t} &= 0,\\ d^{\mathrm{FE},m}_{\tau^m_{k-1},t} &= \frac{1}{2} \sum_{\alpha=1}^d e_{\alpha} \otimes e_{\alpha} E[(B^\alpha_{\tau^m_{k-1},t})^2] - {\mathbb{B}}_{\tau^m_{k-1},t}, & \hat{\epsilon}^{\mathrm{FE},m}_{\tau^m_{k-1},t} &= 0. \end{align}\]

The Crank-Nicolson scheme leads to a slightly complicated situation. For the Crank-Nicolson scheme \(Y^{\mathrm{CN},m}\), we set \(d^{\mathrm{CN},m}_{\tau^m_{k-1},t} = \frac{1}{2} B_{\tau^m_{k-1},t} \otimes B_{\tau^m_{k-1},t} - {\mathbb{B}}_{\tau^m_{k-1},t}\), that is, the same one as the case of implementable Milstein scheme. Once \(d^{\mathrm{CN},m}_{\tau^m_{k-1},t}\) is defined, \(\hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}\) is automatically determined by the identity 21 . For \(\omega\in\Omega_0\setminus \Omega_0^{(m)}\), from \(Y^{\mathrm{CN},m}_t=\xi\) and 21 , we easily see \[\begin{align} \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t} = -\sigma(\xi)B_{\tau^m_{k-1},t} -((D\sigma)[\sigma])(\xi){\mathbb{B}}_{\tau^m_{k-1},t} -c(\xi)d^{\mathrm{CN},m}_{\tau^m_{k-1},t} -b(\xi)(t-\tau^m_{k-1}). \label{definition32of32hepm32for32CN} \end{align}\tag{22}\] For \(\omega\in\Omega_0^{(m)}\), we set \[\begin{align} \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t} &= \frac{1}{2} \bigg( \int_0^1 \bigg( (D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}]\nonumber\\ &\qquad\qquad\qquad -(D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})[\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t}]\bigg)d\theta \bigg) B_{\tau^m_{k-1},t}\nonumber\\ &\qquad\qquad + \frac{1}{2} \left( \int_0^1(Db) (Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}] d\theta \right) (t-\tau^m_{k-1}). \label{explicit32form32of32hepmCN} \end{align}\tag{23}\] Then we see that the recurrence relation 21 holds and that \(\hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}\) admits good estimates as follows.

Lemma 2. Let \(\omega\in \Omega_0^{(m)}\).

  1. The Crank-Nicolson approximate solution satisfies 21 with \(d^m_{\tau^m_{k-1},t}=d^{\mathrm{CN},m}_{\tau^m_{k-1},t}\) and \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}=\hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}\).

  2. There exists a positive constant \(C\) such that \[\begin{align} |\hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}| \le C|t-\tau^m_{k-1}|^{3H^-} \qquad \text{for all} \qquad 1\le k\le 2^m. \end{align}\] Here, \(C\) depends on \(\sigma\) and \(b\) polynomially.

  3. There exist bounded Lipschitz continuous functions \(\varphi_{\alpha,\beta,\gamma} :{\mathbb{R}}^n\to{\mathbb{R}}^n\) and \(\psi_{\alpha} :{\mathbb{R}}^n\to{\mathbb{R}}^n\) such that \[\begin{gather} \Big| \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},\tau^m_k} - \sum_{1\le \alpha,\beta,\gamma\le d} \varphi_{\alpha,\beta,\gamma}(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{k-1},\tau^m_k} - \sum_{1\le \alpha\le d} \psi_{\alpha}(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) B^{\alpha}_{\tau^m_{k-1},\tau^m_k}\Delta_m \Big|\\ \le C \Delta_m^{4H^-} \qquad \text{for all} \qquad 1\le k\le 2^m. \end{gather}\] Here, \(C\) depends on \(\sigma\) and \(b\) polynomially.

Proof. We show (1). From 20 , we have \[\begin{gather} \label{eq4382908401} Y^{\mathrm{CN},m}_t-Y^{\mathrm{CN},m}_{\tau^m_{k-1}} = \sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} +b(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})(t-\tau^m_{k-1})\\ + \frac{1}{2} \left( \sigma(Y^{\mathrm{CN},m}_t)-\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) \right) B_{\tau^m_{k-1},t} + \frac{1}{2} \left( b(Y^{\mathrm{CN},m}_t)-b(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) \right) (t-\tau^m_{k-1}). \end{gather}\tag{24}\] Hence applying the Taylor formula and writing \(Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}=Y^{\mathrm{CN},m}_t-Y^{\mathrm{CN},m}_{\tau^m_{k-1}}\), we have \[\begin{align} Y^{\mathrm{CN},m}_t -Y^{\mathrm{CN},m}_{\tau^m_{k-1}} -\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t} -b(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})(t-\tau^m_{k-1})\\&= \frac{1}{2} \left( \int_0^1 (D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}]d\theta \right) B_{\tau^m_{k-1},t} \\ &\phantom{=}\qquad \qquad \qquad + \frac{1}{2} \left( \int_0^1 (Db)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}]d\theta \right) (t-\tau^m_{k-1}) \\ &= ((D\sigma)[\sigma])(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) \left[ \frac{1}{2} B_{\tau^m_{k-1},t} \otimes B_{\tau^m_{k-1},t} \right] + \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}\\ &= ((D\sigma)[\sigma])(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) {\mathbb{B}}_{\tau^m_{k-1},t} + c(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})d^{\mathrm{CN},m}_{\tau^m_{k-1},t} + \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}.\end{align}\]

We show (2). From 20 , we have \[\begin{align} \max_k \sup_{\tau^m_{k-1}\leq t\leq\tau^m_k} |Y^{\mathrm{CN},m}_t-Y^{\mathrm{CN},m}_{\tau^m_{k-1}}| \le C |t-\tau^m_{k-1}|^{H^-}. \end{align}\] This estimate and 24 imply \[\begin{align} \max_k \sup_{\tau^m_{k-1}\leq t\leq\tau^m_k} |Y^{\mathrm{CN},m}_t-Y^{\mathrm{CN},m}_{\tau^m_{k-1}}-\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t}| \le C |t-\tau^m_{k-1}|^{2H^-}. \end{align}\] Hence, by substituting \[\begin{align} (D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[Y^{\mathrm{CN},m}_{\tau^m_{k-1},t}]\\&= (D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}+\theta Y^{\mathrm{CN},m}_{\tau^m_{k-1},t})[\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t}] + O(|t-\tau^m_{k-1}|^{2H^-})\\ &= (D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})[\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},t}] + O(|t-\tau^m_{k-1}|^{2H^-})\end{align}\] into 23 , we can estimate the first term in 23 . Because the second term can be estimated in the same way, we arrive at (2).

We show (3). By a similar calculation to the above, we have \[\begin{gather} \Big| \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},\tau^m_k} -\frac{1}{2}(D\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) \left[ \frac{1}{2}(D\sigma)[\sigma](Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) [(B_{\tau^m_{k-1},\tau^m_k})^{\otimes 2}] +b(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})\Delta_m\right] B_{\tau^m_{k-1},\tau^m_k}\\ -\frac{1}{4}(D^2\sigma)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}}) \left[ \sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k},\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} \right]B_{\tau^m_{k-1},\tau^m_k}\\ -\frac{1}{2}(Db)(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})\left[\sigma(Y^{\mathrm{CN},m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k}\right]\Delta_m \Big| \le C\Delta_m^{4H^-}. \end{gather}\] Note that the above constants depend on \(\sigma,b\) polynomially because \(\omega\in \Omega_0^{(m)}\). The proof completed. ◻

Remark 7. Let \(d^m=d^{\mathrm{IM},m}\), \(d^{\mathrm{M},m}\), \(d^{\mathrm{FE},m}\), \(d^{\mathrm{CN},m}\) and \(\hat{\epsilon}^{m}=\hat{\epsilon}^{\mathrm{IM},m}\), \(\hat{\epsilon}^{\mathrm{M},m}\), \(\hat{\epsilon}^{\mathrm{FE},m}\), \(\hat{\epsilon}^{\mathrm{CN},m}\). For every \(s,t\in D_m\) with \(s\le t\) , define \(d^m_t\), \(d^m_{s,t}\), \(\hat{\epsilon}^{m}_t\) and \(\hat{\epsilon}^{m}_{s,t}\) in the same way as 2 with \(\eta_{\tau^m_{i-1},\tau^m_i}=d^m_{\tau^m_{i-1},\tau^m_i},\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}\).

2.3 Statement of main results↩︎

Now we are in a position to state our main results (Theorem 12, Corollary 1 and Theorem 15). In Section 2.2, we recalled four approximation schemes and we wrote the solutions as \(\hat{Y}^{m}_t\). They are continuous processes but the values at the discrete times \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) well approximate \(\{\hat{Y}^{m}_t\}_{t\in [0,1]}\). Also it is natural to consider approximate schemes defined at discrete times \(D_m\) only for implementation. As stated in Introduction, in Theorem 12, we consider the recurrence relations of \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) can be obtained by adding extra two terms containing \(d^m\) and \(\hat{\epsilon}^{m}\) to the recurrence relation of the Milstein scheme. Since the Milstein scheme converges, we can expect that \(\hat{Y}^{m}_t\) also converges to \(Y_t\) if \(d^m\) and \(\hat{\epsilon}^{m}\) are small in a certain sense. Based on this idea, we introduce smallness conditions as Conditions 8\(\sim\)11 and address approximate solutions and estimates of the errors at discrete times \(D_m\). This is stated as Theorem 12, which is a result in a general setting not limited to the four schemes and fBms. Corollary 1 is a continuous version of Theorem 12. In Theorem 15, we give estimates of errors for the four schemes and fBms. Note that we can check Conditions 8\(\sim\)11 to use Corollary 1 for the four schemes except Crank-Nicolson scheme in the case of fBm with the Hurst parameter \(\frac{1}{3}<H\leq\frac{1}{2}\). Although the Crank-Nicolson scheme can also be reduced to a setting satisfying the conditions, it requires additional considerations.

Here we reset the notation to state Theorem 12. For \(\omega\in\Omega_0\), we define \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) by the following recurrence relation: \(\hat{Y}^{m}_0=\xi\) and \[\begin{align} \hat{Y}^{m}_{\tau^m_k} &= \hat{Y}^{m}_{\tau^m_{k-1}}+\sigma(\hat{Y}^{m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} +((D\sigma)[\sigma])(\hat{Y}^{m}_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},\tau^m_k} +b(\hat{Y}^{m}_{\tau^m_{k-1}})\Delta_m\nonumber\\ &\qquad\qquad\qquad\qquad +c(\hat{Y}^{m}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k} +\hat{\epsilon}^m_{\tau^m_{k-1},\tau^m_k}, \quad 1\le k\le 2^m. \label{equation32of32ymh} \end{align}\tag{25}\] Here \(c\in C^3_b({\mathbb{R}}^n, L({\mathbb{R}}^d\otimes {\mathbb{R}}^d,{\mathbb{R}}^n))\) is a function and \(d^m=\{d^m_{\tau^m_{k-1},\tau^m_k}\}_{1\leq k\leq 2^m}\subset {\mathbb{R}}^d\otimes {\mathbb{R}}^d\) and \(\hat{\epsilon}^{m}=\{\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}\}_{1\leq k\leq 2^m}\subset {\mathbb{R}}^n\) are random variables defined on \(\Omega_0\). We now state our smallness conditions on \(d^m\) and \(\hat{\epsilon}^{m}\).

Condition 8. There exist two pairs of positive numbers \((\varepsilon_0,2H^-)\) and \((\varepsilon_1,\lambda_1)\) with \(\varepsilon_i>0\) \((i=0,1)\) and \(\lambda_1+H^->1\) and non-negative random variables \(G_0=G_0(\varepsilon_0,2H^-)\) and \(G_1=G_1(\varepsilon_1,\lambda_1)\) which belong to \(\cap_{p\ge 1}L^p(\Omega_0)\) such that \[\begin{align} |d^m_{s,t}| \le \min \left\{ \Delta_m^{\varepsilon_0}G_0|t-s|^{2H^-}, \Delta_m^{\varepsilon_1} G_1 |t-s|^{\lambda_1} \right\} \quad \text{for all s,t\in D_m with s<t.} \end{align}\]

Although the reader might be interested in the reason why two exponents \(2H^-\) and \(\lambda_1\) are introduced, we defer the explanation to Remark 13 and proceed to state the conditions. We next explain a condition on \(\hat{\epsilon}^{m}\). In this condition, although (1-a) follows from (2), we state (1-a) independently because it is used in Section 4. Below, \(B^{\alpha,\beta,\gamma}_{s,t}\) \((0\le s\le t\le 1)\) denotes the \(e_{\alpha}\otimes e_{\beta}\otimes e_{\gamma}\)-component of the third level rough paths which are constructed from \((B,{\mathbb{B}})\).

Condition 9.

    1. There exists a positive constant \(C\) such that \[\begin{align} \label{eq4839012814214} |\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}| \le C \Delta_m^{3H^-} \qquad \text{for all} \qquad 1\le k\le 2^m,\quad \omega\in \Omega_0^{(m)}. \end{align}\tag{26}\] Here, \(C\) depends on \(\sigma\), \(b\) and \(c\) polynomially.

    2. There exists a positive constant \(C\) such that \[\begin{align} \label{eq4583902014} |\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}| \le C \Delta_m^{3H^-} \qquad \text{for all} \qquad 1\le k\le 2^m,\quad \omega\in \Omega_0\setminus \Omega_0^{(m)} \end{align}\tag{27}\] Here, \(C\) depends on \(\sigma\), \(b\), \(c\) and \(C(B)\) polynomially.

  1. There exist bounded Lipschitz continuous functions \(\varphi_{\alpha,\beta,\gamma} :{\mathbb{R}}^n\to{\mathbb{R}}^n\) and \(\psi_{\alpha} :{\mathbb{R}}^n\to{\mathbb{R}}^n\) such that \[\begin{gather} \Big| \hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k} - \sum_{1\le \alpha,\beta,\gamma\le d} \varphi_{\alpha,\beta,\gamma}(\hat{Y}_{\tau^m_{k-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{k-1},\tau^m_k} - \sum_{1\le \alpha\le d} \psi_{\alpha}(\hat{Y}^{m}_{\tau^m_{k-1}}) B^{\alpha}_{\tau^m_{k-1},\tau^m_k}\Delta_m \Big|\\ \le C \Delta_m^{4H^-} \qquad \text{for all} \qquad 1\le k\le 2^m, \quad \omega\in \Omega_0^{(m)}. \end{gather}\] Here, \(C\) depends on \(\sigma, b\) and \(c\) polynomially.

Here we state the main non-trivial condition assumed in our main results. For \(c\in C^3_b({\mathbb{R}}^n, L({\mathbb{R}}^d\otimes {\mathbb{R}}^d,{\mathbb{R}}^n))\), which is used in 25 , set \[\begin{align} \label{defIm} I^m_t = I^m_t(c,d^m) &= \sum_{i=1}^{\lfloor 2^mt \rfloor} J_{\tau^m_{i-1}}^{-1} c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}. \end{align}\tag{28}\] Let \(I^m|_{D_m}\) denote the discrete process defined as the restriction of \(I^m\) on \(D_m\).

Condition 10. Let \(I^m|_{D_m}\) be as above. For all \(p\geq 1\), we have \[\begin{align} \sup_m \| \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-} \|_{L^p}<\infty. \end{align}\]

We explain the final condition. Let \(d^{m,\alpha,\beta}_{\tau^m_{i-1},\tau^m_i} =(d^{m}_{\tau^m_{i-1},\tau^m_i},e_{\alpha}\otimes e_{\beta})\). We set \[\begin{align} \tilde{\mathcal{K}}^3_m &= \Big\{ \big\{ d^{m,\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}B^{\gamma}_{\tau^m_{i-1},\tau^m_i} \big\}_{i=1}^{2^m}, \, \big\{ B^{\alpha,\beta,\gamma}_{\tau^m_{i-1},\tau^m_i} \big\}_{i=1}^{2^m}, \, \big\{B^{0,\alpha}_{\tau^m_{i-1},\tau^m_i}\big\}, \, \big\{B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}\big\} ~\Big|~1\le \alpha,\beta,\gamma\le d \Big\}. \end{align}\] and \[\begin{align} \label{eqK} \mathcal{K}^3_m &= \left\{ \{K^m_t\}_{t\in D_m} \,\middle|\, K^m_t=\sum_{i=1}^{\lfloor 2^mt \rfloor}K^m_{\tau^m_{i-1},\tau^m_i} \,\, \text{for some}\,\, \big\{ K^m_{\tau^m_{i-1},\tau^m_i} \big\}_{i=1}^{2^m}\in \tilde{\mathcal{K}}^3_m \right\}. \end{align}\tag{29}\] Here we set \(K^m_0=0\) with convention. Note that \(B^{0,\alpha}_{\tau^m_{i-1},\tau^m_i}\), \(B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}\) are defined in (16 ).

Condition 11. There exist a pair of positive numbers \((\varepsilon_2,\lambda_2)\) with \(\lambda_2+H^->1\) and a non-negative random variable \(G_2=G_2(\varepsilon_2,\lambda_2)\in \cap_{p\ge 1}L^p(\Omega_0)\) such that for all discrete processes \(\{K^m_t\}_{t\in D_m}\in \mathcal{K}^3_m\), \[\begin{align} \big|(2^m)^{2H-\frac{1}{2}} K^{m}_{s,t}\big| &\le \Delta_m^{\varepsilon_2} G_2 |t-s|^{\lambda_2} \quad \text{for all s,t\in D_m.} \end{align}\]

In the above condition, we consider \(B^{\alpha,\beta,\gamma}_{\tau^m_{i-1},\tau^m_i}\) only in a subset of Wiener chaos of order 3 which can be obtained by iterated integrals of \(B\). However, noting the relation, \[\begin{align} \label{product32of32iterated32integral} \left\{ \begin{aligned} B^{\alpha,\beta}_{s,t}B^\gamma_{s,t} &= B^{\alpha,\beta,\gamma}_{s,t} +B^{\gamma,\alpha,\beta}_{s,t} +B^{\alpha,\gamma,\beta}_{s,t},\\ B^{\alpha}_{s,t}B^{\beta}_{s,t}B^{\gamma}_{s,t} &= \frac{1}{2} \left( B^{\alpha,\beta,\gamma}_{s,t}+ B^{\beta,\alpha,\gamma}_{s,t}+ B^{\beta,\gamma,\alpha}_{s,t}+ B^{\alpha,\gamma,\beta}_{s,t}+ B^{\gamma,\alpha,\beta}_{s,t}+ B^{\gamma,\beta,\alpha}_{s,t} \right), \end{aligned} \right. \end{align}\tag{30}\] which follows from the geometric property of \((B,{\mathbb{B}})\), we obtain similar estimates for sum processes defined by the above increments.

We now state our first main result. Note that we always assume Condition 3 on \(B\).

Theorem 12. Let \(Y_t\) be the solution to RDE 10 . Let \(c\in C^3_b({\mathbb{R}}^n, L({\mathbb{R}}^d\otimes {\mathbb{R}}^d,{\mathbb{R}}^n))\). Let \(d^m=\{d^m_{\tau^m_{k-1},\tau^m_k}\}_{k=1}^{2^m} \subset {\mathbb{R}}^d\otimes {\mathbb{R}}^d\) and \(\hat{\epsilon}^{m}=\{\hat{\epsilon}^{m}_{\tau^m_{k-1},t}\}_{k=1}^{2^m} \subset {\mathbb{R}}^n\) be random variables defined on \(\Omega_0\). Consider the approximate solution \(\hat{Y}^{m}_t\) \((t\in D_m)\) defined by 25 . Let \(I^m\) be the weighted sum process defined by 28 . Set \[\begin{align} \label{remainder32term32discrete} R^m_t &= \hat{Y}^{m}_t-Y_t-J_tI^m_t,\quad t\in D_m. \end{align}\tag{31}\] Let \(\frac{1}{2}(H+\frac{1}{4})<H^-<H\). Assume that Conditions  \(\sim\)  hold. Then for \(0<\varepsilon<\min \{3H^--1,4H^- - 2H - \frac{1}{2},\varepsilon_1,\varepsilon_2\}\), we have \(2^{m(2H-\frac{1}{2}+\varepsilon)} \max_{t\in D_m}|R^m_t|\to 0\) in \(L^p\) for all \(p\ge 1\) and almost surely.

The next is a remark on how to use Condition 8.

Remark 13. In our proof, we will use the Hölder estimate of \(d^m\) given by the pair \((\varepsilon_0,2H^-)\) to estimate an approximation of the Jacobian and its inverse (we write them as \(\tilde{J}^{m,\rho}\) and \((\tilde{J}^{m,\rho})^{-1}\) later) by using Cass-Litterer-Lyons’ estimate. On the other hand, the Hölder estimate given by the pair \((\varepsilon_1,\lambda_1)\) determines the convergence rate of the remainder term \(R^m_t\) in our main theorems. More precisely, \(\varepsilon_1\) is one of upper bounds of the convergence rate and we obtain a good convergence rate if we can choose large \(\varepsilon_1\).

A trivial choice of \((\varepsilon_1,\lambda_1)\) is \((\varepsilon_0,2H^-)\). In general, there is a trade-off between the Hölder exponent and the value of the Hölder norm. Hence for \(\lambda_1<2H^-\) we may be able to take \(\varepsilon_1>\varepsilon_0\). This is a good situation for our application. In fact we can implement this situation in our application. Therefore we may be able to take large \(\varepsilon_1\) for small \(\lambda_1\). We refer the readers for this to Remark 20.

In the above theorem, \(d^m_{s,t}\) and \(\hat{\epsilon}^{m}_{s,t}\) are defined only at the discrete times \((s,t)=(\tau^m_{k-1},\tau^m_k)\) \((1\le k\le 2^m)\). However, they are defined at \(\{\{(s, t)\}_{s=\tau^m_{k-1}, t\in [\tau^m_{k-1},\tau^m_k]}\}_{k=1}^{2^m}\) in some cases as in the four schemes we explained. As a corollary of this theorem, we have the following result in such a situation.

Corollary 1. We consider the same situation as in Theorem . Further we assume \(d^m_{s,t}\) and \(\hat{\epsilon}^{m}_{s,t}\) are defined at \(\{\{(s, t)\}_{s=\tau^m_{k-1}, t\in [\tau^m_{k-1},\tau^m_k]}\}_{k=1}^{2^m}\) and assume that there exists a positive random variable \(\hat{X}\in \cap_{p\ge 1}L^p(\Omega_0)\) such that \[\begin{align} \label{eq48902834} |d^m_{\tau^m_{k-1},t}|&\le \hat{X}|t-\tau^m_{k-1}|^{2H^-}, & |\hat{\epsilon}^{m}_{\tau^m_{k-1},t}|&\le \hat{X}|t-\tau^m_{k-1}|^{3H^-} \end{align}\tag{32}\] for all \(\tau^m_{k-1}<t\leq \tau^m_k\) and \(1\le k\le 2^m\). We define \(\hat{Y}^{m}_t\) \((0\leq t\leq 1)\) as an extension of \(\hat{Y}^{m}_t\) \((t\in D_m)\) via 25 , with \(\tau^m_k\) replaced by \(t(\in [\tau^m_{k-1},\tau^m_k])\). Set \[\begin{align} \label{remainder32term32conti} R^m_t=\hat{Y}^{m}_t-Y_t-J_tI^m_t, \quad 0\leq t\leq 1. \end{align}\tag{33}\] Then for the same constant \(\varepsilon\) as in Theorem \(\ref{main32theorem}\), we have \(2^{m(2H-\frac{1}{2}+\varepsilon)} \sup_{0\le t\le 1}|R^m_t|\to 0\) in \(L^p\) for all \(p\ge 1\) and almost surely.

We will prove the above results in Section 5. We make a remark on the estimate of \(\varepsilon\) in the above theorem.

Remark 14. We fix \(H^-\) and lift \(B\) to an \(H^-\)-Hölder rough path. It is necessary to give the meaning of the solutions \(Y_t\) and \(J_t\) of the differential equations. That is, they depends on the choice of \(H^-\). However, note that each \(\hat{Y}^{m}, Y_t, I^m_t\) are all almost surely defined for any choice of \(\frac{1}{3}<H^-<H\) in our problem because any versions of \((B,{\mathbb{B}})\) are identical almost all \(\omega\) for any \(H^-\) as noted in Remark 4. Therefore, the optimal constant of the estimate of \(\varepsilon\) in Theorem 12 should be independent of the choice of \(H^-\).

We now return to the four schemes stated in Section 2.2. We assume that \(B\) is an fBm. The following is the second main theorem.

Theorem 15. Let \(B\) be an fBm with the Hurst parameter \(\frac{1}{3}<H\leq \frac{1}{2}\). Let \(Y_t\) be the solution to RDE 10 . Consider the implementable Milstein, Crank-Nicolson, Milstein or first-order Euler scheme and let \(\hat{Y}^{m}_t\) and \(I^m_t\) be their counterparts. Let \(R^m_t\) \((0\leq t\leq 1)\) be defined by 33 . Then for \(0<\varepsilon<3H-1\), we have \(2^{m(2H-\frac{1}{2}+\varepsilon)}\sup_t|R^m_t|\to 0\) in \(L^p\) for all \(p\ge 1\) and almost surely.

We will show Theorem 15 for the four schemes in Section 2.4 with the help of Corollary 1. For the implementable Milstein, Milstein, and first-order Euler schemes, we can check the conditions assumed in Corollary 1. The Crank-Nicolson scheme satisfies Condition 9 only partially. Namely, while Lemma 2 implies that Condition 9 (1-a) and (2) holds, expression 22 yields that Condition 9 (1-b) does not hold. Hence we cannot use Corollary 1 directly. However, it is easy to reduce the problem of Crank-Nicolson scheme to the case which can be treated in Corollary 1.

We conclude this section with remarks on Theorem 15.

Remark 16. When we consider the Milstein scheme, we have \(d^m_{\tau^m_{k-1},\tau^m_k}=d^{\mathrm{M},m}_{\tau^m_{k-1},\tau^m_k}=0\). From Theorem 15, for any \(\kappa>0\), we have \((2^m)^{5H-\frac{3}{2}-\kappa}\sup_{t}|\hat{Y}^{m}_t-Y_t|\to 0\) as \(m\to\infty\) in \(L^p\) for all \(p\ge 1\) and almost surely. For the other schemes, we have \((2^m)^{2H-1/2-\kappa}\sup_t|\hat{Y}^{m}_t-Y_t|\to 0\) in the same sense. We will explain related weaker results in Theorem 28 and Remark 29.

Remark 17. We mention related study with the above results. Ueda [9] studied the estimate of the remainder term in one-dimensional case. By “one-dimensional”, we mean that the solution \(Y_t\) and the driving fBm \(B_t\) is one-dimensional. In this case, \(H\) can be arbitrary positive number less than \(1\). His study also is based on analysis of interpolation processes between the solutions and approximate solutions.

Remark 18. We make remarks on weak convergence of \((2^m)^{2H-\frac{1}{2}}I^m_t\) in the case of fBm. Let \(B\) be an fBm. Let \(d^m=d^{\mathrm{IM},m}=d^{\mathrm{CN},m}\). In this case, \(d^{m,\alpha,\beta}_{\tau^m_{k-1},\tau^m_k}=(d^{m}_{\tau^m_{k-1},\tau^m_k},e_{\alpha}\otimes e_{\beta})\) is given by \[\begin{align} d^{m,\alpha,\beta}_{\tau^m_{k-1},\tau^m_k}=\frac{1}{2}B^{\alpha}_{\tau^m_{k-1},\tau^m_k} B^{\beta}_{\tau^m_{k-1},\tau^m_k}-B^{\alpha,\beta}_{\tau^m_{k-1},\tau^m_k}. \end{align}\] Note that \(d^{m,\alpha,\beta}_{\tau^m_{k-1},\tau^m_k}=-d^{m,\beta,\alpha}_{\tau^m_{k-1},\tau^m_k}\) holds because the rough path is geometric. Furthermore, we see that \(\{(2^m)^{2H-\frac{1}{2}}J_tI^m_t\}_{0\le t\le 1}\) weakly converges to \[\begin{align} \left\{C\sum_{1\le \alpha, \beta\le d} J_t\int_0^tJ_s^{-1}(D\sigma)(Y_s)[\sigma(Y_s)e_{\alpha}]e_{\beta} dW^{\alpha,\beta}_s\right\}_{0\le t\le 1} \label{limit32process} \end{align}\tag{34}\] in \(D([0,1],{\mathbb{R}}^n)\) with respect to the Skorokhod \(J_1\)-topology. Here

  1. \(\{W^{\alpha,\beta}_t\}\) \((1\le \alpha<\beta\le d)\) is a \(\frac{1}{2}d(d-1)\)-dimensional standard Brownian motion which is independent of the fBm \((B_t)\) and \(W^{\beta,\alpha}_t=-W^{\alpha,\beta}_t\) \((\beta>\alpha)\), \(W^{\alpha,\alpha}_t=0\) \((1\le \alpha\le d)\).

  2. Let \(\alpha\ne\beta\). The constant \(C\) is given by \[\begin{align} C= & \Bigg\{ E[(B^{\alpha,\beta}_{0,1})^2] +2\sum_{k=1}^{\infty}E[B^{\alpha,\beta}_{0,1}B^{\alpha,\beta}_{k,k+1}] -\frac{1}{4}(E[(B^{\alpha}_{0,1})^2])^2 -\frac{1}{2} \sum_{k=1}^{\infty}E[B^{\alpha}_{0,1}B^{\alpha}_{k,k+1}]^2\Bigg\}^{\frac{1}{2}}. \end{align}\]

We proved this convergence in [12] under the assumption \(\sigma, b\in C^{\infty}_b\). Note that \(I^m_t\equiv 0\) in the case where \(d^m=d^{\mathrm{M},m}\). Also a similar convergence is proved in the case where \(d^m=d^{\mathrm{FE},m}\) by Liu-Tindel [7] too. See also [12].

Remark 19 (Weak convergence via Remark 18 and Theorem 15). Combining Remark 18 and Theorem 15, we can prove \(\{(2^m)^{2H-\frac{1}{2}}(\hat{Y}^{m}_t-Y_t)\}\) weakly converges to the weak limit of \(\{(2^m)^{2H-\frac{1}{2}}J_tI^m_t\}\) in \(D([0,1],{\mathbb{R}}^n)\) in the Skorokhod topology. This follows from the following more general result. Let \(\{Z^m_t\}_{0\le t\le 1}\), \(\{\tilde{Z}^m_t\}_{0\le t\le 1}\) and \(\{R^m_t\}_{0\le t\le 1}\) be \({\mathbb{R}}^n\)-valued càdlàg processes such that \(Z^m_t=\tilde{Z}^m_t+R^m_t\) holds almost surely. Suppose that \(\tilde{Z}^m\) converges weakly in \(D([0,1],{\mathbb{R}}^n)\) and \(\lim_{m\to\infty}E[\sup_t|R^m_t|]=0\). Then \(Z^m\) also converges weakly to the same limit of \(\tilde{Z}^m\). The reason is as follows. \(D([0,1],{\mathbb{R}}^n)\) is a Polish space with respect to a metric \(\rho\) on \(D([0,1],{\mathbb{R}}^n)\) which satisfies \(\rho(x,y)\le \sup_t|x_t-y_t|\). To prove the convergence and the coincidence of the limit, it suffices to show that \(\lim_{m\to\infty}E[\varphi(Z^m)-\varphi(\tilde{Z}^m)]=0\) for any bounded Lipschitz continuous function \(\varphi\) on \(D([0,1],{\mathbb{R}}^n)\). Clearly, this can be proved by using \[|\varphi(Z^m)-\varphi(\tilde{Z}^m)|\le \|\varphi\|_{{\rm Lip}} \rho(Z^m,\tilde{Z}^m)\le \|\varphi\|_{{\rm Lip}}\sup_t|R^m_t|,\] and the assumption on \(R^m\).

2.4 Proof of Theorem 15↩︎

In this section, we show Theorem 15. First, in the case of the four schemes, the implementable Milstein, Crank-Nicolson, Milstein and first-order Euler schemes, we show that Conditions , and hold, in this order, and then give a proof of Theorem 15.

Lemma 3. Assume that \(B\) is a \(d\)-dimensional fBm with \(\frac{1}{3}<H\leq\frac{1}{2}\). Let \(d^m\) be \(d^{\mathrm{IM},m}\), \(d^{\mathrm{CN},m}\), \(d^{\mathrm{M},m}\) or \(d^{\mathrm{FE},m}\). Then Condition  is satisfied for the pairs \((\varepsilon_1,\lambda_1)\) and \((\varepsilon_0,2H^-)\), where \(0<\varepsilon_1<3H^--1\), \(\lambda_1=1+2H-3H^-\) and \(0<\varepsilon_0<2(H-H^-)\).

Proof. Since \[\begin{align} d^{\mathrm{FE},m}_{\tau^m_{i-1},\tau^m_i} &= - \sum_{1\le \alpha\neq \beta\le d} B^{\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}e_\alpha\otimes e_\beta - \sum_{\alpha=1}^d \frac{1}{2} \left\{(B^{\alpha}_{\tau^m_{i-1},\tau^m_i})^2-\Delta_m^{2H}\right\} e_{\alpha}\otimes e_{\alpha}, \end{align}\] all components of \(d^m_{\tau^m_{i-1},\tau^m_i}\), \(d^{m,\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}=(d^m_{\tau^m_{i-1},\tau^m_i},e_\alpha\otimes e_\beta)\), are written by a linear combination of \[\begin{align} \label{eq43902498109} B^{\alpha}_{\tau^m_{i-1},\tau^m_i}B^{\beta}_{\tau^m_{i-1},\tau^m_i}, \quad B^{\alpha,\beta}_{\tau^m_{i-1},\tau^m_i},\quad (B^{\alpha}_{\tau^m_{i-1},\tau^m_i})^2-\Delta_m^{2H}, \quad \alpha\ne \beta. \end{align}\tag{35}\] Hence we may assume \(d^m_{\tau^m_{i-1},\tau^m_i}\) to be one of the above without loss of generality. These quantities are considered in several papers; for example [12], [7], [17], and [3]. In what follows, we assume \(\frac{1}{3}<H<\frac{1}{2}\). For the case \(H=\frac{1}{2}\), we can easily modify the discussion.

For \(k<l\), we have \[\begin{gather} \left| E\left[B^\alpha_{\tau^m_{k-1},\tau^m_k}B^\beta_{\tau^m_{k-1},\tau^m_k}B^\alpha_{\tau^m_{l-1},\tau^m_l}B^\beta_{\tau^m_{l-1},\tau^m_l}\right] \right| \le C \left( \frac{|k-l|^{2H-2}}{2^{2mH}} \right)^2,\\ \left| E\left[B^{\alpha,\beta}_{\tau^m_{k-1},\tau^m_k}B^{\alpha,\beta}_{\tau^m_{l-1},\tau^m_l}\right] \right| \le C \left( \frac{|k-l|^{2H-2}}{2^{2mH}} \right)^2,\\ \left| E \left[ (B^{\alpha}_{\tau^m_{k-1},\tau^m_k})^2-\Delta_m^{2H}) (B^{\alpha}_{\tau^m_{l-1},\tau^m_l})^2-\Delta_m^{2H}) \right] \right| \leq C \left( \frac{|k-l|^{2H-2}}{2^{2mH}} \right)^2. \end{gather}\] For \(k=l\), the terms above can be estimate by \(C(2^{-2mH})^2\). We refer the readers for these estimates to Lemma 3.4 in [7]. Also we can find these estimates in Lemma 7.2 (1) in [12]. These estimates imply \[\begin{align} E[|d^m_{s,t}|^{2}]&\le C \left(\frac{1}{2^m}\right)^{4H-1}(t-s)\qquad \text{for s,t\in D_m with s<t.} \end{align}\] Note that all constants \(C\) above are independent of \(m\) and \(H\). By using the hypercontractivity of the Ornstein-Uhlenbeck semigroup, we get \[\begin{align} E[|d^m_{s,t}|^{p}]&\le C_p \left(\frac{1}{2^m}\right)^{(2H-\frac{1}{2})p}(t-s)^{\frac{p}{2}} \qquad \text{for s,t\in D_m with s<t.} \label{moment32estimate32for32dm} \end{align}\tag{36}\] This estimate implies the next assertion. For \(0<\kappa<\frac{1}{2}\), set \[\begin{align} G_{m,\kappa}=(2^m)^{2H-\frac{1}{2}}\max_{s,t\in D_m, s\ne t}\frac{|d^m_{s,t}|}{|t-s|^{\frac{1}{2}-\kappa}}. \end{align}\] Then \[\begin{align} &\qquad\qquad \sup_m\|G_{m,\kappa}\|_{L^p}<\infty & &\text{for all p\ge 1}, \tag{37}\\ & |d^m_{s,t}|\le \Delta_m^{2H-\frac{1}{2}} |t-s|^{\frac{1}{2}-\kappa}G_{m,\kappa} & & \text{for all s,t\in D_m with s<t} \tag{38}. \end{align}\] This can be checked as follows. Since we see 38 from the definition of \(G_{m,\kappa}\), we show integrability 37 . Let \(\{\tilde{d}^m_t\}_{t\in [0,1]}\) be the piecewise linear extension of \(\{d^m_t\}_{t\in D_m}\). By (36 ), we have \[\begin{align} E[|\tilde{d}^m_{s,t}|^p]&\le 3^{p-1}C_p \left(\frac{1}{2^m}\right)^{(2H-\frac{1}{2})p}|t-s|^{\frac{p}{2}}. \end{align}\] By the Garsia-Rodemich-Rumsey inequality, we have for any \(p,\theta>0\) \[\begin{align} \left(\sup_{s,t, s\ne t}\frac{|\tilde{d}^m_{s,t}|}{|t-s|^{\theta}}\right)^p \le 2\int_0^1\int_0^t\frac{|\tilde{d}^m_{s,t}|^p}{|t-s|^{2+p\theta}} dsdt. \end{align}\] Combining these two inequalities and setting \(\theta=\frac{1}{2}-\kappa\), we get \[\begin{align} E[G_{m,\kappa}^p]&\le 2\cdot(2^m)^{(2H-\frac{1}{2})p} \int_0^1\int_0^t\frac{E[|\tilde{d}^m_{s,t}|^p]}{|t-s|^{2+p\theta}}dsdt \le 2\cdot 3^{p-1} C_p\int_0^1\int_0^t|t-s|^{\kappa p-2}dsdt. \end{align}\] If \(p>\kappa^{-1}\), then the right-hand side is bounded and we get \[\begin{align} E[G_{m,\kappa}^p]\le 2\cdot 3^{p-1}C_p \left(\kappa p(\kappa p-1)\right)^{-1}, \end{align}\] which proves 37 .

By using 37 and 38 , we show the assertion. Let us choose \(0<\varepsilon<2H-\frac{1}{2}\) and \(0<2\kappa<\varepsilon\). Using \(\Delta_m\le t-s\), we get \[\begin{align} \text{(RHS of \eqref{estimate32of32dm322})} &=\Delta_m^{\varepsilon-\kappa} \Delta_m^{2H-\frac{1}{2}-\varepsilon+\kappa}|t-s|^{\frac{1}{2}-\kappa}G_{m,\kappa}\\ &\le \Delta_m^{\varepsilon-\kappa}|t-s|^{2H-\varepsilon}G_{m,\kappa}\\ &=\Delta_m^{\varepsilon-2\kappa}|t-s|^{2H-\varepsilon}\Delta_m^{\kappa}G_{m,\kappa} \end{align}\] Let \(G_1=\sum_{m=1}^{\infty}\Delta_m^{\kappa}G_{m,\kappa}\). This infinite series converges for \(\mu\) almost all \(\omega\). Because for all \(p\ge 1\), \[\begin{align} \|G_1\|_{L^p} \le \sum_{m=1}^{\infty} \Delta_m^{\kappa} \sup_m\|G_{m,\kappa}\|_{L^p} < \infty. \end{align}\] Combining the trivial estimate \(\Delta_m^{\kappa}G_{m,\kappa}\le G_1\), we get \[\begin{align} |d^m_{s,t}| &\le \Delta_m^{\varepsilon-2\kappa}|t-s|^{2H-\varepsilon}G_1. \end{align}\] To check the validity of the statements for the pairs \((\varepsilon_1,\lambda_1)\) and \((\varepsilon_0,2H^-)\), it suffices to set \(\varepsilon=3H^--1(<2H-\frac{1}{2})\) and \(\varepsilon=2(H-H^-)(<2H-\frac{1}{2})\) respectively and choose \(\kappa\) to be sufficiently small. This completes the proof. ◻

Remark 20. We make a remark on the numbers appeared in Lemma 3. Recall that \(\lambda_1=1+2H-3H^-\) and that \(3H^--1\) and \(2(H-H^-)\) are the upper bounds of \(\varepsilon_1\) and \(\varepsilon_0\), respectively. We see that both inequalities \(\lambda_1<2H^-\) and \(3H^--1>2(H-H^-)\) are equivalent to \(5H^--2H>1\). The inequality \(5H^--2H>1\) holds true if \(H^-\) is sufficiently close to \(H\) because \(H>\frac{1}{3}\). Hence we see that the good situation stated in Remark 13 is fulfilled.

Lemma 4. Assume that \(B\) is a \(d\)-dimensional fBm with \(\frac{1}{3}<H\leq\frac{1}{2}\). Let \(d^m\) be \(d^{\mathrm{IM},m}\), \(d^{\mathrm{CN},m}\), \(d^{\mathrm{M},m}\) or \(d^{\mathrm{FE},m}\). Then Condition  is satisfied for \(\varepsilon_2<3H^--1+(\frac{1}{2}-H)\) and \(\lambda_2=1+2H-3H^-\).

Proof. In what follows, we assume \(\frac{1}{3}<H<\frac{1}{2}\). In the case where \(H=\frac{1}{2}\), we can easily modify the discussion. Let \((K^m_t)\in \mathcal{K}^3_m\). First, we give estimates for variance of \(K^m_{s,t}\). We have for \(s,t\in D_m\) with \(s<t\),

E[|K^m_s,t|^2] C _m^6H-1|t-s| &
C _m^2H+1|t-s| &

Note that if the schemes are implementable Milstein or Crank-Nicolson scheme, then it is enough to consider the case \(K^m=B^{\alpha,\beta,\gamma}\) only for the proof of ([K32estimate321]) because of the identities (30 ). Therefore, in those cases, from [2], we see ([K32estimate321]) holds. In [12], the same estimates are obtained in a little bit different way. If the scheme is the first-order Euler scheme, then by the same reasoning as above, it is sufficient to estimate \(E[(\Delta_m^{2H} B^{\gamma}_{s,t})^2]\). For this, we have \[\begin{align} E[(\Delta_m^{2H} B^{\gamma}_{s,t})^2]&\le C\Delta_m^{4H}|t-s|^{2H}\\ &=C\Delta_m^{4H}\cdot|t-s|^{2H-1}|t-s|\\ &\le C\Delta_m^{4H}\Delta_m^{2H-1}|t-s|=C\Delta_m^{6H-1}|t-s|. \end{align}\] Actually we use Condition 3 only to obtain this estimate.

Now we consider ([K32estimate322]). Let \(K^m_{\tau^m_{i-1},\tau^m_i}=B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}=\int_{\tau^m_{i-1}}^{\tau^m_i}B^{\alpha}_{\tau^m_{i-1},u}du\). By using \(|E[B^{\alpha}_{\tau^m_{i-1},u}B^{\alpha}_{\tau^m_{j-1},v}]| \leq |E[B^{\alpha}_{\tau^m_{i-1},\tau^m_i}B^{\alpha}_{\tau^m_{j-1},\tau^m_j}]|\) for \(\tau^m_{i-1}\leq u\leq \tau^m_i\leq \tau^m_{j-1}\leq v\leq \tau^m_j\), we have \[\begin{align} | E [ K^m_{\tau^m_{i-1},\tau^m_i} K^m_{\tau^m_{j-1},\tau^m_j} ] | &\leq \int_{\tau^m_{i-1}}^{\tau^m_i} du \int_{\tau^m_{j-1}}^{\tau^m_j} dv |E[B^{\alpha}_{\tau^m_{i-1},\tau^m_i}B^{\alpha}_{\tau^m_{j-1},\tau^m_j}]|\\ &\leq 2^{-2m} 2^{-2Hm} |E[B^{\alpha}_{0,1}B^{\alpha}_{j-i-1,j-i}]|. \end{align}\] Noting \(E[B^{\alpha}_{0,1}B^{\alpha}_{k-1,k}]\sim -H(1-2H)k^{2H-2}\) as \(k\to\infty\), we have for \(k2^{-m}=s<t=l2^{-m}\), \[\begin{align} E[(K^m_{s,t})^2]\le C\sum_{i,j=k+1}^l 2^{-2m}2^{-2Hm}|j-i|^{2H-2}\le C(2^{-m})^{2H+1}|t-s|. \end{align}\] As for \(B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}\), we have \(B^{0,\alpha}_{\tau^m_{i-1},\tau^m_i}=B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}-\Delta_mB^{\alpha}_{\tau^m_{i-1},\tau^m_i}\). Hence, we need to estimate \(E[(\Delta_m B^{\alpha}_{s,t})^2]\). Since \(\Delta_m\le \Delta_m^{2H}\), this term is smaller than \(E[(\Delta_m B^{\alpha}_{s,t})^2]\) and we get desired estimate.

Because \(6H-1\le 2H+1\), consequently, for all cases, we have \(E[|K^m_{s,t}|^2]\le C\Delta_m^{6H-1}|t-s|\). Combining the hypercontractivity of the Ornstein-Uhlenbeck semigroup and the estimates above, for all \(p\ge 2\), we obtain \[\begin{align} E[|K^{m}_{s,t}|^{p}] \le C_p\left(2^{-m}\right)^{(3H-\frac{1}{2})p}(t-s)^{\frac{p}{2}} \quad \text{for all s,t\in D_m}. \end{align}\] From the same argument as in (38 ), for any \(\frac{1}{2}>\kappa>0\) and \(m\), there exists a positive random variable \(G'_{m,\kappa}\) satisfying \(\sup_m\|G'_{m,\kappa}\|_{L^p}<\infty\) for all \(p\ge 1\) such that \[\begin{align} |K^{m}_{s,t}|&\le \Delta_m^{3H-\frac{1}{2}}|t-s|^{\frac{1}{2}-\kappa}G'_{m,\kappa}\quad \text{for all s,t\in D_m}, \end{align}\] which implies \[\begin{align} |(2^{m})^{2H-\frac{1}{2}}K^{m}_{s,t}|&\le \Delta_m^{\frac{1}{2}-H} \Delta_m^{2H-\frac{1}{2}} |t-s|^{\frac{1}{2}-\kappa}G'_{m,\kappa}\quad \text{for all s,t\in D_m}.\label{estimate32of32294-mK94m} \end{align}\tag{39}\] Note that \(\Delta_m^{2H-\frac{1}{2}}|t-s|^{\frac{1}{2}-\kappa}\) appears in the proof of Lemma 3 (see 38 ).

Let us choose \(0<\varepsilon<2H-\frac{1}{2}\) and \(0<2\kappa<\varepsilon\). Then again using \(\Delta_m\le t-s\) and similarly to the estimate of \(d^m_{s,t}\), we get \[\begin{align} |(2^{m})^{2H-\frac{1}{2}}K^{m}_{s,t}| &\le \Delta_m^{\frac{1}{2}-H} \Delta_m^{\varepsilon-2\kappa} |t-s|^{2H-\varepsilon} \Delta_m^{\kappa} G'_{m,\kappa} \end{align}\] and set \(G_2=\sum_{m=1}^{\infty}\Delta_m^{\kappa}G'_{m,\kappa}\) which converges \(\mu\)-a.s. \(\omega\) and \(\|G_2\|_{L^p}<\infty\) for all \(p\ge 1\). Again by using the trivial estimate \(\Delta_m^{\kappa}G'_{m,\kappa}\le G_2\), we get \[\begin{align} |(2^{m})^{2H-\frac{1}{2}}K^{m}_{s,t}| &\le \Delta_m^{\frac{1}{2}-H} \Delta_m^{\varepsilon-2\kappa} |t-s|^{2H-\varepsilon} G_2. \end{align}\] Putting \(\varepsilon=3H^--1(<2H-\frac{1}{2})\), we completes the proof. ◻

Lemma 5. Assume that \(B\) is a \(d\)-dimensional fBm with \(\frac{1}{3}<H\leq\frac{1}{2}\). Let \(d^m\) be \(d^{\mathrm{IM},m}\), \(d^{\mathrm{CN},m}\), \(d^{\mathrm{M},m}\) or \(d^{\mathrm{FE},m}\). Then Condition  holds.

Proof. Recall \(c=(D\sigma)[\sigma]\in C^3_b\). We show the case \(\frac{1}{3}<H<\frac{1}{2}\). We use the result by Liu-Tindel [7]. They considered similar problems (Proposition 4.7 and Corollary 4.9 in [7]). We can use their result to show the assertion as follows. Note that \(f_t=J^{-1}_tc(Y_t)\in \mathcal{L}({\mathbb{R}}^d\otimes{\mathbb{R}}^d,{\mathbb{R}}^n)\) and \(g_t\in \mathcal{L}({\mathbb{R}}^d,\mathcal{L}({\mathbb{R}}^d\otimes{\mathbb{R}}^d,{\mathbb{R}}^n))\) defined by \(g_tv=(-J^{-1}_tD\sigma(Y_t)v)c(Y_t)+J^{-1}_tDc(Y_t)[\sigma(Y_t)v]\) for \(v\in{\mathbb{R}}^d\) satisfy [7] because \(Y\) and \(J^{-1}\) are solutions to 10 and 12 respectively and they belong to \(L^p\) for all \(p\ge 1\). The integrability of \(J_t^{-1}\) is due to [13] (see also Remark 29). Hence from Corollary 4.9 in [7], we get \(\|(2^m)^{2H-\frac{1}{2}}I^m_{s,t}\|_{L^p} \leq C(t-s)^{\frac{1}{2}}\) for some constant \(C\). This and the Garsia-Rodemich-Rumsey inequality imply the assertion. While the above proof is based on the result by Liu-Tindel [7], we can provide another proof of the assertion under the assumption that \(\sigma, b\in C^{\infty}_b\) (see [12]).

Finally, we consider the case where \(H=\frac{1}{2}\). Actually, it is not difficult to check this case by using the Itô calculus. For the reader’s convenience, we include the proof. Recall that \(I^m_t\) in Condition  is defined by \(I^m_t=\sum_{i=1}^{2^mt}F_{\tau^m_{i-1}}d^m_{\tau^m_{i-1},\tau^m_i}\) \((t\in D_m)\), where \(F_{t}=J_t^{-1}c(Y_t)\). We give an estimate of \(E[|I^m_{s,t}|^{2p}]\) by applying martingale theory. Since all components of \(d^{m}_{\tau^m_{i-1},\tau^m_i}\), \(d^{m,\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}=(d^m_{\tau^m_{i-1},\tau^m_i},e_\alpha\otimes e_\beta)\), are written by a linear combination of 35 , the desired estimates follow from those of \[\begin{align} \label{eq439028014214} & \sum_{i=1}^{2^mt} F^{\alpha,\beta}_{\tau^m_{i-1}} B^{\alpha}_{\tau^m_{i-1},\tau^m_i}B^{\beta}_{\tau^m_{i-1},\tau^m_i}, & & \sum_{i=1}^{2^mt} F^{\alpha,\beta}_{\tau^m_{i-1}} B^{\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}, & & \sum_{i=1}^{2^mt} F^{\alpha,\alpha}_{\tau^m_{i-1}} \{ (B^{\alpha}_{\tau^m_{i-1},\tau^m_i})^2-\Delta_m \}, \end{align}\tag{40}\] where \(F^{\alpha,\beta}_t=F_t(e_\alpha\otimes e_\beta)\) and \(\alpha\neq\beta\). For \(t\in[0,1]\), let \[\tilde{I}^m_t = \sum_{i=1}^{2^{m}} F^{\alpha,\beta}_{\tau^m_{i-1}} B^{\alpha,\beta}_{\tau^m_{i-1}\wedge t,\tau^m_i\wedge t}.\] Clearly \(I^m_t=\tilde{I}^m_t\) \((t\in D_m)\) holds. Note that \[B^{\alpha}_{s,t}B^{\beta}_{s,t}=B^{\alpha,\beta}_{s,t}+B^{\beta,\alpha}_{s,t} \quad (\alpha\ne\beta), \quad (B^{\alpha}_{s,t})^2-(t-s)=\int_s^tB^{\alpha}_{s,u}dB^{\alpha}_u,\] where the integral in the second identity is the Itô integral. Therefore, for all cases in (40 ), it suffices to give the moment estimate of \[\begin{align} \tilde{I}^{m,\alpha,\beta}_t&= \int_0^tF^{m,\alpha,\beta}_udB^{\beta}_u,\qquad 1\le \alpha,\beta\le d, \end{align}\] where the integral is an Itô integral and \(F^{m,\alpha,\beta}_u=\sum_{i=1}^{2^m}F^{\alpha,\beta}_{\tau^m_{i-1}} B^{\alpha}_{\tau^m_{i-1},u}1_{[\tau^m_{i-1},\tau^m_i)}(u).\) Let \(p>1\). We have \[\begin{align} \nonumber E\left[|\tilde{I}^{m,\alpha,\beta}_{s,t}|^{2p}\right] &\le CE\left[\left(\int_s^t|F^{m,\alpha,\beta}_u|^{2}du\right)^p\right]\\ &\le C(t-s)^{p-1}E\left[\int_s^t|F^{m,\alpha,\beta}_u|^{2p}du\right] \le C'\left(\frac{t-s}{2^m}\right)^p,\label{moment32estimate32H611472} \end{align}\tag{41}\] where we have used the Burkholder-Davis-Gundy and the Hölder inequalities, and the estimate \[E[|F^{m,\alpha,\beta}_u|^{2p}]\le CE[|F^{\alpha,\beta}_{\tau^m_{i-1}}|^{2p}]E[(B^{\alpha}_{\tau^m_{i-1},u})^{2p}] \le C2^{-pm}\sup_tE[|F_t|^{2p}], \quad \tau^m_{i-1}\le u<\tau^m_i.\] By the estimate (41 ) and a similar argument to the estimate (37 ) of \(d^m_{s,t}\), we see that the assertion holds.

We conclude this proof with mentioning that, under the assumption of this lemma, Condition  holds for all \(H^-<\frac{1}{2}\) and that we can choose \(H^-\) close to \(\frac{1}{2}\). ◻

We now prove Theorem 15.

Proof of Theorem 15. First, we prove the case of the implementable Milstein, Milstein and first-order Euler schemes. Note that in these cases, \(\hat{\epsilon}^{m}\equiv 0\) holds for the approximate solution \(\hat{Y}^{m}_t\). Hence Condition  is clearly satisfied. From Lemmas 3, 4, and 5, we see that Conditions , and hold. From the definition, 32 also holds. Hence the conditions assumed in Corollary 1 are satisfied. By Corollary 1, for any \(\varepsilon<\min\{3H^--1,4H^--2H-\frac{1}{2}\}\), we have \((2^m)^{2H-\frac{1}{2}+\varepsilon}\sup_{0\leq t\leq 1}|R^m_t|\to 0\) in \(L^p\) \((p\ge 1)\) and almost surely. Since \(H^-\) can be any positive number less than \(H\) and \(3H-1\leq 2H-\frac{1}{2}\), the proof is completed.

We consider the case of the Crank-Nicolson approximate solution \(Y^{\mathrm{CN},m}_t\). We cannot directly apply Corollary 1 to the Crank-Nicolson scheme since it satisfies only Condition 9 (1-a) and (2). However we can reduce it to Corollary 1. To this end, we introduce an auxiliary approximate solution \(\hat{Y}^{m}_t\) defined via 25 , with \(\tau^m_k\) replaced by \(t(\in [\tau^m_{k-1},\tau^m_k])\) and \[\begin{align} d^m_{\tau^m_{k-1},t} &= d^{\mathrm{CN},m}_{\tau^m_{k-1},t}, & \hat{\epsilon}^{m}_{\tau^m_{k-1},t} &= \begin{cases} \hat{\epsilon}^{\mathrm{CN},m}_{\tau^m_{k-1},t}, & \omega\in\Omega_0^{(m)},\\ 0, & \omega\in (\Omega_0^{(m)})^{\complement}. \end{cases} \end{align}\] Lemmas 2 and the definition above imply that \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}\) satisfies Condition . From Lemmas 3, 4, and 5, we see that Conditions , and hold. We see that \(d^m_{\tau^m_{k-1},t}\) and \(\hat{\epsilon}^{m}_{\tau^m_{k-1},t}\) satisfy 32 . Hence we can apply Corollary 1 to \(\hat{Y}^{m}_t\) defined above. By using \(\sup_{0\le t\le 1}|J_t|\in L^p\) \((p\ge 1)\) which is due to [13], as a consequence of Corollary 1, we see that \(\sup_m\sup_{0\le t\le 1}|\hat{Y}^{m}_t|\in L^p\) \((p\ge 1)\). Note that we will give selfcontained proof of the integrability of \(J_t\) and \(J_t^{-1}\) in Remark 29 (2) and the integrability of \(\hat{Y}^{m}_t\) holds under weaker assumption as in Lemma 9 since (32 ) holds. Let \(R^m_t=\hat{Y}^{m}_t-Y_t-J_tI^m_t\) and \(R^{\mathrm{CN},m}_t=Y^{\mathrm{CN},m}_t-Y_t-J_tI^m_t\). Then using \(\hat{Y}^{m}_t=Y^{\mathrm{CN},m}_t\) \((\omega\in \Omega_0^{(m)})\) and \(Y^{\mathrm{CN},m}_t\equiv \xi\) \((\omega\in (\Omega_0^{(m)})^{\complement})\), we have \[\begin{align} R^{\mathrm{CN},m}_t &= R^m_t+Y^{\mathrm{CN},m}_t-\hat{Y}^{m}_t\\ &= R^m_t + \{Y^{\mathrm{CN},m}_t-\hat{Y}^{m}_t\} 1_{\Omega_0^{(m)}} + \{Y^{\mathrm{CN},m}_t-\hat{Y}^{m}_t\} 1_{(\Omega_0^{(m)})^\complement}\\ &= R^m_t+\{\xi-\hat{Y}^{m}_t\} 1_{(\Omega_0^{(m)})^\complement}. \end{align}\] By Corollary 1, we have \((2^m)^{2H-\frac{1}{2}+\varepsilon}\sup_{0\le t\le 1}|R^m_t|\to 0\) for all \(p\ge 1\) and almost surely. By the integrability of \(\sup_m\sup_{0\le t\le 1}|\hat{Y}^{m}_t|\) and the estimate (9 ), we have \((2^m)^{2H-\frac{1}{2}+\varepsilon}\sup_{0\le t\le 1}|(\xi-\hat{Y}^{m}_t) 1_{(\Omega_0^{(m)})^\complement}|\to 0\) in \(L^p\) and almost surely. This completes the proof. ◻

2.5 Small order nice discrete process↩︎

We introduce a class of discrete stochastic processes, which includes \(d^m_t\) satisfying Condition 8. Before doing so, we need to define a subset of \(\Omega_0^{(m)}\). For a positive number \(\lambda_1\) satisfying \(\lambda_1+H^->1\), we introduce the following set: \[\begin{align} \Omega_0^{(m,d^m)} &= \{\, \omega\in \Omega_0^{(m)} ~|~ \|d^m(\omega)\|_{2H^-}\le 1,\quad \|d^m(\omega)\|_{\lambda_1}\le 1 \}. \end{align}\] Similarly to the estimate of the complement of \(\Omega^{(m)}_0\), if Condition 8 holds with the same exponent \(\lambda_1\) in the definition of \(\Omega_0^{(m,d^m)}\), we can prove that for any \(p\ge 1\), there exists \(C_p>0\) such that \[\begin{align} \label{complement32of32Omegam} \mu\left((\Omega_0^{(m,d^m)})^\complement\right)\le C_{p}2^{-mp} \end{align}\tag{42}\] which implies the complement of \(\Omega_0^{(m,d^m)}\) is also negligible set for our problem.

Definition 3.

  1. Let \(\eta=\{(\eta^m_t)_{t\in D_m}; m\ge m_0\}\) be a sequence of Banach space valued random variables such that \(\eta^m_0=0\) and \(\{\eta^m_t\}_{t\in D_m}\) is defined on \(\Omega_0^{(m,d^m)}\) for each \(m\), where \(m\ge m_0\) and \(m_0\) is a non-random constant and depends on the sequence. Let \(\{a_m\}\) be a positive sequence which converges to \(0\). Let \(\lambda\) be a positive number such that \(\lambda+H^->1\). We say that \(\eta=(\eta^m)\) is a \(\{a_m\}\)-order nice discrete process with the Hölder exponent \(\lambda\) if there exists a positive random variable \(X\in \cap_{p\ge 1}L^p(\Omega_0)\) which is independent of \(m\) such that \[\begin{align} \|\eta^m_t-\eta^m_s\|\le a_m X(\omega)|t-s|^{\lambda} \qquad \text{for all m\ge m_0,\, t,s\in D_m,\, \omega\in \Omega_0^{(m,d^m)}}. \label{small32order32discrete32process32inequality} \end{align}\tag{43}\]

  2. Let \(\{v^m_{\theta}\}_{\theta\in \Theta}\) be a family of Banach space valued random variables defined on \(\Omega_0^{(m,d^m)}\), where \(m\ge m_0\). Let \(\{a_m\}\) be a positive sequence which converges to \(0\). If there exists a non-negative random variable \(X\in \cap_{p\ge 1}L^p(\Omega_0)\) which does not depend on \(m\) such that \[\begin{align} \sup_{\theta\in \Theta}\|v^m_{\theta}\|\le a_m X(\omega) \quad \text{for all m and \omega\in \Omega_0^{(m,d^m)}}, \end{align}\] then we write \[\begin{align} \sup_{\theta\in \Theta}\|v^m_{\theta}\|=O(a_m). \end{align}\]

Remark 21. Here we give examples of small order nice discrete processes.

  1. Let \(\epsilon^m_{\tau^m_{k-1},\tau^m_k}\) be given by 13 . Assume that Conditions 8, 9 (1) and 11 are satisfied. Let \(\varepsilon_1,\lambda_1, \varepsilon_2, \lambda_2\) be the numbers appeared in Condition 8 and 11. Set \(a_m=\max\{\Delta_m^{3H^{-}-1},\Delta_m^{\varepsilon_1},\Delta_m^{\varepsilon_2}\}\) and \(\lambda=\min\{2H^-,\lambda_1,\lambda_2\}\). Let \(\omega\in\Omega_0\). Then there exists a non-negative random variable \(X\in \cap_{p\ge 1}L^p(\Omega_0)\) which is independent of \(m\) such that \[\begin{align} |d^m_{s,t}| +|\epsilon^m_{s,t}| +|\hat{\epsilon}^{m}_{s,t}| +|(2^m)^{2H-\frac{1}{2}}K^m_{s,t}| \le a_m X |t-s|^\lambda \quad \text{for all} \quad s, t\in D_m. \label{condition32on32dm32ep32eph} \end{align}\tag{44}\] In particular, \(d^m_t\), \(\epsilon^m_t\), \(\hat{\epsilon}^m_t\) and \((2^m)^{2H-\frac{1}{2}}K^m_t\) are \(\{a_m\}\)-order nice discrete processes with the Hölder exponent \(\lambda\). We need to check \(\epsilon^m\) and \(\hat{\epsilon}^{m}\) satisfy the inequality. For \(s=\tau^m_l\) and \(t=\tau^m_k\), Lemma 1 and Condition 9 (1) imply \[\begin{align} |\epsilon^m_{s,t}| + |\hat{\epsilon}^{m}_{s,t}| = \sum_{i=l+1}^k \left\{ |\epsilon^m_{\tau^m_{i-1},\tau^m_i}| + |\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}| \right\} \le C(k-l)\Delta_m^{3H^-} \le C\Delta_m^{3H^--1}|t-s|^{2H^-}, \end{align}\] where the constant \(C\) depends \(\sigma\), \(b\), \(c\) and \(C(B)\) polynomially. If we consider the pair \((\varepsilon_0, 2H^-)\), we can prove that there exist \(\tilde{X}\in \cap_{p\ge 1}L^p(\Omega_0)\) and \(\tilde{a}_m=\max\{\Delta_m^{\varepsilon_0},\Delta_m^{3H^--1}\}\) such that \[\begin{align} |d^m_{s,t}|+|\epsilon^m_{s,t}|+|\hat{\epsilon}^{m}_{s,t}|\le \tilde{a}_m \tilde{X}|t-s|^{2H^-}. \end{align}\] We use the estimate (44 ) in Sections 4.2 and 4.4.

  2. In the above definition of \(\{a_m\}\)-order nice discrete processes, we assume the strong assumption on \(X\) such that \(X\in \cap_{p\ge 1}L^p(\Omega_0)\). Under Conditions 3 and 8, we have many examples which satisfy this strong conditions.

Remark 22. Suppose a Banach space valued discrete process \(F=\{(F^m_t)_{t\in D_m};\) \(m\ge m_0\}\) defined on \(\Omega_0^{(m,d^m)}\) satisfy the Hölder continuity \[\begin{align} & \|F^m_t-F^m_s\|\le X_F(\omega)|t-s|^{H^{-}}\quad \,\, \text{for all m\ge m_0,\, s,t\in D_m,\, \omega\in \Omega_0^{(m,d^m)}}, \\ & \sup_m\|F^m_0(\omega)\|\le Y_F(\omega)\quad\quad \text{for}~~ \omega\in \Omega_0^{(m,d^m)}. \end{align}\] Here \(X_F, Y_F\in \cap_{p\ge 1}L^p(\Omega_0)\) are random variables independent of \(m\). If \(\eta=(\eta^m)\) is a real valued \(\{a_m\}\)-order nice discrete process with the Hölder exponent \(\lambda\), then \[\begin{align} \tilde{\eta}^m_{\tau^m_k}=\sum_{i=1}^kF^m_{\tau^m_{i-1}}\eta^m_{\tau^m_{i-1},\tau^m_i} \end{align}\] is also a \(\{a_m\}\)-order nice discrete process with the Hölder exponent \(\lambda\) by the estimate of the (discrete) Young integral (see [11]): \[\begin{align} \|\tilde{\eta}^m\|_{\lambda}&\le C\left(\|F_0^m\|+\|F^m\|_{H^-}\right) \|\eta^m\|_{\lambda}, \end{align}\] where \(C\) is a constant depending only on \(H^-\) and \(\lambda\). Note that we used \(\lambda+H^->1\).

This property is very nice for our purpose. However, in our application, since the estimate on \(F^m\) is satisfied only on \(\Omega_0^{(m,d^m)}\), we cannot require 43 for all \(\omega\in\Omega_0\) to be nice discrete processes.

Remark 23. In what follows, we use the following elementary summation by parts formula several times: For sequences \(\{f_i\}_{i=0}^n\), \(\{g_i\}_{i=0}^n\), we have \[\begin{align} \sum_{i=1}^n f_{i-1}g_{i-1,i} &= f_ng_n-f_0g_0 - \sum_{i=1}^nf_{i-1,i}g_i. \label{summation32by32parts32formula} \end{align}\tag{45}\] We will use this formula when we give estimates of discrete Young integral.

3 An interpolation of discrete rough differential equations↩︎

Let \(Y_t\) and \(\hat{Y}^{m}_t\) be a solution to 10 and an approximate solution given by 25 , respectively. In previous section, we observe that the discrete stochastic processes \(\{Y_t\}_{t\in D_m}\) and \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) corresponding to the solution and our approximate solutions respectively of the RDE satisfy the following common recurrence form: \(Y_0=\hat{Y}^{m}_0=\xi\) and, for \(1\leq k\leq 2^m\), \[\begin{align} Y_{\tau^m_k} &= Y_{\tau^m_{k-1}} +\sigma(Y_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} +((D\sigma)[\sigma])(Y_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},\tau^m_k} +b(Y_{\tau^m_{k-1}})\Delta_m +\epsilon^m_{\tau^m_{k-1},\tau^m_k},\\ \hat{Y}^{m}_{\tau^m_k} &= \hat{Y}^{m}_{\tau^m_{k-1}} +\sigma(\hat{Y}^{m}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} +((D\sigma)[\sigma])(\hat{Y}^{m}_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},\tau^m_k} +b(\hat{Y}^{m}_{\tau^m_{k-1}})\Delta_m\\ &\qquad\qquad\qquad +c(\hat{Y}^{m}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k} +\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}. \end{align}\] We now introduce an interpolation process between \(\{Y_t\}_{t\in D_m}\) and \(\{\hat{Y}^{m}_t\}_{t\in D_m}\) to study the difference \(\hat{Y}^{m}_t-Y_t\). Moreover, we introduce a matrix valued process \(\tilde{J}^{m,\rho}_t\) which approximates the derivative process \(J_t\) when \(m\to \infty\). Note that, in this section, we do not use any specific forms of \(d^m\) and \(\hat{\epsilon}^{m}\) which were given in Section 2. Taking a look at the recurrence equations, we see that the different points between \(\hat{Y}^{m}_t\) and \(Y_t\) are the terms \(c(\hat{Y}^{m}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k}\), \(\hat{\epsilon}^m_{\tau^m_{k-1},\tau^m_k}\) and \(\epsilon^m_{\tau^m_{k-1},\tau^m_k}\). In view of this, we define a sequence \(\{Y^{m,\rho}_t\}_{t\in D_m}\) by the following recurrence relation: \(Y^{m,\rho}_0=\xi\) and, for \(1\leq k\leq 2^m\), \[\begin{align} Y^{m,\rho}_{\tau^m_k} &= Y^{m,\rho}_{\tau^m_{k-1}} +\sigma(Y^{m,\rho}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} +((D\sigma)[\sigma])(Y^{m,\rho}_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},\tau^m_k} +b(Y^{m,\rho}_{\tau^m_{k-1}})\Delta_m \nonumber \\ &\qquad\qquad\qquad +\rho c(Y^{m,\rho}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k} +\rho\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k} +(1-\rho)\epsilon^m_{\tau^m_{k-1},\tau^m_k}. \label{def95interplation} \end{align}\tag{46}\] Note that \(Y^{m,0}_t=Y_t\) and \(Y^{m,1}_t=\hat{Y}^{m}_t\) (\(t\in D_m\)). In this paper, we call this recurrence relation a discrete RDE. The function \([0,1]\ni\rho\mapsto Y^{m,\rho}_t\) is smooth and \[\begin{align} \hat{Y}^{m}_t-Y_t=\int_0^1\partial_{\rho}Y^{m,\rho}_t d\rho \end{align}\] holds. We give the estimate for \(\hat{Y}^{m}_t-Y_t\) by using the estimate of \(Z^{m,\rho}_t=\partial_{\rho}Y^{m,\rho}_t\). Then \(\{Z^{m,\rho}_t\}_{t\in D_m}\) satisfies \(Z^{m,\rho}_0=0\) and, for \(1\leq k\leq 2^m\), \[\begin{align} Z^{m,\rho}_{\tau^m_k} &= Z^{m,\rho}_{\tau^m_{k-1}} +(D\sigma)(Y^{m,\rho}_{\tau^m_{k-1}})[Z^{m,\rho}_{\tau^m_{k-1}}]B_{\tau^m_{k-1},\tau^m_k} +\left(D((D\sigma)[\sigma])\right)(Y^{m,\rho}_{\tau^m_{k-1}})[Z^{m,\rho}_{\tau^m_{k-1}}]{\mathbb{B}}_{\tau^m_{k-1},\tau^m_k} \nonumber \\ &\phantom{=}\qquad +(Db)(Y^{m,\rho}_{\tau^m_{k-1}})[Z^{m,\rho}_{\tau^m_{k-1}}]\Delta_m +\rho (Dc)(Y^{m,\rho}_{\tau^m_{k-1}})[Z^{m,\rho}_{\tau^m_{k-1}}]d^m_{\tau^m_{k-1},\tau^m_k}\nonumber \\ &\phantom{=}\qquad +c(Y^{m,\rho}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k} +\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k} -\epsilon^m_{\tau^m_{k-1},\tau^m_k}, \label{zmr32identity} \end{align}\tag{47}\] where \[\begin{align} \label{DDsigma2} (D((D\sigma)[\sigma]))(y)[\eta] v\otimes w = D^2\sigma (y)[\eta,\sigma(y)v]w+D\sigma(y)[D\sigma(y)[\eta]v]w \end{align}\tag{48}\] for \(y,\eta\in{\mathbb{R}}^n\) and \(v,w\in{\mathbb{R}}^d\) (see also 5 ).

We introduce the \(\mathcal{L}({\mathbb{R}}^n)\)-valued, that is, matrix valued process \(\{\tilde{J}^{m,\rho}_t\}_{t\in D_m}\) to obtain the estimates of \(\{Z^{m,\rho}_t\}_{t\in D_m}\). Let \(\{\tilde{J}^{m,\rho}_t\}_{t\in D_m}\) be the solution to the following recurrence relation: \(\tilde{J}^{m,\rho}_0=I\) and, for \(1\leq k\leq 2^m\), \[\begin{align} \tilde{J}^{m,\rho}_{\tau^m_k} &= \tilde{J}^{m,\rho}_{\tau^m_{k-1}} +[D\sigma](Y^{m,\rho}_{\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}]B_{\tau^m_{k-1},\tau^m_k} +\left(D((D\sigma)[\sigma])\right)(Y^{m,\rho}_{\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}]{\mathbb{B}}_{\tau^m_{k-1},\tau^m_k}\nonumber\\ & \qquad \qquad +(Db)(Y^{m,\rho}_{\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}]\Delta_m+\rho (Dc)(Y^{m,\rho}_{\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}]d^m_{\tau^m_{k-1},\tau^m_k}. \label{Mmr} \end{align}\tag{49}\] Clearly, we can represent \(\{Z^{m,\rho}_t\}_{t\in D_m}\) by using \(\{\tilde{J}^{m,\rho}_t\}_{t\in D_m}\) and \(\{(\tilde{J}^{m,\rho}_t)^{-1}\}_{t\in D_m}\) if \(\tilde{J}^{m,\rho}_t\) are invertible by a constant variation method. Actually, such kind of representation holds in general case too. To show this, and for later purpose, we consider discrete RDEs which are driven by time shift process of \(B_t\).

Let \(u\in D_m\) with \(u\leq 1-\Delta_m\). For \(\tau^m_k\leq 1-u\), we introduce time shift variables: \[\begin{align} (\theta_uB)_{\tau^m_{k-1},\tau^m_k} &= B_{u+\tau^m_{k-1},u+\tau^m_k}, & (\theta_u{\mathbb{B}})_{\tau^m_{k-1},\tau^m_k} &={\mathbb{B}}_{u+\tau^m_{k-1},u+\tau^m_k},\\ (\theta_ud^m)_{\tau^m_{k-1},\tau^m_k} &= d^m_{u+\tau^m_{k-1},u+\tau^m_k},\\ (\theta_u\epsilon^m)_{\tau^m_{k-1},\tau^m_k} &=\epsilon^m_{u+\tau^m_{k-1},u+\tau^m_k}, & (\theta_u\hat{\epsilon}^{m})_{\tau^m_{k-1},\tau^m_k} &=\hat{\epsilon}^{m}_{u+\tau^m_{k-1},u+\tau^m_k}. \end{align}\] For general \(x\in {\mathbb{R}}^n\), we define a discrete process \(\{Y^{m,\rho}_t(x)\}_{t\in D_m,0\leq t\leq 1-u}\) by \(Y^{m,\rho}_0(x)=x\) and, for \(\tau^m_k\leq 1-u\), \[\begin{align} Y^{m,\rho}_{\tau^m_k}(x) &= Y^{m,\rho}_{\tau^m_{k-1}}(x) +\sigma(Y^{m,\rho}_{\tau^m_{k-1}}(x))(\theta_uB)_{\tau^m_{k-1},\tau^m_k}\\ &\qquad\qquad +((D\sigma)[\sigma])(Y^{m,\rho}_{\tau^m_{k-1}}(x)) (\theta_u{\mathbb{B}})_{\tau^m_{k-1},\tau^m_k} +b(Y^{m,\rho}_{\tau^m_{k-1}}(x))\Delta_m\\ &\qquad\qquad +\rho c(Y^{m,\rho}_{\tau^m_{k-1}}(x))(\theta_ud^m)_{\tau^m_{k-1},\tau^m_k} +\rho(\theta_u\hat{\epsilon}^{m})_{\tau^m_{k-1},\tau^m_k} +(1-\rho)(\theta_u\epsilon^m)_{\tau^m_{k-1},\tau^m_k}. \end{align}\] To make clear the dependence of the driving process, we may denote the solution of the above equation by \(Y^{m,\rho}_t(x,\theta_uB)\). For simplicity, we write \(Y^{m,\rho}_t\) for \(Y^{m,\rho}_t(\xi,B)\). Using these notation, we have \(Y^{m,\rho}_t(Y^{m,\rho}_u(\xi,B),\theta_uB)=Y^{m,\rho}_{u+t}(\xi,B)\). We consider the case where \(x=Y^{m,\rho}_u\) (\(u\in D_m\) with \(u\leq 1-\Delta_m\)) below.

We now explain explicit representation of \(\tilde{J}^{m,\rho}_t\). For given \(x\in {\mathbb{R}}^n\), let \[\begin{align} \label{defEmr} E^{m,\rho}(x,\theta_tB) &= I +(D\sigma)(x)B_{t,t+\Delta_m} +D((D\sigma)[\sigma])(x){\mathbb{B}}_{t,t+\Delta_m} \nonumber \\ &\phantom{=}\qquad +(Db)(x)\Delta_m +\rho (Dc)(x)d^m_{t,t+\Delta_m}. \end{align}\tag{50}\] Then for \(t\in D_m\) with \(t>0\), we have \[\begin{align} \tilde{J}^{m,\rho}_t&= E^{m,\rho}(Y^{m,\rho}_{t-\Delta_m},\theta_{t-\Delta_m}B) E^{m,\rho}(Y^{m,\rho}_{t-2\Delta_m},\theta_{t-2\Delta_m}B) \cdots E^{m,\rho}(\xi,B). \end{align}\] Since \(\tilde{J}^{m,\rho}_t\) depends on \(\xi\) and \(B\), we may denote \(\tilde{J}^{m,\rho}_t\) by \(\tilde{J}^{m,\rho}_t(\xi,B)\). Next we define \(\tilde{J}^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB)\) similarly to \(Y^{m,\rho}_t(x,\theta_uB)\). That is, \(\tilde{J}^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB)\) is defined by substituting \(Y^{m,\rho}_u(=Y^{m,\rho}_u(\xi,B))\), \(\theta_uB\), \(\theta_u{\mathbb{B}}\), \(\theta_ud^m\) for \(\xi\), \(B\), \({\mathbb{B}}\), \(d^m\) in the equation 49 of \(\tilde{J}^{m,\rho}_t(=\tilde{J}^{m,\rho}_t(\xi,B))\). Using \(Y^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB)=Y^{m,\rho}_{u+t}(\xi,B)\), we see that \(\tilde{J}^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB)\) satisfies \(\tilde{J}^{m,\rho}_0(Y^{m,\rho}_u, \theta_uB)=I\) and, for \(\tau^m_k\leq 1-u\), \[\begin{align} \tilde{J}^{m,\rho}_{\tau^m_k}(Y^{m,\rho}_u,\theta_uB) &= \tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB) +[D\sigma](Y^{m,\rho}_{u+\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB)]B_{u+\tau^m_{k-1},u+\tau^m_k} \\ &\qquad\qquad +\left(D((D\sigma)[\sigma])\right)(Y^{m,\rho}_{u+\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB)]{\mathbb{B}}_{u+\tau^m_{k-1},u+\tau^m_k}\\ &\qquad\qquad +(Db)(Y^{m,\rho}_{u+\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB)]\Delta_m\\ &\qquad\qquad +\rho (Dc)(Y^{m,\rho}_{u+\tau^m_{k-1}})[\tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB)]d^m_{u+\tau^m_{k-1},u+\tau^m_k}. \end{align}\] From this equation, we obtain \[\begin{align} \tilde{J}^{m,\rho}_{\tau^m_k}(Y^{m,\rho}_u,\theta_uB) &= E^{m,\rho}(Y^{m,\rho}_{u+\tau^m_{k-1}},\theta_{u+\tau^m_{k-1}}B) \tilde{J}^{m,\rho}_{\tau^m_{k-1}}(Y^{m,\rho}_u,\theta_uB),\label{EMmrk} \end{align}\tag{51}\] which implies \[\begin{gather} \tilde{J}^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB)\\ = E^{m,\rho}(Y^{m,\rho}_{u+t-\Delta_m},\theta_{u+t-\Delta_m}B) E^{m,\rho}(Y^{m,\rho}_{u+t-2\Delta_m},\theta_{u+t-2\Delta_m}B) \cdots E^{m,\rho}(Y^{m,\rho}_u,\theta_uB)\label{time32shift32of32Mmrk} \end{gather}\tag{52}\] Also we have, for \(s,t\in D_m\) with \(s+t\leq 1-u\), \[\begin{align} \tilde{J}^{m,\rho}_{s+t}(Y^{m,\rho}_u,\theta_uB) &= \tilde{J}^{m,\rho}_s(Y^{m,\rho}_{u+t},\theta_{u+t}B) \tilde{J}^{m,\rho}_t(Y^{m,\rho}_u,\theta_uB).\label{Mmrk3943k} \end{align}\tag{53}\] The proof of 53 is as follows. By 52 , we have \[\begin{align} \tilde{J}^{m,\rho}_{s+t}(Y^{m,\rho}_u,\theta_uB) &= E^{m,\rho}(Y^{m,\rho}_{u+t+s-\Delta_m},\theta_{u+t+s-\Delta_m}B) \cdots E^{m,\rho}(Y^{m,\rho}_{u+t},\theta_{u+t}B)\\ &\qquad \cdot E^{m,\rho}(Y^{m,\rho}_{u+t-\Delta_m},\theta_{u+t-\Delta_m}B) \cdots E^{m,\rho}(Y^{m,\rho}_{u},\theta_{u}B)\\ &= \tilde{J}^{m,\rho}_s(Y^{m,\rho}_{u+t},\theta_{u+t}B) \tilde{J}^{m,\rho}_t(Y^{m,\rho}_{u},\theta_{u}B). \end{align}\]

We have the following lemma for the invertibility of \(\tilde{J}^{m,\rho}_t\).

Lemma 6. For \(1\leq k\leq 2^m\), we have \[\begin{align} \tilde{J}^{m,\rho}_{\tau^m_k} &= E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)\tilde{J}^{m,\rho}_{\tau^m_{k-1}}\nonumber\\ &= \Bigl( I +(D\sigma)(Y^{m,\rho}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k} +D((D\sigma)[\sigma])(Y^{m,\rho}_{\tau^m_{k-1}}){\mathbb{B}}_{\tau^m_{k-1},\tau^m_k}\nonumber\\ &\qquad \qquad \qquad \qquad \qquad +\rho (Dc)(Y^{m,\rho}_{\tau^m_{k-1}})d^m_{\tau^m_{k-1},\tau^m_k} +(Db)(Y^{m,\rho}_{\tau^m_{k-1}})\Delta_m \Bigr) \tilde{J}^{m,\rho}_{\tau^m_{k-1}}, \label{expansion32of32Mmr2} \end{align}\tag{54}\] and for large \(m\), \(\tilde{J}^{m,\rho}_t\) are invertible. For example, for any \(\omega\in \Omega_0^{(m,d^m)}\), if \(m\) satisfies \[\begin{align} \Delta_m^{H^-}\|D\sigma\|+\Delta_m^{2H^-} \|D\left((D\sigma)[\sigma]\right)\|+ \Delta_m^{2H^-}\|Dc\|+\Delta_m\|Db\|\le \frac{1}{2}, \label{assumption32on32m} \end{align}\tag{55}\] then \(E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)\) is invertible and it holds that \[\begin{align} \left|E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)^{-1}-I+(D\sigma)(Y^{m,\rho}_{\tau^m_{k-1}})B_{\tau^m_{k-1},\tau^m_k}\right|&\le C\Delta_m^{2H^-}, \qquad 1\le k\le 2^m,\label{estimate32of32Emr94-1} \end{align}\tag{56}\] where \(C\) depends on \(\sigma,b,c\) polynomially.

Proof. Under the assumption, \(E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)^{-1}\) is given by the Neumann series of \(A^{m,\rho}_{\tau^m_{k-1}} = I - E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)\). The estimate of the residual terms implies 56 . ◻

Remark 24. When we consider the inverse \((\tilde{J}^{m,\rho}_t)^{-1}\), we always assume that \(\omega\in \Omega_0^{(m,d^m)}\) and \(m\) satisfies 55 .

We have the following representation of \(Z^{m,\rho}_t\).

Lemma 7. For any \(t\in D_m\) with \(t>0\), we have \[\begin{align} Z^{m,\rho}_t &= \sum_{i=1}^{2^mt} \tilde{J}^{m,\rho}_{t-\tau^m_i}(Y^{m,\rho}_{\tau^m_i},\theta_{\tau^m_i}B) \left( c(Y^{m,\rho}_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i} +\hat{\epsilon}^m_{\tau^m_{i-1},\tau^m_i} -\epsilon^m_{\tau^m_{i-1},\tau^m_i} \right). \label{rep32for32zmrk} \end{align}\tag{57}\] If all \(Z^{m,\rho}_s(\xi,B)\) \((s\in D_m,0\leq s\leq t)\) are invertible, \[\begin{align} Z^{m,\rho}_t &= \tilde{J}^{m,\rho}_t \sum_{i=1}^{2^mt} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} \left( c(Y^{m,\rho}_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i} +\hat{\epsilon}^m_{\tau^m_{i-1},\tau^m_i} -\epsilon^m_{\tau^m_{i-1},\tau^m_i} \right). \end{align}\]

Proof. The second statement follows from 53 and 57 . We show 57 . Write \(k=2^mt\) and denote by \(\zeta_k\) the quantity on the right-hand side of 57 . For simplicity we write \[\begin{align} c_{i-1}d_{i-1,i}= c(Y^{m,\rho}_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i},\quad \epsilon_{i-1,i}= \hat{\epsilon}^m_{\tau^m_{i-1},\tau^m_i}- \epsilon^m_{\tau^m_{i-1},\tau^m_i}. \end{align}\] From 51 , we have \[\begin{align} \zeta_k-(\zeta_{k-1}+c_{k-1}d_{k-1,k}+\epsilon_{k-1,k})\\&= \sum_{i=1}^{k-1} \left\{ \tilde{J}^{m,\rho}_{\tau^m_{k-i}}(Y^{m,\rho}_{\tau^m_i},\theta_{\tau^m_i}B) - \tilde{J}^{m,\rho}_{\tau^m_{k-i-1}}(Y^{m,\rho}_{\tau^m_i},\theta_{\tau^m_i}B) \right\} (c_{i-1}d_{i-1,i}+\epsilon_{i-1,i}) \\ &= \sum_{i=1}^{k-1} \{ E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)-I \} \tilde{J}^{m,\rho}_{\tau^m_{k-i-1}}(Y^{m,\rho}_{\tau^m_i},\theta_{\tau^m_i}B) (c_{i-1}d_{i-1,i}+\epsilon_{i-1,i})\\ &= \{ E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)-I \} \sum_{i=1}^{k-1} \tilde{J}^{m,\rho}_{\tau^m_{k-i-1}}(Y^{m,\rho}_{\tau^m_i},\theta_{\tau^m_i}B) (c_{i-1}d_{i-1,i}+\epsilon_{i-1,i})\\ &= \{ E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)-I \} \zeta_{k-1},\end{align}\] which implies \[\begin{align} \zeta_k = E^{m,\rho}(Y^{m,\rho}_{\tau^m_{k-1}},\theta_{\tau^m_{k-1}}B)\zeta_{k-1} +c_{k-1}d_{k-1,k}+\epsilon_{k-1,k}. \end{align}\] Comparing the above with 47 , we complete the proof. ◻

Remark 25.

  1. We do not use the notation \(J^{m,\rho}_t\) to denote the solution of 49 . The reason is as follows. It is natural to use \((Y^{m,\rho}_t,J^{m,\rho}_t)\) to denote the interpolation process between \((Y_t, J_t)\) and its approximate solution, that is, we expect that \((Y^{m,0}_t,J^{m,0}_t)\) and \((Y^{m,1}_t,J^{m,1}_t)\) coincide \((Y_t, J_t)\) and its approximate solution, respectively. However, \(\tilde{J}^{m,\rho}_t\) is not such an process. In fact, \(\tilde{J}^{m,0}_t\) is not equal to \(J_t\). Differently from this, in the case of the implementable Milstein and Milstein schemes, \((\hat{Y}^{m}_t,\tilde{J}^{m,1}_t)\) is identical to the corresponding approximate solution of \((Y_t, J_t)\).

  2. When we consider quantity associated with \(\{Y^{m,\rho}_t\}\), \(\{a_m\}\)-order nice discrete process \(\eta\) may depend on a parameter \(\rho\) \((0\le \rho\le 1)\). For \(\eta^{\rho}=\{(\eta^{m,\rho}_t)_{t\in D_m}; m=1,2,\ldots \}\), if we can choose the random variable \(X\) in (43 ) independently of \(\rho\), we say that \(\eta^{\rho}\) is a \(\{a_m\}\)-order nice discrete process independent of \(\rho\).

For later use, we introduce the following.

Definition 4. When \(\tilde{J}^{m,\rho}_t\) is invertible, we define \(\tilde{Z}^{m,\rho}_t=(\tilde{J}^{m,\rho}_t)^{-1}Z^{m,\rho}_t\) for \(t\in D_m\). Explicitly, \[\begin{align} \label{def32tzmr} \tilde{Z}^{m,\rho}_t = \sum_{i=1}^{2^mt} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} \left( c(Y^{m,\rho}_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i} +\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i} -\epsilon^m_{\tau^m_{i-1},\tau^m_i} \right). \end{align}\tag{58}\]

Proposition 26. We assume (55 ) holds. For any \(\omega\in \Omega_0^{(m,d^m)}\), we obtain the following neat expression \[\begin{align} \hat{Y}^{m}_t-Y_t&=\int_0^1 \tilde{J}^{m,\rho}_t \tilde{Z}^{m,\rho}_t d\rho. \end{align}\]

Below, we prove that under appropriate assumptions: as \(m\to\infty\),

  1. \(\tilde{J}^{m,\rho}_t\to J_t\), \((\tilde{J}^{m,\rho}_t)^{-1}\to J_t^{-1}\), \(Y^{m,\rho}_t\to Y_t\) uniformly in \(t\in D_m\) for all \(\omega\in \Omega_0^{(m,d^m)}\).

  2. \((2^m)^{2H-\frac{1}{2}} \sum_{i=1}^{2^mt} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} \big( \hat{\epsilon}^m_{\tau^m_{i-1},\tau^m_i}-\epsilon^m_{\tau^m_{i-1},\tau^m_i} \big)\) converges to \(0\) uniformly in \(t\in D_m\).

Hence it is reasonable to conjecture the main theorem holds true by Proposition 26. We prove our main theorem by using estimates for \(\tilde{Z}^{m,\rho}\).

Remark 27 (List of notations).

  • \(Y_t\): solution of RDE

  • \(\hat{Y}^{m}_t\): discrete approximate solution of \(Y_t\)

  • \(Y^{m,\rho}_t\): an interpolated process between \(Y_t(=Y^{m,0}_t)\) and \(\hat{Y}^{m}_t(=Y^{m,1}_t)\)

  • \(J_t=\partial_{\xi}Y_t(\xi,B)\)

  • \(\tilde{J}^{m,\rho}_t\): \(\mathcal{L}({\mathbb{R}}^n)\)-valued process defined by \(Y^{m,\rho}_t\) which approximates \(J_t\)

  • \(\tilde{J}^m_t=\tilde{J}^{m,0}_t\)

  • \(Z^{m,\rho}_t=\partial_{\rho}Y^{m,\rho}_t\)

  • \(\tilde{Z}^{m,\rho}_t=(\tilde{J}^{m,\rho}_t)^{-1}Z^{m,\rho}_t\) (see Definition 4)

  • \(E^{m,\rho}(Y^{m,\rho}_s,\theta_sB)=\tilde{J}^{m,\rho}_t(\tilde{J}^{m,\rho}_s)^{-1}\) for \(t-s=\Delta_m\) (see 50 and Lemma 6)

4 Estimates of \(Y^{m,\rho}_t\) and \(\tilde{J}^{m,\rho}_t\)↩︎

In this section, we give estimates for \(Y^{m,\rho}_t\), \(\tilde{J}^{m,\rho}_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}\) which do not depend on \(\rho\). Recall that \(\{Y^{m,\rho}_t\}_{t\in D_m}\) satisfies \(Y^{m,\rho}_0=\xi\) and 46 . This equation is defined by the data of random variables \(d^m=\{d^m_{\tau^m_{k-1},\tau^m_k}\}_{k=1}^{2^m}\), \(\hat{\epsilon}^{m}=\{\hat{\epsilon}^{m}_{\tau^m_{k-1},\tau^m_k}\}_{k=1}^{2^m}\) \((m=1,2,\ldots)\) and \(c\in C^3_b({\mathbb{R}}^n,L({\mathbb{R}}^d\otimes{\mathbb{R}}^d,{\mathbb{R}}^n))\). \(d^m\) and \(\hat{\epsilon}^{m}\) need not to be corresponding quantities defined in Section 2.2 and it is not necessary that \(c=(D\sigma)[\sigma]\). Note that we define \(d^m_{s,t},\hat{\epsilon}^{m}_{s,t}\) for general \(s,t\in D_m\) with \(s<t\) by 2 with \(\eta_{\tau^m_{i-1},\tau^m_i}=d^m_{\tau^m_{i-1},\tau^m_i},\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}\). We choose \(0<\lambda_1<1\) so that \(\lambda_1+H^->1\) arbitrarily and fix it. Note that \(\|d^m\|_{\lambda_1}<\infty\) because \(d^m_t\) is defined on the finite set \(D_m\).

In Section 4.1, for \(\omega\in \Omega_0\), by applying Davie’s method [10], we give an estimate for \(Y^{m,\rho}_t\) in terms of the three constants \(C\) given in 14 , 26 , and 27 , and \(\|d^m\|_{\lambda_1}\).

In Section 4.2, we give estimates for \(\max_{t\in D_m}\big\{|\tilde{J}^{m,\rho}_t|,|(\tilde{J}^{m,\rho}_t)^{-1}|\big\}1_{\Omega_0^{(m,d^m)}}\). The coefficient of the discrete RDE for which \(\tilde{J}^{m,\rho}\) satisfies is not bounded but linear growth. Hence, we cannot apply the estimate in Section 4.1. To overcome the difficulty, we view the \(H^{-1}\)-Hölder rough path \((B_{s,t},{\mathbb{B}}_{s,t})\) as a rough path of finite \((H^{-})^{-1}\)-variation norm. Note that we assume Condition 3 on \(B_t\) and so we can apply the result due to Cass-Litterer-Lyons [13] (see Lemma 19 below) to obtain the estimate of \(\tilde{J}^{m,\rho}\) and \((\tilde{J}^{m,\rho})^{-1}\) similarly to \(J_t\) and \((J_t)^{-1}\). In Section 4.3, we give estimates for \(J_t-\tilde{J}^m_t\) and \(J^{-1}_t-(\tilde{J}^m_t)^{-1}\) on \(\Omega_0^{(m)}\) by using the results in Section 4.2. In Section 4.4 , we give estimates for \(\tilde{J}^{m,\rho}_t-J_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}-J^{-1}_t\).

4.1 Estimates of \(Y^{m,\rho}_t\) on \(\Omega_0\)↩︎

For \(s,t\in D_m\) with \(s\le t\), let \[\begin{align} \label{Ist32def} I_{s,t} = Y^{m,\rho}_t -Y^{m,\rho}_s -\sigma(Y^{m,\rho}_s)B_{s,t} -((D\sigma)[\sigma])(Y^{m,\rho}_s){\mathbb{B}}_{s,t} -\rho c(Y^{m,\rho}_s)d^m_{s,t} -b(Y^{m,\rho}_s)(t-s). \end{align}\tag{59}\] First, we prove the following.

Lemma 8. Assume that Condition  (1) holds and let \(\omega\in \Omega_0\). Let \(\lambda_1\) be a positive number satisfying \(\lambda_1+H^->1\). Set \(\lambda=\min\{\lambda_1,2H^-\}\). There exist \(0<\delta\le 1\) and \(C_1>0\) such that \[\begin{align} |I_{s,t}|\le C_1|t-s|^{\lambda+H^-},\quad \quad \text{s,t\in D_m \,\, with \,\, |t-s|\le \delta}. \label{estimate32of32Ist} \end{align}\tag{60}\] Here \(\delta^{-1}\) and \(C_1\) depend only on \(\sigma, b, c\), \(C(B)\) and \(\|d^m\|_{\lambda_1}\) polynomially.

Proof. Below, \(C\) is a constant depending only on \(\sigma\), \(b\), \(c\), \(C(B)\) and \(\|d^m\|_{\lambda_1}\) polynomially. By using \(C\), we determine \(\delta\) and \(C_1\) so that 60 holds. For simplicity we write \(\tau^m_i=t_i\). Let \(s=t_k, t=t_{k+l}\). By \(I_{t_k,t_{k+1}}=(1-\rho)\epsilon^m_{t_k,t_{k+1}} +\rho\hat{\epsilon}^m_{t_k,t_{k+1}}\) and the estimate of \(\hat{\epsilon}^m\), we see that 60 holds for any \(\delta\) and for the maximum of three constants \(C\) stated in 14 , 26 , and 27 . Let \(K\ge 1\). Suppose the following estimate: there exists \(M>0\) such that \[\begin{align} |I_{s,t}|&\le M|t-s|^{\lambda+H^-} \end{align}\] holds for \(\{(s,t)=(t_k, t_{k+l})~|~0\le k\le 2^m-1, l\le K, \,\, |t-s|\le \delta \}.\) Here \(M\) should be larger than the number \(C_1\) which is determined by the case \(K=1\).

We consider the case \(K+1\). We rewrite \(s=t_k\) and \(t=t_{k+K+1}\). Choose maximum \(u=t_l\) satisfying \(|u-s|\le |t-s|/2\). Then \(|t-t_{l+1}|\le |t-s|/2\) holds. Note that \(l-k\le K\) and \(K+1-(l+1)\le K\). Hence by the assumption, we have \[\begin{align} \max\{|I_{s,u}|, |I_{t_{l+1},t}|\} & \le M\left|\frac{t-s}{2}\right|^{\lambda+H^-},\tag{61}\\ \max\{|Y^{m,\rho}_u-Y^{m,\rho}_s|,|Y^{m,\rho}_{t}-Y^{m,\rho}_{t_{l+1}}|\} &\le M\left|\frac{t-s}{2}\right|^{\lambda+H^-} +C|t-s|^{H^{-}}.\tag{62} \end{align}\] Next we estimate \((\delta I)_{s,u,t}=I_{s,t}-I_{s,u}-I_{u,t}\). Denote by \((\delta I)_{s,u,t}^\sigma\), \((\delta I)_{s,u,t}^b\) and \((\delta I)_{s,u,t}^c\) the terms in \((\delta I)_{s,u,t}\) being concerned with \(\sigma\), \(b\) and \(c\), respectively. Then \[\begin{align} (\delta I)_{s,u,t}^b &= -b(Y^{m,\rho}_s)(t-s) +b(Y^{m,\rho}_s)(u-s) +b(Y^{m,\rho}_u)(t-u)\\ &= \{b(Y^{m,\rho}_u)-b(Y^{m,\rho}_s)\}(t-u),\\ (\delta I)_{s,u,t}^c &= \rho \{c(Y^{m,\rho}_u)-c(Y^{m,\rho}_s)\} d^m_{u,t} \end{align}\] and \[\begin{align} (\delta I)_{s,u,t}^\sigma &= \{\sigma(Y^{m,\rho}_u)-\sigma(Y^{m,\rho}_s)\}B_{u,t} -((D\sigma)[\sigma])(Y^{m,\rho}_s)[{\mathbb{B}}_{s,t}-{\mathbb{B}}_{s,u}-{\mathbb{B}}_{u,t}]\\ &\qquad -\big\{((D\sigma)[\sigma])(Y^{m,\rho}_s) -((D\sigma)[\sigma])(Y^{m,\rho}_u)\big\}{\mathbb{B}}_{u,t}\\ &= \big\{\sigma(Y^{m,\rho}_u)-\sigma(Y^{m,\rho}_s)-D\sigma(Y^{m,\rho}_s)[Y^{m,\rho}_u-Y^{m,\rho}_s]\big\}B_{u,t}\\ &\qquad + D\sigma(Y^{m,\rho}_s) [ I_{s,u} +((D\sigma)[\sigma])(Y^{m,\rho}_s){\mathbb{B}}_{s,u} +\rho c(Y^{m,\rho}_s)d^m_{s,u} +b(Y^{m,\rho}_s)(u-s) ] B_{u,t}\\ &\qquad - \big\{ ((D\sigma)[\sigma])(Y^{m,\rho}_s) - ((D\sigma)[\sigma])(Y^{m,\rho}_u) \big\} {\mathbb{B}}_{u,t}. \end{align}\] Here we used Chen’s identity and definition of \(I_{s,u}\). By (61 ) and (62 ), we obtain \[\begin{align} |(\delta I)_{s,u,t}| &\le C\big\{1+M\delta^{H^-}+(M\delta^{H^-})^2\big\}|t-s|^{\lambda+H^-}. \end{align}\] Similarly, we obtain \(|(\delta I)_{t_l,t_{l+1},t}| \le C|t-s|^{3H^{-}}\). By \[\begin{align} I_{s,t} &= I_{s,u}+I_{t_l,t_{l+1}}+I_{t_{l+1},t} +(\delta I)_{t_{l},t_{l+1},t} +(\delta I)_{s,u,t}, \end{align}\] we have \(|I_{s,t}|\leq f(C,M,\delta)|t-s|^{\lambda+H^-}\), where \[\begin{align} f(C,M,\delta) = 2^{1-(\lambda+H^-)}M +C\big\{1+M\delta^{H^-}+(M\delta^{H^-})^2\big\}. \end{align}\] Note that the function \(f\) and \(C\) do not depend on \(K\). Let \((M,\delta)\) be a pair such that \(f(C,M,\delta)\le M\) holds and \(M\) is greater than or equal to the maximum of three constants \(C\) stated in 14 , 26 , and 27 . Then (60 ) holds for \((C_1,\delta)=(M,\delta)\). One choice is as follows. \[\begin{align} M=\frac{3C}{1-2^{1-(\lambda+H^-)}},\quad \delta= \min\bigg\{\bigg(\frac{3C}{1-2^{1-(\lambda+H^-)}}\bigg)^{-\frac{1}{H^-}}, 1\bigg\}, \end{align}\] where \(C\) is greater than or equal to the maximum of three constants \(C\) stated in 14 , 26 , and 27 . This completes the proof. ◻

Lemma 9. Assume that Condition  (1) holds and let \(\omega\in \Omega_0\). Let \(\lambda_1\) be a positive number satisfying \(\lambda_1+H^->1\). Set \(\lambda=\min\{\lambda_1,2H^-\}\). Then there exist a positive number \(C_2\) which depends on \(\sigma,b,c\), \(C(B)\) and \(\|d^m\|_{\lambda_1}\) polynomially such that \[\begin{align} |I_{s,t}|\le C_2|t-s|^{\lambda+H^-}, \qquad s,t\in D_m. \end{align}\]

Proof. Below, \(C\) denote constants depending only on \(\sigma, b, c\), \(C(B)\) and \(\|d^m\|_{\lambda_1}\) polynomially. We have proved the case where \(s,t\) with \(t-s\le \delta\). Suppose \(t-s>\delta\). In this case, from the definition of \(I_{s,t}\) and \((\delta^{-1}|t-s|)^\lambda \geq 1\), we have \[\begin{align} |I_{s,t}| \leq |Y^{m,\rho}_{s,t}| +C|t-s|^{H-} \leq |Y^{m,\rho}_{s,t}| +C\delta^{-1}|t-s|^{\lambda+H^-}. \end{align}\] Here we wrote \(Y^{m,\rho}_{s,t}=Y^{m,\rho}_t-Y^{m,\rho}_s\). In what follows, we will give an estimates of \(|Y^{m,\rho}_{s,t}|\).

First, we consider the case \(2^{-m}\ge \delta\). For \(s=2^{-m}k<t=2^{-m}l\), we have \[\begin{align} |Y^{m,\rho}_{s,t}|&=\left|\sum_{i=k+1}^lY^{m,\rho}_{\tau^m_{i-1},\tau^m_i}\right| \le C (l-k)\Delta_m^{H^-} =C (2^m)^\lambda(l-k)^{1-(\lambda+H^-)}|t-s|^{\lambda+H^-}. \end{align}\] Noting \((2^m)^\lambda\le \delta^{-\lambda}\), we obtain \(|Y^{m,\rho}_{s,t}|\le C \delta^{-\lambda}|t-s|^{\lambda+H^-}\).

We next consider the case \(2^{-m}<\delta\). Let \(\tau^m_K=\max\{\tau^m_k~|~\tau^m_k\le \delta\}\). Then \(2^{-1}\delta\le \tau^m_K\). Let \(s_i=s+i\tau^m_K\) \((0\leq i\leq N-1)\), where \(N\) is a positive integer such that \(0\le t-s_{N-1}<\tau^m_K\). For notational simplicity, we set \(s_N=t\). Then we have \(N\le (\tau^m_K)^{-1}(t-s)+1\le 2(t-s)(\tau^m_K)^{-1} \le 4\delta^{-1}(t-s).\) By the estimate in Lemma 8, we have \[\begin{align} |Y^{m,\rho}_{s_{i-1},s_i}| \leq C \big\{ |t-s|^{\lambda+H^-}+|t-s|^{H^-}+ |t-s|^{2H^-}+|t-s|^\lambda+|t-s| \big\} \leq C |t-s|^{H^-}. \end{align}\] Hence \[\begin{align} |Y^{m,\rho}_{s,t}| \le \sum_{i=1}^N|Y^{m,\rho}_{s_i}-Y^{m,\rho}_{s_{i-1}}| \le \delta^{-1}|t-s| \cdot C |t-s|^{H^-}. \end{align}\] Since \(1>\lambda\), we obtain \(|Y^{m,\rho}_t-Y^{m,\rho}_s|\le C\delta^{-1}|t-s|^{\lambda+H^-}\). Since \(\delta^{-1}\) depends on \(\sigma\), \(b\), \(c\), \(C(B)\), \(\|d^m\|_{\lambda_1}\) polynomially, we complete the proof. ◻

For \(f\in C^2_b({\mathbb{R}}^n, \mathcal{L}({\mathbb{R}}^d,{\mathbb{R}}^K))\), \(g\in C^1_b({\mathbb{R}}^n, {\mathbb{R}}^K)\), and \(h\in C^1_b({\mathbb{R}}^n,\mathcal{L}({\mathbb{R}}^d\otimes{\mathbb{R}}^d,{\mathbb{R}}^K))\), and \(s,t\in D_m\) with \(s<t\), we define an \({\mathbb{R}}^K\)-valued random variable by \[\begin{align} \Xi(f,g,h)_{s,t}&= f(Y^{m,\rho}_{s})B_{s,t} +(Df)[\sigma](Y^{m,\rho}_{s}){\mathbb{B}}_{s,t} +g(Y^{m,\rho}_{s})(t-s) +h(Y^{m,\rho}_{s})d^m_{s,t}, \end{align}\] where \((Df)[\sigma](y)[v\otimes w] = Df(y)[\sigma(y)v]w\) for \(y\in{\mathbb{R}}^n\), \(v,w\in{\mathbb{R}}^d\) (see also 5 ). For a sub-partition \({\cal P}=\{u_i\}_{i=0}^l\subset D_m\) \((s=u_0,t=u_l)\), let \[\begin{align} I(f,g,h ; {\cal P})_{s,t}= \sum_{i=0}^l \Xi(f,g,h)_{u_{i-1},u_i}. \end{align}\]

Lemma 10. Assume that Condition  (1) holds and let \(\omega\in \Omega_0\). Let \(\lambda_1\) be a positive number satisfying \(\lambda_1+H^->1\). Set \(\lambda=\min\{\lambda_1,2H^-\}\). Then \[\begin{align} |I(f,g,h ; {\cal P})_{s,t}-\Xi(f,g,h)_{s,t}|\le C|t-s|^{\lambda+H^-}, \end{align}\] where \(C\) depends on \(\sigma,b,c,C(B),\|d^m\|_{\lambda_1}\) polynomially.

Proof. Let \(I_{st}\) be the function defined in 59 . \[\begin{align} \delta \Xi(f,g,h)_{s,u,t} = \Xi(f,g,h)_{s,t}-\Xi(f,g,h)_{s,u}-\Xi(f,g,h)_{u,t}\\&= -\left\{f(Y^{m,\rho}_u)-f(Y^{m,\rho}_s) -(Df)(Y^{m,\rho}_s)[Y^{m,\rho}_u-Y^{m,\rho}_s]\right\}B_{u,t}\\ &\qquad -(Df)(Y^{m,\rho}_s) \left[ I_{s,u} +((D\sigma)[\sigma])(Y^{m,\rho}_s){\mathbb{B}}_{s,u} +\rho c(Y^{m,\rho}_{s})d^m_{s,u}+b(Y^{m,\rho}_s)(u-s) \right] B_{u,t}\\ &\qquad + \left\{ (Df)(Y^{m,\rho}_s)[\sigma(Y^{m,\rho}_s)]-(Df)(Y^{m,\rho}_u)[\sigma(Y^{m,\rho}_u)] \right\}{\mathbb{B}}_{u,t}\nonumber\\ &\qquad +\left\{g(Y^{m,\rho}_s)-g(Y^{m,\rho}_u)\right\}(t-u) +\left\{h(Y^{m,\rho}_s)-h(Y^{m,\rho}_u)\right\}d^m_{u,t}.\end{align}\] Hence \(|\delta I(f,g,h)_{s,u,t}|\le C|t-s|^{\lambda+H^-}\). By a standard argument (for example, use the sewing lemma (see [14])), we complete the proof of the lemma. ◻

4.2 Estimates of \(\tilde{J}^{m,\rho}_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}\) on \(\Omega_0^{(m,d^m)}\)↩︎

We next proceed to the estimate of \(\tilde{J}^{m,\rho}_t(\omega)\) and their inverse. From now on, we always assume that \(\omega\in \Omega_0^{(m,d^m)}\) and \(m\) satisfies 55 ; see Remark 24. For \(\omega\in \Omega_0^{(m,d^m)}\), both estimates \(\|d^m(\omega)\|_{2H^-}\le 1\) and \(\|d^m(\omega)\|_{\lambda_1}\le 1\) hold. However note that we use one or the other only of the two estimates in the proofs of some statements in this section. Since \(\tilde{J}^{m,\rho}\) is also a solution to a discrete RDE, one may expect similar estimates for \(\tilde{J}^{m,\rho}\) to \(Y^{m,\rho}\). However, the coefficient of the RDE of \(\tilde{J}^{m,\rho}\) is unbounded, we cannot apply the same proof as the one of \(Y^{m,\rho}\) and we need to prove the boundedness of \(\tilde{J}^{m,\rho}\) in advance. We give an estimate of \(\tilde{J}^{m,\rho}\) by combining the group property of \(\tilde{J}^{m,\rho}\) and a similar argument to the estimate of \(Y^{m,\rho}\). The difference from \(Y^{m,\rho}\) is that we use the estimate \(\|d^m(\omega)\|_{2H^-}\le 1\) and the variation norm of \((B,{\mathbb{B}})\) (see Definition 5) to obtain the boundedness of \(\tilde{J}^{m,\rho}\). After obtaining the boundedness, we see estimates on \(\tilde{J}^{m,\rho}_t\) and their inverse by using the estimate \(\|d^m(\omega)\|_{\lambda_1}\le 1\) and the Hölder norm of \((B,{\mathbb{B}})\).

First, we observe the following. For \(s\le t\), \(s,t,\tau\in D_m\) with \(t+\tau\le 1\), let us define \[\begin{align} I_{s,t}(Y^{m,\rho}_\tau,\theta_\tau B)&= \tilde{J}^{m,\rho}_{t}(Y^{m,\rho}_\tau,\theta_\tau B)-\tilde{J}^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B) \\ &\qquad\quad- (D\sigma)(Y^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B)) [\tilde{J}^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B)] (\theta_\tau B)_{s,t}\\ &\qquad\quad -D\left((D\sigma)[\sigma]\right)(Y^{m,\rho}_{s}(Y^{m,\rho}_\tau, \theta_\tau B)) [\tilde{J}^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B)] (\theta_\tau{\mathbb{B}})_{s,t}\\ &\qquad\quad -\rho (Dc)(Y^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B)) [\tilde{J}^{m,\rho}_{s}(Y^{m,\rho}_\tau,\theta_\tau B)] (\theta_\tau d^m)_{s,t}\\ &\qquad\quad -(Db)(Y^{m,\rho}_{s}(Y^{m,\rho}_\tau ,\theta_\tau B)) [\tilde{J}^{m,\rho}_{s}(Y^{m,\rho}_\tau ,\theta_\tau B)](t-s). \end{align}\] We may write \(I_{s,t}(\xi,B)=I_{s,t}\) for simplicity. Note that \[\begin{align} I_{0,t-u}(Y^{m,\rho}_u,\theta_uB)&= \tilde{J}^{m,\rho}_{t-u}(Y^{m,\rho}_u,\theta_uB) -I-(D\sigma)(Y^{m,\rho}_u)[I]B_{u,t} -D((D\sigma)[\sigma])(Y^{m,\rho}_u)[I]{\mathbb{B}}_{u,t}\nonumber\\ &\quad - \rho (Dc)(Y^{m,\rho}_u)[I]d^m_{u,t}- (Db)(Y^{m,\rho}_u)[I](t-u),\label{I0t-u} \end{align}\tag{63}\] where \(I\) denotes the identity operator and we refer the notation \(D((D\sigma)[\sigma])(Y^{m,\rho}_u)[I]{\mathbb{B}}_{u,t}\) to (48 ). By (63 ), if \(I_{0,t-u}(Y^{m,\rho}_u,\theta_uB)\) and \(t-u\) is sufficiently small, then we see \(\tilde{J}^{m,\rho}_{t-u}(Y^{m,\rho}_u,\theta_uB)\) is invertible.

Lemma 11. Let \(s,t,\tau, \tau' \in D_m\) with \(\tau'\le s\le t\) and \(t+\tau\le 1\). Then \[\begin{align} I_{s,t}(Y^{m,\rho}_{\tau},\theta_\tau B)& =I_{0,t-s}(Y^{m,\rho}_{s+\tau},\theta_{s+\tau}B)\tilde{J}^{m,\rho}_{s} (Y^{m,\rho}_{\tau},\theta_{\tau}B)\nonumber\\ &=I_{s-\tau',t-\tau'}(Y^{m,\rho}_{\tau'+\tau},\theta_{\tau'+\tau} B) \tilde{J}^{m,\rho}_{\tau'} (Y^{m,\rho}_{\tau},\theta_\tau B). \end{align}\]

Proof. These follows from the definition and the following identity. Let \(u\ge s\). \[\begin{align} Y^{m,\rho}_u(Y^{m,\rho}_{\tau},\theta_{\tau}B) &= Y^{m,\rho}_{u+\tau}(\xi,B) = Y^{m,\rho}_{u-s}(Y^{m,\rho}_{s+\tau},\theta_{s+\tau} B),\\ \tilde{J}^{m,\rho}_u(Y^{m,\rho}_\tau,\theta_\tau B) &= \tilde{J}^{m,\rho}_{u-s}(Y^{m,\rho}_{s+\tau},\theta_{s+\tau}B) \tilde{J}^{m,\rho}_s(Y^{m,\rho}_\tau,\theta_\tau B),\\ (\theta_\tau \Xi)_{u,t} &= (\theta_{s+\tau}\Xi)_{u-s,t-s} \qquad \text{for} \qquad \Xi=B,{\mathbb{B}},d^m. \end{align}\] ◻

Definition 5. Let \(p=(H^-)^{-1}\). For \((1,B_{s,t},{\mathbb{B}}_{s,t})_{0\le s\le t\le 1}\), we define \[\begin{align} w(s,t)&=\|B\|_{[s,t], p\mathchar`-var}^{p}+ \|{\mathbb{B}}\|_{[s,t], \frac{p}{2}\mathchar`-var}^{\frac{p}{2}},\qquad 0\le s\le t\le 1, \end{align}\] where \(\|~\|_{[s,t], r\mathchar`-var}\) denotes the \(r\)-variation norm. Also we define \(\tilde{w}(s,t)=w(s,t)+|t-s|\).

Note that the variables \(s,t\) move in \([0,1]\) and \(B\) and \({\mathbb{B}}\) are random variables defined on \(\Omega_0\) and so are \(w(s,t)\) and \(\tilde{w}(s,t)\).

We give estimates for \(\tilde{J}^{m,\rho}\) and \(I_{s,t}(Y^{m,\rho}_\tau,\theta_\tau B)\) by using \(\tilde{w}\). First we note that the following estimate.

Lemma 12. Assume that Condition  (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). There exist \(0<\delta\le 1\) and \(C_3>0\) such that for all \(s,t\in D_m\) with \(0\le s<t\le 1\) and \(\tilde{w}(s,t)\le\delta\), the following estimate holds: \[\begin{align} & \left|Y^{m,\rho}_t-Y^{m,\rho}_s- \sigma(Y^{m,\rho}_s)B_{s,t}-((D\sigma)[\sigma])(Y^{m,\rho}_s){\mathbb{B}}_{s,t}- \rho c(Y^{m,\rho}_s)d^m_{s,t}-b(Y^{m,\rho}_s)(t-s) \right|\\ &\qquad \le C_3\tilde{w}(s,t)^{3H^-}, \end{align}\] where \(\delta\) and \(C_3\) are constants depending only on \(\sigma, b, c, H^-\).

Proof. The proof of this lemma is similar to that of Lemma 8 and is done by induction. The difference is that we do not use 14 and 27 and use 15 and 26 . Here we give a sketch of the proof. Below, \(\tau^m_i=t_i\) and \(C\) denotes a constant depending only on \(\sigma\), \(b\), \(c\), and \(H^-\) polynomially.

The first step of the induction is as follows. Note \(I_{t_k,t_{k+1}} = (1-\rho)\epsilon^m_{t_k,t_{k+1}} +\rho\hat{\epsilon}^m_{t_k,t_{k+1}}\). The estimates 15 and 26 imply \(|\epsilon^m_{t_{k-1},t_k}| + |\hat{\epsilon}^{m}_{t_{k-1},t_k}| \le C \tilde{w}(t_{k-1},t_k)^{3H^-}\) for all \(1\le k\le 2^m\) and \(\omega\in \Omega_0^{(m)}\). Hence \(|I_{t_k,t_{k+1}}|\leq C\tilde{w}(t_{k-1},t_k)^{3H^-}\). The induction works well by noting \[\begin{align} |B_{s,t}|&\le \tilde{w}(s,t)^{H^-}, & |{\mathbb{B}}_{s,t}|&\le \tilde{w}(s,t)^{2H^-}, & |d^m_{s,t}|&\le \tilde{w}(s,t)^{2H^-} \qquad \text{for all}\qquad s,t\in D_m. \end{align}\] The last estimate above follows from \(\omega\in\Omega_0^{(m,d^m)}\). For example, we need to change the sentence “maximum \(u=t_l\) satisfying \(|u-s|\le |t-s|/2\)” to “maximum \(u=t_l\) satisfying \(\tilde{w}(s,u)\le \tilde{w}(s,t)/2\)”. For this \(l\), we see \(\tilde{w}(t_{l+1},t)\le \frac{1}{2}\tilde{w}(s,t)\). We omit the details. ◻

Lemma 13. Assume that Condition  (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). There exist \(0<\delta\le 1\) and \(C_4>0\) such that for any \(t,\tau\in D_m\) with \(\tilde{w}(\tau,\tau+t)\le \delta\) and \(t+\tau\le 1\), the following estimate holds. \[\begin{align} |I_{0,t}(Y^{m,\rho}_{\tau},\theta_\tau B)|\le C_4 \tilde{w}(\tau,\tau+t)^{3H^{-}},\label{Ist} \end{align}\tag{64}\] where \(\delta\) and \(C_4\) are constants depending only on \(\sigma, b, c, H^{-}\).

Proof. Below, we write \(\tilde{w}_{\tau}(s,t)=\tilde{w}(s+\tau,t+\tau)\) and \(C\) is a constant depending only on \(\sigma,b,c, H^-\) which may change line by line. The proof is similar to that of Lemma 8. We take \(\delta\) smaller than \(\delta\) in Lemma 12. For simplicity we write \(t_k=\tau^m_k\). It suffices to consider the case where \(\tau\le 1-2^{-m}\). We consider the following claim depending on a positive integer \(K\).

(Claim \(K\)) (64 ) holds for all \(\tau\) and \(t_k\) satisfying \(\tau+t_k\le 1\), \(\tilde{w}_\tau(0,t_k)\le \delta\) and \(1\le k\le K\).

Since \(I_{0,t_1}= I_{0,t_1}(Y^{m,\rho}_{\tau},\theta_{\tau}B)=0\) holds for all \(\tau\), (Claim 1) holds for \(C_4=0\) and any \(\delta\). We assume (Claim \(K\)) holds and we will find the condition on \(C_4\) and \(\delta\) independent of \(K\) under which (Claim \(K+1\)) holds. Assume the case \(K\) holds for a positive constant \(C_4\) and \(\delta\). Suppose \(\tau+t_{K+1}\le 1\) and \(\tilde{w}_\tau(0,t_{K+1})\le \delta\), where \(K\ge 1\). Define \(0\le t_l<t_{K+1}\) as the maximum number such that \(\tilde{w}_{\tau}(0,t_l)\le \tilde{w}_{\tau}(0,t_{K+1})/2\). On the other hand, for \(t_{l+1}\), we have \(\tilde{w}_{\tau}(t_{l+1},t_{K+1})\le \tilde{w}_{\tau}(0,t_{K+1})/2\). We will write \(u=t_l\) and \(t=t_{K+1}\). By (Claim \(K\)), we have \[\begin{align} \tag{65} |I_{0,u}(Y^{m,\rho}_\tau,\theta_\tau B)| &\le C_4(\tilde{w}_{\tau}(0,t)/2)^{3H^{-}},\\ \tag{66} |I_{0,t-t_{l+1}}(Y^{m,\rho}_{t_{l+1}+\tau},\theta_{t_{l+1}+\tau}B)| &\le C_4(\tilde{w}_{\tau}(0,t)/2)^{3H^{-}}. \end{align}\] The estimate (65 ) implies \[\begin{align} \tag{67}|\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)-I| &\le C_4(\tilde{w}_{\tau}(0,t)/2)^{3H^{-}} +C\tilde{w}_{\tau}(0,t)^{H^{-}} +C\tilde{w}_{\tau}(0,t)^{2H^{-}} \\ &\le \{C_4(\delta/2)^{2H^{-}}+C\} \tilde{w}_{\tau}(0,t)^{H^{-}}, \\ \tag{68} | \tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B) -I -(D\sigma)(Y^{m,\rho}_{\tau})B_{\tau,u+\tau} | \le \{C_4(\delta/2)^{H^{-}}+C\} \tilde{w}_{\tau}(0,t)^{2H^{-}}. \end{align}\] For simplicity, we write \(I_{0,t}=I_{0,t}(Y^{m,\rho}_\tau,\theta_\tau B)\) and set \((\delta I)_{0,u,t}=I_{0,t}-I_{0,u}-I_{u,t}\). Hereafter we will estimate \((\delta I)_{0,u,t}\) and \(I_{u,t}\). By the results on them and the inductive assumption, we will obtain a bound of \(I_{0,t}\)

First we consider \((\delta I)_{0,u,t}\). Denote by \((\delta I)_{0,u,t}^\sigma\), \((\delta I)_{0,u,t}^b\) and \((\delta I)_{0,u,t}^c\) the terms in \((\delta I)_{0,u,t}\) being concerned with \(\sigma\), \(b\) and \(c\), respectively. Then we have \[\begin{align} (\delta I)_{0,u,t}^b &= -(Db)(Y^{m,\rho}_\tau)[I]t +(Db)(Y^{m,\rho}_\tau)[I]u\\ &\phantom{=}\qquad +(Db)(Y^{m,\rho}_u(Y^{m,\rho}_\tau,\theta_\tau B))[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)](t-u)\\ &= \big\{ (Db)(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)] - (Db)(Y^{m,\rho}_\tau)[I] \big\} (t-u)\\ (\delta I)_{0,u,t}^c &= \rho \big\{ (Dc)(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_u(Y^{m,\rho}_{\tau},\theta_{\tau}B)] - (Dc)(Y^{m,\rho}_\tau)[I] \big\} d^m_{u+\tau,t+\tau} \end{align}\] and \[\begin{align} (\delta I)_{0,u,t}^\sigma &= -(D\sigma)(Y^{m,\rho}_{\tau})[I]B_{u+\tau,t+\tau} - D((D\sigma)[\sigma])(Y^{m,\rho}_{\tau})[I] \left( {\mathbb{B}}_{\tau,\tau+t}-{\mathbb{B}}_{\tau,\tau+u} \right) \\ &\qquad\qquad +(D\sigma)(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)] B_{u+\tau,t+\tau} \\ &\qquad\qquad + D((D\sigma)[\sigma])(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)] {\mathbb{B}}_{u+\tau,t+\tau}. \end{align}\] Here by getting the first and third terms together, we have \[\begin{gather} (D\sigma)(Y^{m,\rho}_{u+\tau}) \Big[ \tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)-I-D\sigma(Y^{m,\rho}_\tau)[I]B_{\tau,\tau+u} \Big] B_{u+\tau,t+\tau}\\ + \Big\{ (D\sigma)(Y^{m,\rho}_{u+\tau})[I] - (D\sigma)(Y^{m,\rho}_{\tau})[I] - D(D\sigma)(Y^{m,\rho}_\tau)[\sigma(Y^{m,\rho}_{\tau})B_{\tau,u+\tau}] \Big\} B_{u+\tau,t+\tau}\\ + \uwave{ (D\sigma)(Y^{m,\rho}_{u+\tau}) \left[ D\sigma(Y^{m,\rho}_\tau)[I]B_{\tau,\tau+u} \right] B_{u+\tau,t+\tau} } + \uwave{D(D\sigma)(Y^{m,\rho}_\tau)[\sigma(Y^{m,\rho}_{\tau})B_{\tau,u+\tau}]B_{u+\tau,t+\tau}}. \end{gather}\] Because of Chen’s identity, the summation of the second and fourth terms gives \[\begin{gather} \Big\{ D((D\sigma)[\sigma])(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)] - D((D\sigma)[\sigma])(Y^{m,\rho}_{\tau})[I] \Big\} {\mathbb{B}}_{u+\tau,t+\tau}\\ + \uwave{ ( - D((D\sigma)[\sigma])(Y^{m,\rho}_{\tau})[I] \left\{ B_{\tau,\tau+u}\otimes B_{\tau+u,\tau+t} \right\} ) }. \end{gather}\] Since the summation of terms with    vanishes due to 48 , we have \[\begin{align} (\delta I)_{0,u,t}^\sigma &= (D\sigma)(Y^{m,\rho}_{u+\tau}) \Big[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)-I-(D\sigma)(Y^{m,\rho}_{\tau})B_{\tau,u+\tau}\Big] B_{u+\tau,t+\tau}\\ &\qquad + \Big\{ (D\sigma)(Y^{m,\rho}_{u+\tau}) - (D\sigma)(Y^{m,\rho}_{\tau})-D(D\sigma)(Y^{m,\rho}_\tau)[\sigma(Y^{m,\rho}_{\tau})B_{\tau,u+\tau}] \Big\} B_{u+\tau,t+\tau} \\ &\qquad + \Big\{ D((D\sigma)[\sigma])(Y^{m,\rho}_{u+\tau})[\tilde{J}^{m,\rho}_{u}(Y^{m,\rho}_\tau,\theta_\tau B)] - D((D\sigma)[\sigma])(Y^{m,\rho}_{\tau})[I] \Big\} {\mathbb{B}}_{u+\tau,t+\tau}. \end{align}\] Thus, combining Lemma 12, (67 ) and (68 ) , we get \[\begin{align} |(\delta I)_{0,u,t}^\sigma| &\le C\tilde{w}_{\tau}(0,t)^{3H^{-}} + C\big\{ 1 + C_4(\delta/2)^{H^-} \big\}\tilde{w}_{\tau}(0,t)^{3H^-},\\ |(\delta I)_{0,u,t}^b| &\le C\big\{ 1 + C_4(\delta/2)^{2H^-} \big\}\tilde{w}_{\tau}(0,t)^{1+H^-},\\ |(\delta I)_{0,u,t}^c| &\le C\big\{ 1 + C_4(\delta/2)^{2H^-} \big\}\tilde{w}_{\tau}(0,t)^{3H^-}. \end{align}\] Hence, \[\begin{align} |(\delta I)_{0,u,t}| &\le C\big\{1+C_4\delta^{H^-}\big\}\tilde{w}_{\tau}(0,t)^{3H^-}. \end{align}\]

We estimate \(I_{u,t}\). We have \(I_{u,t}=I_{t_l,t} = (\delta I)_{t_l,t_{l+1},t}+ I_{t_l,t_{l+1}}+I_{t_{l+1},t}\). It is clear that \(I_{t_l,t_{l+1}}=0\). First we consider \((\delta I)_{t_l,t_{l+1},t}\). Using Lemma 11 and (67 ), we get \[\begin{align} |(\delta I)_{t_l,t_{l+1},t}| &= \big| \big\{ I_{0,t-t_l}(Y^{m,\rho}_{t_l+\tau},\theta_{t_l+\tau}B) -I_{0,t_{l+1}-t_l}(Y^{m,\rho}_{t_l+\tau}, \theta_{t_l+\tau}B)\\ &\qquad\qquad\qquad -I_{t_{l+1}-t_l,t-t_l}(Y^{m,\rho}_{t_l+\tau},\theta_{t_l+\tau}B) \big\} \cdot \tilde{J}^{m,\rho}_{t_l}(Y^{m,\rho}_\tau,\theta_\tau B) \big|\\ &\le C \big\{ 1 + C_4\delta^{H^-} \big\} \tilde{w}_{\tau+t_l}(0,t-t_l)^{3H^-} \big|\tilde{J}^{m,\rho}_{t_l}(Y^{m,\rho}_\tau,\theta_\tau B)\big|, \end{align}\] where we have used a similar estimate of \((\delta I)_{0,t_{l+1}-t_l,t-t_l}\) to \((\delta I)_{0,u,t}\) and note \(\tilde{w}_{\tau+t_l}(0,t-t_l)=\tilde{w}_{\tau}(t_l,t)\leq \tilde{w}_{\tau}(0,t)\). Next we consider \(I_{t_{l+1},t}\). Lemma 11 implies \[\begin{align} I_{t_{l+1},t} &= I_{0,t-t_{l+1}}(Y^{m,\rho}_{t_{l+1}+\tau},\theta^m_{t_{l+1}+\tau}B) \tilde{J}^{m,\rho}_{t_{l+1}}(Y^{m,\rho}_{\tau},\theta_\tau B)\\ &= I_{0,t-t_{l+1}}(Y^{m,\rho}_{t_{l+1}+\tau},\theta^m_{t_{l+1}+\tau}B) E^{m,\rho}(Y^{m,\rho}_{t_l+\tau},\theta_{t_l+\tau}B) \tilde{J}^{m,\rho}_{t_l}(Y^{m,\rho}_\tau,\theta_\tau B). \end{align}\] By 66 and the definition of \(E^{m,\rho}\) (see 50 ), we obtain \[\begin{align} |I_{t_{l+1},t}| &\le C_4 \left( \frac{1}{2} \tilde{w}_{\tau}(0,t) \right)^{3H^{-}} \big\{ 1 + C\tilde{w}_\tau(0,t)^{H^-} \big\} \big|\tilde{J}^{m,\rho}_{t_l}(Y^{m,\rho}_\tau,\theta_\tau B)\big|. \end{align}\] Hence noting \(|\tilde{J}^{m,\rho}_{t_l}(Y^{m,\rho}_\tau,\theta_\tau B)| \leq 1+C\{1+C_4\delta^{H^-}\}\), we have \[\begin{align} |I_{u,t}| &\leq \big\{ C \big\{ 1+C_4\delta^{H^-} \big\} + C_4 2^{-3H^-} \big\{ 1+C\delta^{H^-} \big\} \big\} \big\{ 1+C\{ 1 + C_4\delta^{H^-} \} \big\} \tilde{w}_\tau(0,t)^{3H^-}\\ &\leq \big\{ C_4 2^{-3H^-} + C \big\{ 1 + C_4\delta^{H^-} \big\} \big\} \big\{ 1+C\{ 1 + C_4\delta^{H^-} \} \big\} \tilde{w}_\tau(0,t)^{3H^-}\\ &\leq \big\{ C_4 2^{-3H^-} + C \big\{ 1+C_4\delta^{H^-}+(C_4\delta^{H^-})^2 \big\} \big\} \tilde{w}_\tau(0,t)^{3H^-}. \end{align}\]

Consequently, noting \(I_{0,t}=I_{0,u}+(\delta I)_{0,u,t}+I_{u,t}\), we obtain \[\begin{align} |I_{0,t}| &\le \big\{ 2C_4 2^{-3H^{-}} + C \big\{1+(C_4\delta^{H^-})+(C_4\delta^{H^-})^2\big\} \big\}\tilde{w}_{\tau}(0,t)^{3H^{-}}. \end{align}\] Hence if \(C_4\) and \(\delta\) satisfies \(C_4 2^{1-3H^{-}} + C \big\{ 1+(C_4\delta^{H^-})+(C_4\delta^{H^-})^2 \big\} \le C_4\), then (64 ) holds in the case of \(K+1\). One choice of \(C_4, \delta\) is \[\begin{align} C_4 &= \frac{3C}{1-2^{1-3H^-}}, & \delta &= \min \bigg\{ \bigg(\frac{3C}{1-2^{1-3H^-}}\bigg)^{-\frac{1}{H^-}}, 1 \bigg\}. \end{align}\]

Under this choice, we see that (64 ) holds for any \(t,\tau\in D_m\) with \(\tilde{w}(\tau,\tau+t)\le \delta\) and \(t+\tau\le 1\). This completes the proof. ◻

In order to obtain \(L^p\) estimate in Theorem 12, we need the estimate obtained by Cass-Litterer-Lyons [13]. To this end, we introduce the number \(N_{\beta}(w)\) which is defined for any control function \(w\) and positive number \(\beta\). We already used the notation \(w\) in Definition 5 and so this is an abuse in a certain sense. For a control function \(w\) and a positive number \(\beta\), let us define \(N_{\beta}(w)\) and a nondecreasing sequence \(\{\sigma_i\}_{i=0}^{\infty}\subset [0,1]\) as follows.

  1. \(\sigma_0=0\).

  2. Let \(i\ge 0\) and write \(A_i=\{s\in [0,1]~|~s\ge \sigma_{i},w(\sigma_{i},s)\ge \beta\}\). Set \(\sigma_{i+1}=\inf A_i\) (resp. \(1\)) if \(A_i\neq \emptyset\) (resp. \(A_i=\emptyset\)).

  3. \(N_{\beta}(w)=\sup\{i\ge 0~|~\sigma_i<1\}\).

We have the following.

Lemma 14. Let \(w, w'\) be any control functions and \(\beta,\beta'>0\).

  1. There exist positive constants \(C_{\beta,\beta'}, C'_{\beta,\beta'}\) which are independent of \(w\) such that \[C_{\beta,\beta'}(N_{\beta'}(w)+1)\le N_{\beta}(w)+1\le C'_{\beta,\beta'}(N_{\beta'}(w)+1).\]

  2. If \(w(s,t)\le w'(s,t)\) \((0\le s\le t\le 1)\) holds, then \(N_{\beta}(w)\le N_{\beta}(w')\).

  3. Let \(\tilde{w}(s,t)=w(s,t)+|t-s|\) \((0\le s\le t\le 1)\). Then for any \(\beta\ge 3\), we have \(N_{\beta}(\tilde{w})\le N_1(w)\).

Proof. We show (1). We use \(\sigma^\beta_i\) to denote the dependence of \(\sigma_i\) on \(\beta\). Assume \(\beta'<\beta\). Then \(\sigma^{\beta'}_i\leq \sigma^\beta_i\) for all \(i\geq 0\), which implies \(N_{\beta'}(w)\geq N_\beta(w)\). Conversely, by setting \(\Lambda_i = \{j:\sigma^\beta_i\leq \sigma^{\beta'}_j,\sigma^{\beta'}_{j+1}\leq \sigma^\beta_{i+1}\}\) for \(0\leq i\leq N_\beta(w)-1\), we have \[\begin{align} \beta = w(\sigma^\beta_i,\sigma^\beta_{i+1}) \geq \sum_{j\in \Lambda_i} w(\sigma^{\beta'}_j,\sigma^{\beta'}_{j+1}) = \sharp \Lambda_i \beta'. \end{align}\] Since the number of \(j\) such that \(\sigma^\beta_i\in(\sigma^{\beta'}_j,\sigma^{\beta'}_{j+1})\) for some \(1\leq i\leq N_\beta(w)\) is bounded by \(N_\beta(w)\) from above and the number of \(j\) such that \((\sigma^{\beta'}_j,\sigma^{\beta'}_{j+1}]\subset(\sigma^\beta_{N_\beta(w)},1]\) is bounded by \(\beta/\beta'\), we have \(\sum_{i=0}^{N_\beta(w)-1} \sharp \Lambda_i \geq N_{\beta'}(w)-N_\beta(w)-\beta/\beta'\). Hence \(\beta N_\beta(w)\geq \beta'(N_{\beta'}(w)-N_\beta(w)-\beta/\beta')\). Hence we see the assertion for \(\beta'<\beta\). It can be generalized easily. We can show (2) easily from the definition. We prove (3). Let \(\{\tilde{\sigma}_i\}_{i=0}^{N_{\beta}(\tilde{w})}\) and \(\{\sigma_i\}_{i=0}^{N_1(w)}\) be corresponding increasing sequences. Then by the definition, we have \(w(\tilde{\sigma}_{i-1},\tilde{\sigma}_i)\ge 2\) for \(1\le i\le N_{\beta}(\tilde{w})\). This implies \(\sigma_i\le \tilde{\sigma}_i\)   \((1\le i\le N_{\beta}(\tilde{w}))\) and so the proof is finished. ◻

In what follows, we write \[\begin{align} \tilde{N}(B)=2^{N_{\beta}(\tilde{w})+1}. \end{align}\]

Lemma 15. Assume that Condition  (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). There exist a positive integer \(m_0\) and a positive number \(\beta\) which depend only on \(\sigma,b,c, H^-\) such that for all \(m\ge m_0\) it holds that \(\tilde{J}^{m,\rho}_t\) are invertible for all \(t\in D_m\) and \[\begin{align} \max_{t\in D_m}\big\{|\tilde{J}^{m,\rho}_t|,|(\tilde{J}^{m,\rho}_t)^{-1}|\big\}\le \tilde{N}(B). \end{align}\]

Proof. Let \(\delta\) and \(C_4\) be numbers given in Lemma 13. Let us take \(m\) satisfying \(2^{-m}\le \delta\). Let \(0<\varepsilon\le\delta\). By Lemma 13, for \(t,\tau\) satisfying \(\tilde{w}(\tau,\tau+t)\le\varepsilon\) and \(\tau+t\le 1\), we have \[\begin{align} |\tilde{J}^{m,\rho}_{t}(Y^{m,\rho}_{\tau},\theta_{\tau}B)-I|&\le C_4\varepsilon^{3H^-}+C(\varepsilon^{H^-}+\varepsilon^{2H^-}+\varepsilon), \end{align}\] where \(C\) is a constant depending only on \(\sigma, b,c\). Hence, for sufficiently small \(\varepsilon\) which depends only on \(C_4, C\), that is, depends only on \(\sigma,b,c\), it holds that for any \(t,\tau\in D_m\) with \(t+\tau\le 1\) and \(\tilde{w}(\tau,t+\tau)\le \varepsilon\), \(\tilde{J}^{m,\rho}_t(Y^{m,\rho}_{\tau},\theta_{\tau}B)\) are invertible and \[\begin{align} \max\big\{|\tilde{J}^{m,\rho}_{t}(Y^{m,\rho}_{\tau},\theta_{\tau}B)|, |\tilde{J}^{m,\rho}_{t}(Y^{m,\rho}_{\tau},\theta_{\tau}B)^{-1}|\big\} &\le 2.\label{small32tJmr} \end{align}\tag{69}\]

By the definition of \(w\), we see that there exists a constant \(C_{H^-}(\ge 1)\) such that for any \(0\le s<u<t\le 1\) \[\begin{align} w(s,t)\le C_{H^-}\left(w(s,u)+w(u,t)\right). \end{align}\] For \(\omega\in \Omega_0^{(m)}\), \(w\left(u,(u+2^{-m})\wedge 1\right)\le 2^{-m}\) holds for any \(0\le u\le 1\). Therefore, we get \[\begin{align} w\left(s,(u+2^{-m})\wedge 1\right)\le C_{H^-}\left(w(s,u)+2^{-m}\right), \quad 0\le s\le u\le 1. \end{align}\] By using this, we get \[\begin{align} \tilde{w}\left(s,(u+2^{-m})\wedge 1\right) \le C_{H^-}\left(\tilde{w}(s,u)+2^{1-m}\right), \quad 0\le s\le u\le 1. \end{align}\] Let us take a positive number \(\beta\) and \(m\) such that \[\begin{align} C_{H^-}\left(\beta+2^{1-m}\right)\le \varepsilon. \end{align}\] Note that \(\beta\) and \(m\) depends on \(C_{H^-}\) and \(\varepsilon\). Let \(\{\tilde{\sigma}_i\}_{i=0}^{N_{\beta}(\tilde{w})}\) be the increasing sequence defined by \(\tilde{w}\) and \(\beta\). Let \(\hat{\sigma}_i=\inf\{t\in D_m~|~t\ge \tilde{\sigma}_i\}\) \((0\le i\le N_{\beta}(\tilde{w}))\). Also set \(\hat{\sigma}_{N_{\beta}(\tilde{w})+1}=1\). Then we have for all \(0\le i\le N_{\beta}(\tilde{w})\) \[\begin{align} \tilde{w}(\hat{\sigma}_{i},\hat{\sigma}_{i+1}) \le \tilde{w}\left(\tilde{\sigma}_{i}, (\tilde{\sigma}_{i+1}+2^{-m})\wedge 1\right) \le C_{H^-}(\tilde{w}(\tilde{\sigma}_{i},\tilde{\sigma}_{i+1})+2^{1-m}) \le \varepsilon.\label{sigmai32and32sigmai431} \end{align}\tag{70}\] Take \(t(\ne 0)\in D_m\) and choose \(j\) so that \(\hat{\sigma}_{j-1}<t\le \hat{\sigma}_j\) \((1\le j\le N_{\beta}(\tilde{w})+1)\). We have \[\begin{align} \tilde{J}^{m,\rho}_t(\xi,B)= \tilde{J}^{m,\rho}_{t-\hat{\sigma}_{j-1}}(Y^{m,\rho}_{\hat{\sigma}_{j-1}}, \theta_{\hat{\sigma}_{i-1}}B)\cdots \tilde{J}^{m,\rho}_{\hat{\sigma}_2-\hat{\sigma}_1}(Y^{m,\rho}_{\hat{\sigma}_1}, \theta_{\hat{\sigma}_1}B) \tilde{J}^{m,\rho}_{\hat{\sigma}_1}(\xi,B). \label{product32of32tJmr1} \end{align}\tag{71}\] By (69 ), (70 ) and (71 ), We obtain \[\begin{align} \max_{t\in D_m}\big\{ |\tilde{J}^{m,\rho}_t(\xi,B)|, |\tilde{J}^{m,\rho}_t(\xi,B)^{-1}| \big\}\le 2^{N_{\beta}(\tilde{w})+1}, \end{align}\] which completes the proof. ◻

Lemma 16. Assume that Condition  (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). Set \(\lambda=\min\{\lambda_1,2H^-\}\). Let \(m\) be a sufficiently large number as in Lemma . There exists a positive number \(C_5\) which does not depend on \(m\) and depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially such that, for all \(t,s\in D_m\), \[\begin{gather} \big|\tilde{J}^{m,\rho}_{t}-\tilde{J}^{m,\rho}_{s} -(D\sigma)(Y^{m,\rho}_{s}) [\tilde{J}^{m,\rho}_{s}]B_{s,t} -D\left((D\sigma)[\sigma]\right)(Y^{m,\rho}_{s})[\tilde{J}^{m,\rho}_{s}]{\mathbb{B}}_{s,t}\\ -\rho (Dc)(Y^{m,\rho}_{s})[\tilde{J}^{m,\rho}_{s}]d^m_{s,t} -(Db)(Y^{m,\rho}_{s})[\tilde{J}^{m,\rho}_{s}](t-s)\big|\le C_5|t-s|^{\lambda+H^-}.\label{estimate32of32remainder32term32of32tJmr} \end{gather}\tag{72}\]

Proof. We already proved that there exists \(\tilde{N}(B)\) such that \(|\tilde{J}^{m,\rho}_t|\le \tilde{N}(B)\) for all sufficiently large \(m\) and \(t\in D_m\). Noting this boundedness, we obtain desired result by the same proofs as in Lemmas 8 and 9. ◻

\((\tilde{J}^{m,\rho}_t)^{-1}\) also satisfies a similar estimate.

Lemma 17. For every \(s,t\in D_m\) with \(s\le t\), set \[\begin{align} \tilde{A}^{m,\rho}_{s,t} &= - \Big[ (D\sigma)(Y^{m,\rho}_{s})B_{s,t}\\ &\qquad\qquad + \sum_{\alpha,\beta} \Bigl\{ (D\sigma)(Y^{m,\rho}_{s}) [(D\sigma)(Y^{m,\rho}_{s})e_{\beta}]e_{\alpha}- (D^2\sigma)(Y^{m,\rho}_{s}) [\cdot,\sigma(Y^{m,\rho}_{s})e_{\alpha}]e_{\beta} \Bigr\} B^{\alpha,\beta}_{s,t}\\ &\qquad\qquad + \rho (Dc)(Y^{m,\rho}_{s})d^m_{s,t} + (Db)(Y^{m,\rho}_{s})(t-s) \Big]. \end{align}\] Assume that Condition  (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). Set \(\lambda=\min\{\lambda_1,2H^-\}\). Let \(m\) be a sufficiently large number as in Lemma .

  1. We define \(\tilde{\epsilon}^{m,\rho}_{\tau^m_{i-1},\tau^m_i}\) by \(\tilde{\epsilon}^{m,\rho}_{\tau^m_{i-1},\tau^m_i} = (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} -(\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} -(\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1}\tilde{A}^{m,\rho}_{\tau^m_{i-1},\tau^m_i}\). Then it holds that \[\begin{align} & |\tilde{\epsilon}^{m,\rho}_{\tau^m_{i-1},\tau^m_i}|\le 2\tilde{N}(B) \Bigl(1+\|D\sigma\|+\|D\left((D\sigma)[\sigma]\right)\|+ \|Dc\|+\|Db\|\Bigr)^3\Delta_m^{\lambda+H^-}. \end{align}\]

  2. For all \(s,t\in D_m\) with \(s\le t\), it holds that there exists a constant \(C_6\) which is defined by a polynomial function of \(\tilde{C}(B)\) and \(\tilde{N}(B)\) such that \[\begin{align} \big|(\tilde{J}^{m,\rho}_{t})^{-1}-(\tilde{J}^{m,\rho}_{s})^{-1} -(\tilde{J}^{m,\rho}_{s})^{-1}\tilde{A}^{m,\rho}_{s,t}\big|\le C_6|t-s|^{\lambda+H^-}. \label{estimate32of32remainder32term32of32tJmr94-1} \end{align}\tag{73}\]

Proof. (1) Set \(A^{m,\rho}_{\tau^m_{i-1},\tau^m_i} = I-E^{m,\rho}(Y^{m,\rho}_{\tau^m_{i-1}},\theta_{\tau^m_{i-1}}B)\). By the equation 54 , we have \[\begin{gather} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1}-(\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} = (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} \left(E^{m,\rho}(Y^{m,\rho}_{\tau^m_{i-1}},\theta_{\tau^m_{i-1}}B)^{-1}-I\right)\\ = (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} \left((I-A^{m,\rho}_{\tau^m_{i-1},\tau^m_i})^{-1}-I\right) = (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} \bigg[ A^{m,\rho}_{\tau^m_{i-1},\tau^m_i} + \sum_{l=2}^{\infty} \left\{ A^{m,\rho}_{\tau^m_{i-1},\tau^m_i} \right\}^l \bigg]. \end{gather}\] By the geometric property \(B^{\alpha,\beta}_{s,t}=B^{\alpha}_{s,t}B^{\beta}_{s,t}-B^{\beta,\alpha}_{s,t}\), we have \[\begin{align} & (D\sigma)(Y^{m,\rho}_s)[(D\sigma)(Y^{m,\rho}_s)B_{s,t}] B_{s,t}- (D\sigma)(Y^{m,\rho}_s)[(D\sigma)(Y^{m,\rho}_s)]{\mathbb{B}}_{s,t} \nonumber\\ &\quad= (D\sigma)(Y^{m,\rho}_s)[(D\sigma)(Y^{m,\rho}_s)e_{\alpha}] e_{\beta}B^{\alpha}_{s,t}B^{\beta}_{s,t}- (D\sigma)(Y^{m,\rho}_s)[(D\sigma)(Y^{m,\rho}_s)e_{\alpha}]e_{\beta} {\mathbb{B}}^{\alpha,\beta}_{s,t}\nonumber\\ &\quad=(D\sigma)(Y^{m,\rho}_s)[(D\sigma)(Y^{m,\rho}_s)e_{\alpha}]e_{\beta} {\mathbb{B}}^{\beta,\alpha}_{s,t}. \end{align}\] Using this and by the assumption of (55 ) and Lemma 15, we obtain the desired estimate.

(2) We have proved that \((\tilde{J}^{m,\rho}_t)^{-1}\) satisfies a similar equation to \(Y^{m,\rho}_t\) and the norm can be estimated as in Lemma 15. Hence, we can complete the proof in the same way as in Lemma 9. ◻

We now give an estimate of discrete rough integral similarly to Lemma 10.

Lemma 18. Let \(\varphi\) be a \(C^2_b\) function on \({\mathbb{R}}^n\times \mathcal{L}({\mathbb{R}}^n)\times \mathcal{L}({\mathbb{R}}^n)\) with values in \(\mathcal{L}({\mathbb{R}}^d,{\mathbb{R}}^l)\) whose all derivatives and itself are at most polynomial order growth. For \(t\in D_m\), set \[\begin{gather} I^{m,\rho}(\varphi)_t\\ = \sum_{i=1}^{2^mt} \Big\{ \varphi \left( Y^{m,\rho}_{\tau^m_{i-1}}, \tilde{J}^{m,\rho}_{\tau^m_{i-1}}, (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} \right) B_{\tau^m_{i-1},\tau^m_i} + \varphi \left( Y^{m,\rho}, \tilde{J}^{m,\rho}, (\tilde{J}^{m,\rho})^{-1} \right)^{\boldsymbol{\cdot}}_{\tau^m_{i-1}} {\mathbb{B}}_{\tau^m_{i-1},\tau^m_i} \Big\}, \end{gather}\] where \(\varphi ( Y^{m,\rho}, \tilde{J}^{m,\rho}, (\tilde{J}^{m,\rho})^{-1} )^{\boldsymbol{\cdot}}_t\) \((t\in D_m)\) is the \(\mathcal{L}({\mathbb{R}}^d\otimes{\mathbb{R}}^d,{\mathbb{R}}^l)\)-valued process such that \[\begin{align} \varphi \left( Y^{m,\rho}, \tilde{J}^{m,\rho}, (\tilde{J}^{m,\rho})^{-1} \right)^{\boldsymbol{\cdot}}_t [v \otimes w] = (D_1\varphi) \left( Y^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1} \right) \left[\sigma\left(Y^{m,\rho}_t\right)v\right]w \\& + (D_2\varphi) \left( Y^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1} \right) \left[ (D\sigma)\left(Y^{m,\rho}_t\right) \left[\tilde{J}^{m,\rho}_t\cdot\right]v \right] w \\ & - (D_3\varphi) \left( Y^{m,\rho}_t, \tilde{J}^{m,\rho}_t, (\tilde{J}^{m,\rho}_t)^{-1} \right) \left[ (\tilde{J}^{m,\rho}_t)^{-1} (D\sigma)\left(Y^{m,\rho}_t\right)[\cdot]v \right]w\end{align}\] for \(v,w\in{\mathbb{R}}^d\). Here \(D_i\) denotes the derivative with respect to the \(i\)-th variable of \(\varphi\).

Assume that Condition 9 (1) holds and let \(\omega\in \Omega_0^{(m,d^m)}\). We have \(\|I^{m,\rho}\|_{H^-}\le C_7,\) where \(C_7\) depends on \(\sigma, b, c, \varphi, C(B), \tilde{N}(B)\) polynomially.

Proof. We already proved Lemma 16 and Lemma 17. Hence the proof is similar to that of Lemma 10. ◻

So far, we have given deterministic estimates of our processes based on \(\tilde{C}(B)\) and \(\tilde{N}(B)\). We now give \(L^p\) estimate of our processes. The following result is due to [13]. See [14] also.

Lemma 19. Assume that the covariance \(R\) satisfies Condition . Let \(w\) be the control function defined in Definition . Then for any \(\beta>0\), there exist positive numbers \(c_1\) and \(c_2\) depending only on \(H\) and \(\beta\) such that \[\begin{align} \mu\left(N_{\beta}(w)\ge r\right)\le c_1 e^{-c_2 r^{4H}}. \label{exp32decay} \end{align}\tag{74}\]

The following is an immediate consequence of Lemma 14 and Lemma 19. Note that \(N_{\beta}(\tilde{w})\) is a random variable defined on \(\Omega_0\).

Corollary 2. Assume the same assumption in Lemma . A similar estimate to \((\ref{exp32decay})\) holds for \(N_{\beta}(\tilde{w})\).

By these results, under additional assumption on the covariance of \((B_t)\), we obtain \(L^p\) estimate of several quantities.

Lemma 20. Assume that Condition  (1) holds. Let \(\tilde{N}(B), C_5, C_6\) and \(C_7\) be the positive numbers defined in Lemmas , , and . Then we have \[\begin{align} \max\big\{\tilde{N}(B),C_5, C_6, C_7\big\}\in \cap_{p\ge 1}L^p(\Omega_0). \end{align}\] In particular \[\begin{align} \sup_m \Big\| \max_{0\leq \rho\leq 1, t\in D_m} \big\{ |\tilde{J}^{m,\rho}_t(\xi,B)|, |\tilde{J}^{m,\rho}_{t}(\xi,B)^{-1}| \big\}1_{\Omega_0^{(m,d^m)}} \Big\|_{L^p} <\infty. \end{align}\]

Consequently we obtain the following estimate. Note that \(\tilde{Z}^{m,\rho}_t\) is a discrete process defined by 58 . Also recall that the notion of \(\{a_m\}\)-order nice discrete process was introduced and the definition of \(\sup_{t,\rho}|Y^{m,\rho}_t-Y_t|=O(a_m)\) was given in Definition 3.

Theorem 28. Assume that Conditions  and  (1) hold. Let \(\varepsilon_1\) be the constant given in Condition 8. Set \(a_m=\max\{\Delta_m^{3H^--1},\Delta_m^{\varepsilon_1}\}\). Then we have the following.

  1. It holds that \(\{\tilde{Z}^{m,\rho}\}_{m}\) is an \(\{a_m\}\)-order nice discrete process with the Hölder exponent \(\lambda=\min\{\lambda_1,2H^-\}\) which is independent of \(\rho\).

  2. It holds that \(\sup_{t,\rho}|Y^{m,\rho}_t-Y_t|=O(a_m)\) in the sense of Definition  .

  3. For any \(p\ge 1\) and \(\kappa>0\), we have \[\begin{align} \lim_{m\to\infty}\|(2^m)^{\min\{3H^--1,\varepsilon_1\}-\kappa} \max_{t\in D_m}|\hat{Y}^{m}_t-Y_t|\|_{L^p}=0. \end{align}\]

Proof. (1) Note that the processes \((\tilde{J}^{m,\rho})^{-1}\) and \(c(Y^{m,\rho})\) appeared in 58 admit the uniform Hölder estimates and that \(d^m\) and \(\hat{\epsilon}^{m}-\epsilon^m\) are \(\{a_m\}\)-order nice discrete processes (see Remark 21). Hence the assertion follows from Remark 22. (2) follows from (1) and Proposition 26. We prove (3). By (2), there exists \(X\in \cap_{p\ge 1}L^p(\Omega)\) such that \(\max_t|\hat{Y}^{m}_t-Y_t|\le a_mX\) on \(\Omega_0^{(m,d^m)}\). Also we have for any \(R>0\), there exists \(C_R>0\) such that \(\mu\Big((\Omega_0^{(m,d^m)})^\complement\Big)\le C_R2^{-mR}\). Using these estimates and the Schwarz inequality, we have \[\begin{align} &\|(2^m)^{\min\{3H^--1,\varepsilon_1\}-\kappa}\max_t|\hat{Y}^{m}_t-Y_t|\|_{L^p}^p\\ &\le E\left[(2^m)^{-\kappa p} X^p ; \Omega_0^{(m,d^m)}\right]+ E\left[(2^m)^{(\min\{3H^--1,\varepsilon_1\}-\kappa)p} \max_t|\hat{Y}^{m}_t-Y_t|^p ;(\Omega_0^{(m,d^m)})^\complement\right]\\ &\le 2^{-mp\kappa}\|X\|_{L^p}^p+ (2^m)^{(\min\{3H^--1,\varepsilon_1\}-\kappa)p-R/2}C_R^{\frac{1}{2}}E[\max_t|\hat{Y}^{m}_t-Y_t|^{2p}]^{\frac{1}{2}}. \end{align}\] Combining this estimate and Lemma 9, we complete the proof. ◻

We remark some consequences of the above results in the case of the Milstein approximate solution.

Remark 29.

  1. Let us consider non-random case. That is, we consider a \(\theta\)-Hölder geometric rough path \((X,{\mathbb{X}})\). The Milstein approximation solution \(\hat{Y}^{m}_t\) (\(t\in D_m\)) is defined by the similar equation to that explained in Section 2.2 replacing \((B,{\mathbb{B}})\) by \((X,{\mathbb{X}})\). Let \(C(X)=\max\{\|X\|_{\theta},\|{\mathbb{X}}\|_{2\theta}\}\). Also we define \(\tilde{N}(X)\) similarly to \(\tilde{N}(B)\). Note that \(d^m\equiv 0\) and \(\hat{\epsilon}^{m}\equiv 0\) and we have the estimate \(|\epsilon^m_{\tau^m_{k-1},\tau^m_k}|\le C\Delta_m^{3\theta}\), where \(C\) depends on \(\sigma, b, C(X)\) polynomially. Let \(\kappa\) be a small positive number and set \(\theta^-=\theta-\kappa\). We can view \((X,{\mathbb{X}})\) as a \(\theta^-\)-Hölder rough path. Then for sufficiently large \(m\), we have \[\begin{align} \sup_{|t-s|\le 2^{-m}} \left|\frac{X_{s,t}}{(t-s)^{\theta^{-}}}\right| + \sup_{|t-s|\le 2^{-m}} \left|\frac{{\mathbb{X}}_{s,t}}{(t-s)^{2\theta^{-}}}\right| \le 2^{-m\kappa+1}C(X)\le \frac{1}{2}. \end{align}\] We can define an interpolated process \(Y^{m,\rho}_t\) and \(\tilde{J}^{m,\rho}_t\) similarly. By the same argument as in this section, we obtain, \[\begin{align} \max_{t\in D_m}|\hat{Y}^{m}_{t}-Y_t|\le C\Delta_m^{3\theta^--1}, \label{Milstein32estimate} \end{align}\tag{75}\] where \(C\) depends on \(\sigma, b\) and \(C(X), \tilde{N}(X)\) polynomially. Similar estimate was obtained by Davie [10] and Friz-Victoir [11]. As for implementable versions, one can find some information in [7]. We think our estimate makes clear how \(C\) depends on \((X,{\mathbb{X}})\) more explicitly in (75 ). In Theorem 28, we deal with an RDE driven by random rough path \((B,{\mathbb{B}})\) for which \(\tilde{N}(B), C(B)\in \cap_{p\ge 1}L^p(\Omega_0)\) holds. Hence, we can obtain \(L^p\) convergence in (3).

  2. We consider RDEs driven by \(B\) which satisfies Condition 3. We can prove \(\sup_t\{|J_t|+|J^{-1}_t|\}\in \cap_{p\ge 1}L^p(\Omega)\) by applying the above results in the case where \(\rho=1\) to the Milstein approximation solution \((\hat{Y}^{m}_t, \tilde{J}^{m,1}_t)\) \((t\in D_m)\). Note that \(\Omega_0^{(m,d^m)}=\Omega_0^{(m)}\) and \(\liminf_{m\to\infty}\Omega_0^{(m)}=\Omega_0\) hold. By Theorem 28, we see that \(\lim_{m\to\infty}\max_{t\in D_m}|\hat{Y}^{m}_t-Y_t|=0\) for all \(\omega\in \Omega_0\). Let \(\hat{J}^{m,1}_t\) and \((\hat{J}^{m,1}_t)^{-1}\) \((t\in [0,1])\) be piecewise linear extensions of \(\tilde{J}^{m,1}_t\) and \((\tilde{J}^{m,1}_t)^{-1}\) \((t\in D_m)\) respectively. Since \(\tilde{J}^{m,1}_t\) and \((\tilde{J}^{m,1}_t)^{-1}\) are uniform Hölder continuous paths on \(D_m\) which follow from Lemmas 9, 16, 17, so are \(\hat{J}^{m,1}_t\) and \((\hat{J}^{m,1}_t)^{-1}\) on \([0,1]\). This implies that for any subsequences of \(\hat{J}^{m,1}_t\) and \((\hat{J}^{m,1}_t)^{-1}\), there exist subsequences of them which converge uniformly on \([0,1]\). By the estimate in Lemmas 16, 17 and the uniqueness of RDEs, any limits of \(\hat{J}^{m,1}_t\) and \((\hat{J}^{m,1}_t)^{-1}\) are equal to \(J_t\) and \(J^{-1}_t\) respectively. This implies that the limits of themselves without taking subsequences exist and the limits \(J_t\) and \(J_t^{-1}\) also satisfy the same estimates as in (72 ) and (73 ) for all \(\omega\in\Omega_0\).

  3. We can improve the estimate in Theorem 28 (3) when the driving process is an fBm as you can see in Theorem 15 and Remark 16.

4.3 Estimates of \(J_t-\tilde{J}^m_t\) and \(J_t^{-1}-(\tilde{J}^m_t)^{-1}\) on \(\Omega_0^{(m)}\)↩︎

Throughout this section, \(Y_t\) and \(J_{t}\) denote the solutions to 10 and 11 , respectively. Recall \(\tilde{J}^m_t=\tilde{J}^{m,0}_t\) is defined by 49 . Note that the recurrence relation for \(\tilde{J}^m\) does not contain the terms \(d^m\) and \(\hat{\epsilon}^{m}\). Hence we do not need assumptions on \(d^m\) and \(\hat{\epsilon}^{m}\) in this section. Again, we assume \(m\) satisfies (55 ). From now on, we will give estimates of \(J_t-\tilde{J}^m_t\) and \(J_t^{-1}-(\tilde{J}^m_t)^{-1}\). We define \(\epsilon(J)_{\tau^m_{k-1},\tau^m_k}\) by \[\begin{align} J_{\tau^m_k} &= J_{\tau^m_{k-1}} +(D\sigma)(Y_{\tau^m_{k-1}})[J_{\tau^m_{k-1}}]B_{\tau^m_{k-1},\tau^m_k} +(D^2\sigma)(Y_{\tau^m_{k-1}}) \left[J_{\tau^m_{k-1}},\sigma(Y_{\tau^m_{k-1}})e_{\alpha}\right]e_{\beta} {\mathbb{B}}^{\alpha,\beta}_{\tau^m_{k-1},\tau^m_k} \nonumber\\ &\quad\qquad + (D\sigma)(Y_{\tau^m_{k-1}})\left[(D\sigma)(Y_{\tau^m_{k-1}})[J_{\tau^m_{k-1}}]e_{\alpha}\right]e_{\beta} {\mathbb{B}}^{\alpha,\beta}_{\tau^m_{k-1},\tau^m_k} +(Db)(Y_{\tau^m_{k-1}})[J_{\tau^m_{k-1}}]\Delta_m \nonumber\\ &\quad\qquad +\epsilon(J)_{\tau^m_{k-1},\tau^m_k}.\label{eq32of32J} \end{align}\tag{76}\]

Lemma 21. Let \(\omega\in \Omega_0^{(m)}\). Let \[\begin{align} \delta^{m}(J)_t=-\sum_{i=1}^{2^mt} \big(\tilde{J}^m_{\tau^m_i}\big)^{-1} \epsilon(J)_{\tau^m_{i-1},\tau^m_i}, \qquad t\in D_m. \end{align}\]

  1. It holds that \[|\epsilon(J)_{\tau^m_{k-1},\tau^m_k}|\le C_5\Delta_m^{3H^-},\quad 1\le k\le 2^m,\] where \(C_5\) is the constant in Lemma .

  2. \(\{\delta^m(J)_t\}_{t\in D_m}\) is a \(\{\Delta_m^{3H^--1}\}\)-order nice discrete process with the Hölder exponent \(2H^-\) and \[\begin{align} \max_{t\in D_m}|\delta^m(J)_t|=O(\Delta_m^{3H^--1}). \end{align}\]

  3. For any natural number \(R\), it holds that \[\begin{align} \tilde{J}^m_t &= J_t \Bigg( I + \sum_{r=1}^R (\delta^{m}(J)_t)^r \Bigg) + (\tilde{J}^m_t-J_t)\delta^{m}(J)_t^R.\label{tJ-J32expansion} \end{align}\tag{77}\] In particular, \[\begin{align} \max_{t\in D_m}\left|\tilde{J}^m_t-J_t\left(I+\sum_{r=1}^R (\delta^{m}(J)_t)^r\right)\right|=O(\Delta_m^{(3H^--1)(R+1)}). \label{estimate32of32tJ-J} \end{align}\tag{78}\]

  4. For any natural numbers \(L\) and \(R\), it holds that \[\begin{gather} \max_{t\in D_m} \Bigg| (\tilde{J}^m_t)^{-1} - \Bigg\{ I + \sum_{l=1}^L \left( - \sum_{r=1}^R (\delta^{m}(J)_t)^r \right)^l \Bigg\} J_t^{-1} \Bigg|\\ = O(\Delta_m^{(3H^--1)(L+1)})+O(\Delta_m^{(3H^--1)(R+1)}). \end{gather}\]

Proof. (1) This follows from Lemma 16 and Remark 29.

(2) Similarly to \(\epsilon^m_t\) and \(\hat{\epsilon}^{m}_t\) (see 2 ), we set \(\epsilon(J)^m_t=\sum_{i=1}^{2^mt} \epsilon(J)_{\tau^m_{i-1},\tau^m_i}\) \((t\in D_m)\). From assertion (1), \(\epsilon(J)^m\) is a \(\{\Delta_m^{3H^--1}\}\)-order nice discrete process. Hence, using the estimate of \(\tilde{J}^m\) and Remark 22, we see assertion (2).

(3) From the definition of \(\tilde{J}^m\) and 76 , we have \[\begin{align} J_t &= \tilde{J}^m_t+\tilde{J}^m_t\sum_{i=1}^{2^mt} \left(\tilde{J}^m_{\tau^m_i}\right)^{-1} \epsilon(J)_{\tau^m_{i-1},\tau^m_i} = \tilde{J}^m_t-\tilde{J}^m_t\delta^{m}(J)_t \end{align}\] Hence \(\tilde{J}^m_t-J_t = J_t \delta^{m}(J)_t + (\tilde{J}^m_t-J_t) \delta^{m}(J)_t\), which implies 77 . Noting \(\tilde{J}^m_t-J_t=\tilde{J}^m_t\delta^m(J)_t\), we get (78 ).

(4) Note that \[\begin{align} J_{t}^{-1}-(\tilde{J}^m_{t})^{-1} &= - (\tilde{J}^m_{t})^{-1} \big(J_{t}-\tilde{J}^m_{t}\big) J_{t}^{-1} \\ &= - J_t^{-1} \big(J_t-\tilde{J}^m_t\big) J_t^{-1} + \big(J_{t}^{-1}-(\tilde{J}^m_{t})^{-1}\big) \big(J_t-\tilde{J}^m_t\big) J_t^{-1}. \end{align}\] Iterating this \(L\) times and using the first identity above, we get \[\begin{align} J_t^{-1}-(\tilde{J}^m_t)^{-1} &= - J_t^{-1} \sum_{l=1}^L \big[ \big(J_t-\tilde{J}^m_t\big) J_t^{-1} \big]^l + (J_t^{-1}-(\tilde{J}^m_t)^{-1}) \big[ \big(J_t-\tilde{J}^m_t\big) J_t^{-1} \big]^L\\ &= - J_t^{-1} \sum_{l=1}^L \big[ \big(J_t-\tilde{J}^m_t\big) J_t^{-1} \big]^l - (\tilde{J}^m_{t})^{-1} \big[ \big(J_t-\tilde{J}^m_t\big) J_t^{-1} \big]^{L+1}. \end{align}\] and \((\tilde{J}^m_{t})^{-1} \big[ \big(J_t-\tilde{J}^m_t\big) J_t^{-1} \big]^{L+1} = O(\Delta_m^{(3H^--1)(L+1)})\). Thus \[\begin{align} J_t^{-1}-(\tilde{J}^m_t)^{-1}\\&= - J_t^{-1} \sum_{l=1}^L \left[ \left( - J_t \sum_{r=1}^R (\delta^{m}(J)_t)^r + O(\Delta_m^{(3H^--1)(R+1)}) \right) J_t^{-1} \right]^l + O(\Delta_m^{(3H^--1)(L+1)}) \\ &= - J_t^{-1} \sum_{l=1}^L \left[ \left(- J_t\sum_{r=1}^R (\delta^{m}(J)_t)^r\right) J_t^{-1} \right]^l +O(\Delta_m^{(3H^--1)(L+1)}) +LO(\Delta_m^{(3H^--1)(R+1)}).\end{align}\] Since we have \[\begin{align} \left[ \left(- J_t \sum_{r=1}^R (\delta^{m}(J)_t)^r\right) J_t^{-1} \right]^l = J_t \left( - \sum_{r=1}^R (\delta^{m}(J)_t)^r \right)^l J_t^{-1}, \end{align}\] we arrive at the conclusion. ◻

Remark 30. Summarizing above, we have the following. By taking \(L=R\) as a positive integer, we have \[\begin{align} \tilde{J}^m_t-J_t&=J_tK^{1,m,R}_t+L^{1,m,R}_t, & (\tilde{J}^m_t)^{-1}-J_t^{-1}&=K^{2,m,R}_tJ_t^{-1}+L^{2,m,R}_t, \end{align}\] where \(K^{1,m,R}\) and \(K^{2,m,R}\) are \(\{\Delta_m^{3H^--1}\}\)-order nice discrete processes with the Hölder exponent \(2H^-\) and \(\max_t\{|L^{1,m,R}_t|+|L^{2,m,R}_t|\}=O(\Delta_m^{(3H^--1)R})\).

4.4 Convergence of \(\tilde{J}^{m,\rho}_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}\)↩︎

Here we show convergence of \(\tilde{J}^{m,\rho}_t\) and \((\tilde{J}^{m,\rho}_t)^{-1}\). To this end we study \(N^{m,\rho}_t=(-\tilde{J}^{m,\rho}_t)^{-1}\partial_{\rho}\tilde{J}^{m,\rho}_t\). Note that \(N^{m,\rho}_t\) is defined on \(\Omega_0^{(m,d^m)}\) and for large \(m\) because \((-\tilde{J}^{m,\rho}_t)^{-1}\) can exist under the same condition.

Lemma 22. Assume that Conditions  and  (1) hold. Let \(\varepsilon_1\) be the constant given in Condition 8. Set \(a_m=\max\{\Delta_m^{3H^--1},\Delta_m^{\varepsilon_1}\}\). Let \(f_1,\dots,f_n\) be the standard basis of \({\mathbb{R}}^n\) and write \(\tilde{Z}^{m,\rho,\nu}_t=(\tilde{Z}^{m,\rho}_t,f_\nu)\) for \(\nu=1,\dots,n\). Note that \(\tilde{Z}^{m,\rho,\nu}\) is a real-valued process.

  1. Let \(\omega\in \Omega_0^{(m,d^m)}\). We have \[\begin{align} N^{m,\rho}_t &= \sum_{\nu=1}^n \tilde{Z}^{m,\rho,\nu}_t I^{m,\rho}(\varphi_\nu)_t + \sum_{\lambda=0}^3 I_\lambda(N^{m,\rho})_t \end{align}\] Here, \(\varphi_\nu(x,M_1,M_2)\) \(\left(x\in {\mathbb{R}}^n, M_1, M_2\in \mathcal{L}({\mathbb{R}}^n)\right)\) is an \(\mathcal{L}\left({\mathbb{R}}^d, \mathcal{L}({\mathbb{R}}^n)\right)\)-valued function defined by \[\begin{align} \varphi_\nu(x,M_1,M_2) = -M_2(D^2\sigma)(x)[M_1f_\nu,M_1] \end{align}\] and \(I^{m,\rho}(\varphi_\nu)\) is a discrete rough integral defined in Lemma . Explicitly, we have, for \(t\in D_m\), \[\begin{align} \varphi_\nu \left( Y^{m,\rho},\tilde{J}^{m,\rho},(\tilde{J}^{m,\rho})^{-1} \right)^{\boldsymbol{\cdot}}_t[v\otimes w]\\&= - (\tilde{J}^{m,\rho}_t)^{-1} (D\sigma)(Y^{m,\rho}_t) \left[ (D^2\sigma)(Y^{m,\rho}_t)\left[\tilde{J}^{m,\rho}_tf_\nu,\tilde{J}^{m,\rho}_t\right]w \right] v\\ &\phantom{=}\qquad + (\tilde{J}^{m,\rho}_t)^{-1} (D^3\sigma)(Y^{m,\rho}_t) \left[ \sigma(Y^{m,\rho}_t)v, \tilde{J}^{m,\rho}_tf_\nu, \tilde{J}^{m,\rho}_t \right] w\\ &\phantom{=}\qquad + (\tilde{J}^{m,\rho}_t)^{-1} (D^2\sigma)(Y^{m,\rho}_t) \left[ (D\sigma)(Y^{m,\rho}_t)\left[\tilde{J}^{m,\rho}_tf_\nu\right]v,\tilde{J}^{m,\rho}_t \right] w\\ &\phantom{=}\qquad + (\tilde{J}^{m,\rho}_t)^{-1} (D^2\sigma)(Y^{m,\rho}_t) \left[ \tilde{J}^{m,\rho}_tf_\nu, (D\sigma)(Y^{m,\rho}_t)\left[\tilde{J}^{m,\rho}_t\right]v \right] w, \qquad v,w\in{\mathbb{R}}^d.\end{align}\] Also \[\begin{align} I_0(N^{m,\rho})_t &= - \sum_{\nu=1}^n \sum_{j=1}^{2^m t} \tilde{Z}^{m,\rho,\nu}_{\tau^m_{j-1},\tau^m_j} I^{m,\rho}(\varphi_\nu)_{\tau^m_j},\\ I_1(N^{m,\rho})_t &= \sum_{j=1}^{2^m t} (-\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} (D^2b)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}] \Delta_m,\\ I_2(N^{m,\rho})_t &= \sum_{j=1}^{2^m t} (-\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} \Big\{(Dc)(Y^{m,\rho}_{\tau^m_{j-1}})[\tilde{J}^{m,\rho}_{\tau^m_{j-1}}]\\ &\qquad\qquad\qquad\qquad\qquad + \rho(D^2c)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}] \Big\} d^m_{\tau^m_{j-1},\tau^m_j}, \end{align}\] and \(I_3(N^{m,\rho})\) is the residual term defined by \[\begin{align} I_3(N^{m,\rho})_t=N^{m,\rho}_t- \sum_{\nu=1}^n \tilde{Z}^{m,\rho,\nu}_t I^{m,\rho}(\varphi_\nu)_t -\sum_{\lambda=0}^2I_{\lambda}(N^{m,\rho})_t. \end{align}\]

  2. \(I_0(N^{m,\rho})\), \(I_1(N^{m,\rho})\), \(I_2(N^{m,\rho})\) and \(I_3(N^{m,\rho})\) are \(\{a_m\}\)-order nice discrete processes with the Hölder exponent \(\lambda=\min\{\lambda_1,2H^-\}\). In addition, \(\sup_{\rho}\|N^{m,\rho}\|_{H^-}=O(a_m)\) in the sense of Definition  .

Proof. From 54 , we have \[\begin{align} N^{m,\rho}_{\tau^m_j} = N^{m,\rho}_{\tau^m_{j-1}} + (-\tilde{J}^{m,\rho}_{\tau^m_j})^{-1} \left\{\partial_{\rho}E^{m,\rho}(Y^{m,\rho}_{\tau^m_{j-1}},\theta^m_{\tau^m_{j-1}}B)\right\} \tilde{J}^{m,\rho}_{\tau^m_{j-1}}. \end{align}\] Using \((\tilde{J}^{m,\rho}_{\tau^m_j})^{-1} = (\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} \{ I -(D\sigma)(Y^{m,\rho}_{\tau^m_{j-1}})B_{\tau^m_{j-1},\tau^m_j} +O(\Delta_m^{2H^-}) \}\) due to Lemma 6 and the expression of \(\partial_\rho E^{m,\rho}(Y^{m,\rho}_{\tau^m_{j-1}},\theta^m_{\tau^m_{j-1}}B)\), we have \[\begin{align} N^{m,\rho}_{\tau^m_j}-N^{m,\rho}_{\tau^m_{j-1}} &= (-\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} \Big\{ (D^2\sigma)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}]B_{\tau^m_{j-1},\tau^m_j} \nonumber\\ &\qquad\qquad\qquad\qquad - (D\sigma)(Y^{m,\rho}_{\tau^m_{j-1}}) \left[ (D^2\sigma)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}]B_{\tau^m_{j-1},\tau^m_j} \right] B_{\tau^m_{j-1},\tau^m_j} \nonumber \\ &\qquad\qquad\qquad\qquad + D^2((D\sigma)[\sigma])(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}]{\mathbb{B}}_{\tau^m_{j-1},\tau^m_j} \Big\} \nonumber \\ &\phantom{=}\quad + (-\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} \Big[ (Dc)(Y^{m,\rho}_{\tau^m_{j-1}})[\tilde{J}^{m,\rho}_{\tau^m_{j-1}}] + \rho(D^2c)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}] \Big]d^m_{\tau^m_{j-1},\tau^m_j} \nonumber \\ &\phantom{=}\quad + (-\tilde{J}^{m,\rho}_{\tau^m_{j-1}})^{-1} \Big[ (D^2b)(Y^{m,\rho}_{\tau^m_{j-1}})[Z^{m,\rho}_{\tau^m_{j-1}},\tilde{J}^{m,\rho}_{\tau^m_{j-1}}]\Delta_m \Big] + O(\Delta_m^{3H^-}). \label{eq4349023211} \end{align}\tag{79}\] Next we take the sum over \(0\leq j\leq 2^mt\). Applying \(B^\alpha_{s,t}B^\beta_{s,t}-B^{\alpha,\beta}_{s,t}=B^{\beta,\alpha}_{s,t}\) and substituting \(Z^{m,\rho}_{\tau^m_{j-1}}=\tilde{J}^{m,\rho}_{\tau^m_{j-1}}\tilde{Z}^{m,\rho}_{\tau^m_{j-1}}=\sum_{\nu=1}^n\tilde{Z}^{m,\rho,\nu}_{\tau^m_{j-1}} \tilde{J}^{m,\rho}_{\tau^m_{j-1}}f_\nu\), we see that the summation of the first term in 79 gives \[\begin{align} \sum_{\nu=1}^n \sum_{j=1}^{2^mt} \tilde{Z}^{m,\rho,\nu}_{\tau^m_{j-1}} I^{m,\rho}(\varphi_\nu)_{\tau^m_j,\tau^m_{j-1}} = \sum_{\nu=1}^n \tilde{Z}^{m,\rho,\nu}_t I^{m,\rho}(\varphi_\nu)_t - \sum_{\nu=1}^n \sum_{j=1}^{2^mt} \tilde{Z}^{m,\rho,\nu}_{\tau^m_j,\tau^m_{j-1}} I^{m,\rho}(\varphi_\nu)_{\tau^m_j}. \end{align}\] The summations of the second and third terms in 79 give \(I_2(N^{m,\rho})\) and \(I_1(N^{m,\rho})\), respectively. The summation of the fourth term \(O(\Delta_m^{3H^-})\) in 79 is \(I_3(N^{m,\rho})\), which is an \(\{a_m\}\)-order nice discrete process. This completes the proof of (1).

We show assertion (2). Recall that the discrete Hölder norm \(\|I^{m,\rho}(\varphi_\nu)\|_{H^-}\) can be estimated by a constant which depends on \(\sigma, b, c\), \(C(B)\) and \(\tilde{N}(B)\) polynomially (see Lemma ) and that \(\tilde{Z}^{m,\rho,\nu}\) is an \(\{a_m\}\)-order nice discrete process (see Theorem 28). Thus, the discrete version of the estimate of Young integrals (Remark 22) implies that \(I_0(N^{m,\rho})\) is an \(\{a_m\}\)-order nice discrete process. Noting that we have good estimates of \(H^-\)-Hölder norm of \(Y^{m,\rho}\), \(\tilde{J}^{m,\rho}\), \((-\tilde{J}^{m,\rho})^{-1}\) (Lemma 9, Lemma 16, Lemma 17) and that \(Z^{m,\rho}\) is an \(\{a_m\}\)-order nice discrete process (Theorem 28), we see that \(I_1(N^{m,\rho})\) is an \(\{a_m\}\)-order nice discrete process. Since \(d^m\) is an \(\{a_m\}\)-order nice discrete process, \(I_2(N^{m,\rho})\) is as well. As for \(I_3^{m,\rho}\), we already proved the assertion. Here we used Lemmas 16, 17, and 18 and Theorem 28. Since \(\sup_{\rho}\|\tilde{Z}^{m,\rho}\|_{H^-}=O(a_m)\) and other terms are \(\{a_m\}\)-order nice discrete processes, we have \(\sup_{\rho}\|N^{m,\rho}\|_{H^-}=O(a_m)\) which completes the proof of assertion (2). ◻

Theorem 31. Assume that Conditions  and  (1) hold. Let \(\varepsilon_1\) be the constant given in Condition 8. Set \(a_m=\max\{\Delta_m^{3H^--1},\Delta_m^{\varepsilon_1}\}\). Then we have \[\begin{align} \sup_{t,\rho}|\tilde{J}^{m,\rho}_t-J_t| &= O(a_m), & \sup_{t,\rho}|(\tilde{J}^{m,\rho}_t)^{-1}-J_t^{-1}| &= O(a_m) \end{align}\] in the sense of Definition  .

Proof. Note that \[\begin{align} \tilde{J}^{m,\rho}_t-\tilde{J}^m_t &= \int_0^\rho \partial_{\rho_1}J^{m,\rho_1}_t d\rho_1 = \int_0^\rho (-J^{m,\rho_1}_t)N^{m,\rho_1}_t d\rho_1,\\ (\tilde{J}^{m,\rho}_t)^{-1}-(\tilde{J}^m_t)^{-1} &= \int_0^{\rho} \partial_{\rho_1} (J^{m,\rho_1}_t)^{-1} d\rho_1 = \int_0^{\rho} N^{m,\rho_1}_t (J^{m,\rho_1}_t)^{-1} d\rho_1. \end{align}\] From Lemmas 15 and 22, we see that \(\sup_{t,\rho}|\tilde{J}^{m,\rho}_t-\tilde{J}^m_t|=O(a_m)\) and \(\sup_{t,\rho}|(\tilde{J}^{m,\rho}_t)^{-1}-(\tilde{J}^m_t)^{-1}|=O(a_m)\). This and Remark 30 yield the assertion. ◻

5 Proof of main theorem↩︎

We prove Theorem 12 and Corollary 1 in Section 5.2. Section 5.1 is a preparation for it.

5.1 Lemmas↩︎

Throughout this section, we assume that Conditions 8\(\sim\)11 hold. Recall that \((\lambda_1,\varepsilon_1, G_1)\) and \((\lambda_2,\varepsilon_2,G_2)\) are the triples of the two constants and the random variable specified in Conditions 8 and 11, respectively. Also, set \[\begin{align} a_m = \max\{\Delta_m^{3H^--1},\Delta_m^{4H^- - 2H - \frac{1}{2}},\Delta_m^{\varepsilon_1},\Delta_m^{\varepsilon_2}\}. \end{align}\] We will give estimate of \(\tilde{Z}^{m,\rho}(\omega)\) for \(\omega\in \Omega_0^{(m,d^m)}\). Precisely, we prove

Lemma 23. There exists a positive integer \(m_0\) such that for all \(p\ge 1\) it holds that \[\begin{align} \sup_{m\ge m_0} \left\|\sup_{0\le \rho\le 1} \|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-}1_{\Omega^{(m,d^m)}_0}\right\|_{L^p} <\infty. \end{align}\]

We refer the readers to Definition 4 and 28 for definition of \(\tilde{Z}^{m,\rho}_t\) and \(I^m\). We decompose as \(\tilde{Z}^{m,\rho}_t-I^m_t=\sum_{i=1}^5 S^{m,\rho,i}_t\), where \[\begin{align} S^{m,\rho,1}_t &= \sum_{i=1}^{2^mt} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} \left(c(Y^{m,\rho}_{\tau^m_{i-1}})-c(Y_{\tau^m_{i-1}})\right) d^m_{\tau^m_{i-1},\tau^m_i},\\ S^{m,\rho,2}_t &= \sum_{i=1}^{2^mt} \left((\tilde{J}^{m,\rho}_{\tau^m_i})^{-1}-(\tilde{J}^m_{\tau^m_i})^{-1}\right) c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i},\\ S^{m,\rho,3}_t &= \sum_{i=1}^{2^mt} \left((\tilde{J}^m_{\tau^m_i})^{-1}-J_{\tau^m_i}^{-1}\right) c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}, \quad S^{m,\rho,4}_t = \sum_{i=1}^{2^mt} J^{-1}_{\tau^m_{i-1},\tau^m_i}c(Y_{\tau^m_{i-1}})d^{m}_{\tau^m_{i-1},\tau^m_i},\\ S^{m,\rho,5}_t &= \sum_{i=1}^{2^mt} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1} \left(\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}-\epsilon^m_{\tau^m_{i-1},\tau^m_i}\right). \end{align}\]

We give estimates for each term \(S^{m,\rho,i}\) (\(1\le i\le 5\)). First, we consider \(S^{m,\rho,1}\).

Lemma 24. Let \(\omega\in \Omega_0^{(m,d^m)}\). Then we have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,1}\|_{\lambda_1}\le a_m C G_1 \sup_{\rho} \|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-}, \end{align}\] where \(C\) depends only on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. Set \(F^{m,\rho}_t=(\tilde{J}^{m,\rho}_{t+\Delta_m})^{-1}(c(Y^{m,\rho}_t)-c(Y_t))\). We have \[\begin{align} c(Y^{m,\rho}_t)-c(Y_t)&= \int_0^{\rho}(Dc)(Y^{m,\rho_1}_t)[Z^{m,\rho_1}_t]d\rho_1 = \int_0^{\rho}(Dc)(Y^{m,\rho_1}_t)[\tilde{J}^{m,\rho_1}_t \tilde{Z}^{m,\rho_1}_t]d\rho_1 \end{align}\] and we obtain Hölder estimate of the discrete process \(\| F^{m,\rho}\|_{H^-} \le C \sup_{\rho}\|\tilde{Z}^{m,\rho}\|_{H^-}\). Here, \(C\) depends on the Hölder norms of \(Y^{m,\rho}\) and \(\tilde{J}^{m,\rho}\). By combining the estimate \(\|d^m\|_{\lambda_1}\le 2^{-m\varepsilon_1}G_1\leq a_m G_1\) (\(\omega\in \Omega_0\)) and Remark 22, we complete the proof. ◻

Next, we consider \(S^{m,\rho,4}\) and \(S^{m,\rho, 5}\).

Lemma 25. Let \(\omega\in \Omega_0^{(m,d^m)}\). We have \[\begin{align} J^{-1}_{\tau^m_{i-1},\tau^m_i}c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i}= -J^{-1}_{\tau^m_{i-1}}(D\sigma)(Y_{\tau^m_{i-1}})[c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}]B_{\tau^m_{i-1},\tau^m_i} + O(\Delta_m^{4H^-}), \end{align}\] where the dominated random variable for the term \(O(\Delta_m^{4H^-})\) depends only on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. This follows from Lemma 17 and Remark 29. We used \(\lambda_1>H^-\). ◻

Lemma 26. Let \(\omega\in \Omega_0^{(m,d^m)}\). There exist \({\mathbb{R}}^n\)-valued bounded Lipschitz functions \(\varphi^{\alpha,\beta,\gamma}\), \(\psi_{\alpha}\), \(F_{\alpha,\beta,\gamma}\), \(F^1_{\alpha}\), \(F^2_{\alpha}\) on \({\mathbb{R}}^n\) \((1\le \alpha,\beta,\gamma\le d)\) such that \[\begin{align} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1}\big(\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}-\epsilon^m_{\tau^m_{i-1},\tau^m_i}\big)\\&= (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1} \Big\{ \sum_{\alpha,\beta,\gamma} \varphi_{\alpha,\beta,\gamma}(\hat{Y}_{\tau^m_{i-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{i-1},\tau^m_i} + \sum_{\alpha} \psi_{\alpha}(\hat{Y}^{m}_{\tau^m_{i-1}}) B^{\alpha}_{\tau^m_{i-1},\tau^m_i}\Delta_m\\ &\quad\quad + \sum_{\alpha,\beta,\gamma} F_{\alpha,\beta,\gamma}(Y_{\tau^m_{i-1}}) B^{\alpha,\beta,\gamma}_{\tau^m_{i-1},\tau^m_i} +\sum_{\alpha}F^1_{\alpha}(Y_{\tau^m_{i-1}})B^{0,\alpha}_{\tau^m_{i-1},\tau^m_i} +\sum_{\alpha}F^2_{\alpha}(Y_{\tau^m_{i-1}})B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i} \Big\}\\ &\quad\quad + O(\Delta_m^{4H^-}).\end{align}\] The dominated random variables for the terms \(O(\Delta_m)^{4H^-}\) depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. From 56 , Condition 9 (1) and Lemma 1 (1), we have \[\begin{align} (\tilde{J}^{m,\rho}_{\tau^m_i})^{-1}\big(\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}-\epsilon^m_{\tau^m_{i-1},\tau^m_i}\big) = (\tilde{J}^{m,\rho}_{\tau^m_{i-1}})^{-1}\big(\hat{\epsilon}^{m}_{\tau^m_{i-1},\tau^m_i}-\epsilon^m_{\tau^m_{i-1},\tau^m_i}\big) + O(\Delta_m^{4H^-}). \end{align}\] Combining this identity with Condition 9 (2) and Lemma 1 (2) yields the desired estimate. ◻

As we have shown in the above lemmas, we need estimates for weighted sum process in Wiener chaos of order \(3\) and sum process of \(d^{m,\alpha,\beta}_{\tau^m_{i-1},\tau^m_i}B^{\gamma}_{\tau^m_{i-1},\tau^m_i}\). We refer the readers to 29 for the definition of \(\mathcal{K}^3_m\).

Lemma 27. Let \(\omega\in \Omega_0^{(m,d^m)}\). Let \(K^{m}\in \mathcal{K}^3_{m}\) and \(\{F^m_t\}_{t\in D_m}\) be a discrete process satisfying \(|F^m_0|+\|F^m\|_{H^-}\le C\), where \(C\) is independent of \(m\) and depends only on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially. Let \(I^m(F^m)_t=\sum_{i=1}^{2^mt}F^m_{\tau^m_{i-1}}K^{m}_{\tau^m_{i-1},\tau^m_i}\) \((t\in D_m)\). Then it holds that \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}I^m(F^m)\|_{\lambda_2} \le a_m C G_2, \end{align}\] where \(C\) depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. By the assumption on the Hölder norm of \(F^m\) and Condition 11 and using Remark 22, we have \(\|(2^m)^{2H-\frac{1}{2}}I^m(F^m)\|_{\lambda_2} \le \Delta_m^{\varepsilon_2} C G_2,\) which implies the assertion. ◻

Lemma 28. Let \(\omega\in \Omega_0^{(m,d^m)}\). We have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,4}\|_{\lambda_2} + \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,5}\|_{\lambda_2} \le a_m C \{G_2+1\}, \end{align}\] where \(C\) depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. We use the decompositions in Lemmas 25 and 26. First, we consider the sum of \(O(\Delta_m^{4H^-})\). Let \(s=\tau^m_k<\tau^m_l=t\). We have \[\begin{align} \left| (2^m)^{2H-\frac{1}{2}} \sum_{i=k}^{l-1} O(\Delta_m^{4H^-}) \right| \le (2^m)^{2H-\frac{1}{2}}(l-k) C\Delta_m^{4H^-} =\Delta_m^{4H^--2H-\frac{1}{2}}C(t-s). \end{align}\] where \(C\) depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially. This term can be estimated as in the assertion. As for sum process \(K^m_{s,t}=\Delta_m B^{\alpha}_{s,t}\) which defined by the term \(\Delta_m B^{\alpha}_{\tau^m_{i-1},\tau^m_i}\) in Lemma 26, we have similar estimate to the elements in \(\mathcal{K}^3_m\). See the proof of Lemma 4. Note that we use Condition 3 only in that proof. The remaining main terms can be handled by Lemma 27 and Condition 11. This completes the proof. ◻

Remark 32. In the above Lemmas 27 and 28, we used the estimate of \(K^m_{s,t}\) which is defined as the sum process of \(B^{0,\alpha}_{\tau^m_{i-1},\tau^m_i}\) and \(B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}\) in Condition 11. If we use the estimate \(|B^{\alpha,0}_{\tau^m_{i-1},\tau^m_i}|\le C\Delta_m^{1+H^-}\), which follows form the Hölder estimate of \(B\) only, we obtain a rough estimate \(|(2^m)^{2H-\frac{1}{2}}K^m_{s,t}|\le C\Delta_m^{H^--(2H-\frac{1}{2})}|t-s|\) similarly to the estimate of \((2^m)^{2H-\frac{1}{2}}\sum O(\Delta_m^{4H^-})\) in the proof of Lemma 28. However, this estimate will give the estimate \(\varepsilon<\min\{3H^--1,H^--(2H-\frac{1}{2}),\varepsilon_1,\varepsilon_2\}\). Clearly this estimate gets worse as \(H\to \frac{1}{2}\).

We consider the estimates of \(S^{m,\rho,3}\). To this end, recall definition 28 of \(I^m\) and set \[\begin{align} \label{def32X} X_m &= \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-}. \end{align}\tag{80}\] Then from Condition 10, we have \(\sup_m\|X_m\|_{L^p}<\infty\) for all \(p\ge 1\).

Lemma 29. Let \(\omega\in \Omega_0^{(m,d^m)}\). We have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,3}\|_{H^-} &\le a_m C \{ X_m +G_2 +1 \}, \end{align}\] where \(C\) depends on \(\tilde{C}(B)\) and \(\tilde{N}(B)\) polynomially.

Proof. Let \(R\) be a positive integer. From Remark 30, we have \[\begin{align} (\tilde{J}^m_{\tau^m_i})^{-1}-J_{\tau^m_i}^{-1} &= K^{2,m,R}_{\tau^m_i}J_{\tau^m_i}^{-1} +L^{2,m,R}_{\tau^m_i} \nonumber \\ &= K^{2,m,R}_{\tau^m_i}J^{-1}_{\tau^m_{i-1}} +K^{2,m,R}_{\tau^m_i}J^{-1}_{\tau^m_{i-1},\tau^m_i} +L^{2,m,R}_{\tau^m_i}. \label{eq48930184231} \end{align}\tag{81}\] where \(K^{2,m,R}\) is an \(\{a_m\}\)-order nice discrete processes and \(L^{2,m,R}\) is a small discrete process. Hence \[\begin{align} S^{m,\rho,3}_t &= \sum_{i=1}^{2^m t} \left( K^{2,m,R}_{\tau^m_i}J^{-1}_{\tau^m_{i-1}} +K^{2,m,R}_{\tau^m_i}J^{-1}_{\tau^m_{i-1},\tau^m_i} +L^{2,m,R}_{\tau^m_i} \right) c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}\\ &= S^{m,\rho,3,1}_t +S^{m,\rho,3,2}_t +S^{m,\rho,3,3}_t, \end{align}\] Then with the help of the summation by parts formula 45 , we have \[\begin{align} S^{m,\rho,3,1}_t = \sum_{i=1}^{2^m t} K^{2,m,R}_{\tau^m_i} I^m_{\tau^m_{i-1},\tau^m_i} = K^{2,m,R}_t I^m_t - \sum_{i=1}^{2^m t} K^{2,m,R}_{\tau^m_{i-1},\tau^m_i} I^m_{\tau^m_{i-1}}. \end{align}\] Recalling that \((2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\) is discrete \(H^-\)-Hölder continuous and using Remark 30, using \(X_m\) defined by 80 , we have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,3,1}\|_{H^-}\\&\leq 2 \|K^{2,m,R}\|_{H^-} \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-} + \left\| \sum_{i=1}^{2^m \cdot} K^{2,m,R}_{\tau^m_{i-1},\tau^m_i} (2^m)^{2H-\frac{1}{2}}I^m_{\tau^m_{i-1}} \right\|_{H^-}\\ &\leq C \{ a_m \cdot X_m + a_m \cdot X_m \}.\end{align}\] In a similar way to Lemma 28, using Lemma 25, we have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}} S^{m,\rho,3,2} \|_{\lambda_2} \le a_m C\{ G_2+1 \}. \end{align}\] The term \(\|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,3,3}\|_{H^-}\) becomes small for large \(R\). The proof is completed. ◻

Finally, we estimate \(S^{m,\rho,2}\). To this end, we use \(N^{m,\rho}_t=(-\tilde{J}^{m,\rho}_t)^{-1}\partial_{\rho}\tilde{J}^{m,\rho}_t\), which is introduced in Section 4.4.

Lemma 30. Let \(L\) be a positive integer. Then it holds that \[\begin{gather} (\tilde{J}^{m,\rho}_t)^{-1}-(\tilde{J}^m_t)^{-1} = \sum_{l=1}^{L-1} \int_{0<\rho_l<\cdots<\rho_1<\rho} d\rho_1 \cdots d\rho_l\, N^{m,\rho_1}_t\cdots N^{m,\rho_l}_t (\tilde{J}^m_t)^{-1}\\ + \int_{0<\rho_{L}<\cdots<\rho_1<\rho} d\rho_1 \cdots d\rho_L\, N^{m,\rho_1}_t \cdots N^{m,\rho_L}_t (\tilde{J}^{m,\rho_L}_t)^{-1}. \end{gather}\]

Proof. Noting \(\partial_{\rho}(\tilde{J}^{m,\rho}_t)^{-1}= -(\tilde{J}^{m,\rho}_t)^{-1}\partial_{\rho}\tilde{J}^{m,\rho}_t (\tilde{J}^{m,\rho}_t)^{-1}=N^{m,\rho}_t(\tilde{J}^{m,\rho}_t)^{-1}\), we have \[\begin{align} (\tilde{J}^{m,\rho}_t)^{-1}-(\tilde{J}^m_t)^{-1} = \int_{0<\rho_1<\rho} d\rho_1\, N^{m,\rho_1}_t (\tilde{J}^{m,\rho_1}_t)^{-1}\\&= \int_{0<\rho_1<\rho} d\rho_1\, N^{m,\rho_1}_t (\tilde{J}^m_t)^{-1} + \int_{0<\rho_1<\rho} d\rho_1 N^{m,\rho_1}_t \left\{ (\tilde{J}^{m,\rho_1}_t)^{-1}-(\tilde{J}^m_t)^{-1} \right\}\\ &= \int_{0<\rho_1<\rho} d\rho_1\, N^{m,\rho_1}_t (\tilde{J}^m_t)^{-1} + \int_{0<\rho_1<\rho} d\rho_1 N^{m,\rho_1}_t \int_{0<\rho_2<\rho_1} d\rho_2\, N^{m,\rho_2}_t (\tilde{J}^{m,\rho_2}_t)^{-1}.\end{align}\] Iterating this calculation, we are done. ◻

Lemma 31. For \(\omega\in\Omega_0^{(m,d^m)}\), we have \[\begin{align} \|(2^m)^{2H-\frac{1}{2}}S^{m,\rho,2}\|_{H^-} \le a_m C \big\{ G_1 \sup_{\rho}\|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-} +X_m +G_2 +1 \big\}, \end{align}\] where \(C\) depends on \(\tilde{C}(B), \tilde{N}(B)\) polynomially.

Proof. We use the same notation as in Lemmas 22 and 30. Set \[\begin{align} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_t &= \prod_{r=1}^l \left\{ N^{m,\rho_r}_t - \sum_{\nu=1}^n \tilde{Z}^{m,\rho_r,\nu}_t I^{m,\rho_r}(\varphi_\nu)_t \right\} = \prod_{r=1}^l \sum_{\lambda=0}^3 I_\lambda(N^{m,\rho_r})_t,\\ R^{m,\rho_1,\ldots,\rho_l}_t &= N^{m,\rho_1}_t\cdots N^{m,\rho_l}_t - \tilde{N}^{m,\rho_1,\ldots,\rho_l}_t. \end{align}\] Note that the product \(\prod_{r=1}^l\) in the above equation should be taken according to the order. Then we have \(S^{m,\rho,2}_t=S^{m,\rho,2,1}_t+S^{m,\rho,2,2}_t+S^{m,\rho,2,3}_t\), where \[\begin{align} S^{m,\rho,2,1}_t &= \sum_{l=1}^{L-1} \int_{0<\rho_l<\cdots<\rho_1<\rho} d\rho_1 \cdots d\rho_l\, \sum_{i=1}^{2^mt} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (\tilde{J}^m_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i},\\ S^{m,\rho,2,2}_t &= \sum_{l=1}^{L-1} \int_{0<\rho_l<\cdots<\rho_1<\rho} d\rho_1 \cdots d\rho_l\, \sum_{i=1}^{2^mt} R^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (\tilde{J}^m_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i},\\ S^{m,\rho,2,3}_t &= \int_{0<\rho_{L}<\cdots<\rho_1<\rho} d\rho_1 \cdots d\rho_L\, \sum_{i=1}^{2^mt} N^{m,\rho_1}_{\tau^m_i} \cdots N^{m,\rho_L}_{\tau^m_i} (\tilde{J}^{m,\rho_L}_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i}. \end{align}\] We estimate the terms above.

By the definition, all terms in the expansion of \(R^{m,\rho_1,\ldots,\rho_l}_t\) are given by the product of \(l\) terms from \(N^{m,\rho_r}_t\) and \(\tilde{Z}^{m,\rho_r,\nu}_t I^{m,\rho_r}(\varphi_\nu)_t\) (\(1\leq r\leq l\), \(1\leq \nu\leq n\)) and each term contains at least one \(\tilde{Z}^{m,\rho_r,\nu}_t I^{m,\rho_r}(\varphi_\nu)_t\). Thus, using Remark 22, Lemmas 18 and 22, we have \[\begin{align} \bigg\| (2^m)^{2H-\frac{1}{2}} \sum_{i=1}^{2^m \cdot} R^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (\tilde{J}^m_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i} \Big\|_{\lambda_1}\\&\leq C \| (2^m)^{2H-\frac{1}{2}} R^{m,\rho_1,\ldots,\rho_l} \|_{H^-} \bigg\| \sum_{i=1}^{2^m \cdot} (\tilde{J}^m_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i} \bigg\|_{\lambda_1}\\ &\leq C \|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-} \cdot a_m C G_1,\end{align}\] from which we obtain an estimate of \(S^{m,\rho,2,2}\). We next consider \(S^{m,\rho,2,1}\). Noting 81 , we have \[\begin{align} \sum_{i=1}^{2^mt} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (\tilde{J}^m_{\tau^m_i})^{-1} c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i} = \sum_{i=1}^{2^mt} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (I+K^{2,m,R}_{\tau^m_i}) J_{\tau^m_{i-1}}^{-1} c(Y_{\tau^m_{i-1}})d^m_{\tau^m_{i-1},\tau^m_i}\\&+ \sum_{i=1}^{2^mt} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} (I+K^{2,m,R}_{\tau^m_i}) J_{\tau^m_{i-1},\tau^m_i}^{-1} c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}\\ &+ \sum_{i=1}^{2^mt} \tilde{N}^{m,\rho_1,\ldots,\rho_l}_{\tau^m_i} L^{2,m,R}_{\tau^m_i} c(Y_{\tau^m_{i-1}}) d^m_{\tau^m_{i-1},\tau^m_i}.\end{align}\] All terms can be treated in the similar way as Lemma 29 because \(\tilde{N}^{m,\rho_1,\ldots,\rho_l}\) is an \(\{a_m\}\)-order nice discrete process independent of \(\rho_1,\ldots,\rho_l\) (see Lemma 22).

Finally, we consider \(S^{m,\rho,2,3}\). Noting that \[\begin{align} \sup_{\rho_1,\ldots,\rho_L}\|N^{m,\rho_1}\cdots N^{m,\rho_L}\|_{H^-} =O(a_m^L), \end{align}\] we see that this term is small for large \(L\). This completes the proof. ◻

Proof of Lemma . We write \(f_m = \sup_{\rho}\|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-}1_{\Omega_0^{(m,d^m)}}\). From the lemmas above, there exist random variables \(\{\varGamma_m\}\) and \(\varGamma\) defined on \(\Omega_0\) which satisfy \(\sup_m\|\varGamma_m\|_{L^p}<\infty\) for all \(p\ge 1\) and \(\varGamma\in \cap_{p\ge 1}L^p(\Omega_0)\) such that \(f_m\le \varGamma_m+a_m \varGamma f_m\). Recalling \(\tilde{Z}^{m,\rho}\) is an \(\{a_m\}\)-order nice discrete process independent of \(\rho\) (Theorem 28), there exists \(\varGamma'\) such that \(f_m\leq (2^m)^{2H-\frac{1}{2}}\varGamma'\) and \(\varGamma'\in \cap_{p\ge 1}L^p(\Omega_0)\). By using this inequality \(L\)-times and Theorem 28, we get \[\begin{align} f_m &\le \left\{ \sum_{l=0}^{L-1} (a_m \varGamma)^l \right\} \varGamma_m + (a_m \varGamma)^{L} f_m \le \left\{ \sum_{l=0}^{L-1} (a_m \varGamma)^l \right\} \varGamma_m + (2^m)^{2H-\frac{1}{2}}(a_m \varGamma)^{L} \varGamma'. \end{align}\] By taking \(L\) to be sufficiently large, we arrive at the conclusion. ◻

Finally, using Lemma 23, we prove an estimate of \(\tilde{Z}^{m,\rho}-I^m\).

Lemma 32. Let \(\varepsilon_1\) and \(\varepsilon_2\) be constants specified in Conditions 8 and 11, respectively. Let \(0<\varepsilon<\min \{3H^--1,4H^--2H-\frac{1}{2},\varepsilon_1,\varepsilon_2\}\). Then, for all \(p\ge 1\) it holds that \[\begin{align} \lim_{m\to\infty} \left\| \sup_{0\le \rho\le 1} \|(2^m)^{2H-\frac{1}{2}+\varepsilon}(\tilde{Z}^{m,\rho}-I^m)\|_{H^-}1_{\Omega^{(m,d^m)}_0}\right\|_{L^p} = 0. \end{align}\]

Proof. Write \(f_m=\sup_{\rho}\|(2^m)^{2H-\frac{1}{2}}\tilde{Z}^{m,\rho}\|_{H^-}1_{\Omega_0^{(m,d^m)}}\). Lemmas 24, 31 imply \[\begin{align} \|(2^m)^{2H-\frac{1}{2}+\varepsilon}S^{m,\rho,1}\|_{H^-}1_{\Omega_0^{(m,d^m)}} &\leq (2^m)^\varepsilon \cdot a_m CG_1 f_m,\\ \|(2^m)^{2H-\frac{1}{2}+\varepsilon_1}S^{m,\rho,2}\|_{H^-}1_{\Omega_0^{(m,d^m)}} &\leq (2^m)^\varepsilon \cdot a_m C \big\{ G_1f_m +X_m +G_2 +1 \big\}. \end{align}\] Lemmas 28 and 29 gives similar estimates for \(\|(2^m)^{2H-\frac{1}{2}+\varepsilon}S^{m,\rho,r}\|_{H^-}1_{\Omega_0^{(m,d^m)}}\) for \(r=3,4,5\). Combining these estimates and Lemma 23, the proof is finished. ◻

5.2 Proofs of Theorem 12 and Corollary 1↩︎

Here we show Theorem 12 and Corollary 1.

Proof of Theorem 12. Recall that \(R^m_t\) \((t\in D_m)\) is defined by 31 . We will first consider \(R^m_t1_{\Omega_0^{(m,d^m)}}\), then \(R^m_t1_{(\Omega_0^{(m,d^m)})^\complement}\). Proposition 26 implies \[\begin{align} R^m_t1_{\Omega_0^{(m,d^m)}} = (\hat{Y}^{m}_t-Y_t-J_tI^m_t)1_{\Omega_0^{(m,d^m)}} = \int_0^1 \{ \tilde{J}^{m,\rho}_t \tilde{Z}^{m,\rho}_t - J_tI^m_t \} 1_{\Omega_0^{(m,d^m)}}\, d\rho. \end{align}\] The integrand scaled by \((2^m)^{2H-\frac{1}{2}+\varepsilon}\) is decomposed into \[\begin{gather} (2^m)^{2H-\frac{1}{2}+\varepsilon} \{ \tilde{J}^{m,\rho}_t \tilde{Z}^{m,\rho}_t - J_tI^m_t \} 1_{\Omega_0^{(m,d^m)}} = \tilde{J}^{m,\rho}_t \cdot (2^m)^{2H-\frac{1}{2}+\varepsilon} \big( \tilde{Z}^{m,\rho}_t-I^m_t \big) 1_{\Omega_0^{(m,d^m)}}\\ + (2^m)^{\varepsilon} \big(\tilde{J}^{m,\rho}_t-J_t\big) 1_{\Omega_0^{(m,d^m)}} \cdot (2^m)^{2H-\frac{1}{2}} I^m_t. \end{gather}\] Hence we have \[\begin{gather} (2^m)^{2H-\frac{1}{2}+\varepsilon} \max_{t\in D_m}|R^m_t1_{\Omega_0^{(m,d^m)}}|\\ \leq \big( \max_t | \tilde{J}^{m,\rho}_t | \big) \big( \sup_{\rho} \|(2^m)^{2H-\frac{1}{2}+\varepsilon}(\tilde{Z}^{m,\rho}-I^m)\|_{H^-} \big) 1_{\Omega^{(m,d^m)}_0}\\ + (2^m)^{\varepsilon} \big( \sup_{t,\rho}|\tilde{J}^{m,\rho}_t-J_t| \big) 1_{\Omega^{(m,d^m)}_0} \cdot \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-}. \end{gather}\] Here, \(I^m|_{D_m}\) denote the discrete process defined as the restriction of \(I^m\) on \(D_m\). The first term in the right-hand side converges to \(0\) due to Lemmas 20 and 32. The second term converges to \(0\) follows from Theorem 31 and Condition 10. From this we have \((2^m)^{2H-\frac{1}{2}+\varepsilon} \max_{t\in D_m}|R^m_t1_{\Omega_0^{(m,d^m)}}|\) converges to \(0\) in \(L^p\).

Next we consider \(R^m_t1_{(\Omega_0^{(m,d^m)})^\complement}\). Noting \[\begin{align} (2^m)^{2H-\frac{1}{2}+\varepsilon} R^m_t1_{(\Omega_0^{(m,d^m)})^\complement} &= \big(\hat{Y}^{m}_t-Y_t\big) \cdot (2^m)^{2H-\frac{1}{2}+\varepsilon} 1_{(\Omega_0^{(m,d^m)})^\complement}\\ &\phantom{\leq}\qquad - J_t \cdot (2^m)^{2H-\frac{1}{2}} I^m_t \cdot (2^m)^{\varepsilon} 1_{(\Omega_0^{(m,d^m)})^\complement}, \end{align}\] we have \[\begin{gather} (2^m)^{2H-\frac{1}{2}+\varepsilon} \max_{t\in D_m} |R^m_t1_{(\Omega_0^{(m,d^m)})^\complement}| \leq \big( \max_t | \hat{Y}^{m}_t-Y_t | \big) \cdot (2^m)^{2H-\frac{1}{2}+\varepsilon} 1_{(\Omega_0^{(m,d^m)})^\complement}\\ + \big( \max_t | J_t | \big) \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-} \cdot (2^m)^{\varepsilon} 1_{(\Omega_0^{(m,d^m)})^\complement}. \end{gather}\] Lemma 9 and Remark 29 imply that \(\max_t | \hat{Y}^{m}_t - Y_t |\) and \(\max_t | J_t |\) are bounded from above by \(\cap_{p\ge 1}L^p\) random variable. By using 42 and Condition 10, both terms of the right-hand side converge to \(0\) in \(L^p\). The proof is completed. ◻

Proof of Corollary . Recall that \(R^m_t\) \((0\leq t\leq 1)\) is defined by 33 . Since \(R^m_t=R^m_{\tau^m_{k-1}}+(R^m_t-R^m_{\tau^m_{k-1}})\) for \(\tau^m_{k-1}\leq t\leq \tau^m_k\), we have \[\begin{align} \max_{0\leq t\leq 1} |R^m_t| \leq \max_{t\in D_m} |R^m_t| + \max_{1\leq k\leq 2^m} \max_{\tau^m_{k-1}\leq t\leq \tau^m_k} |R^m_t-R^m_{\tau^m_{k-1}}|. \end{align}\] Since the first term is estimated in Theorem 12, we give an estimate of the second term. Let \(\tau^m_{k-1}\leq t\leq \tau^m_k\). We decompose \(R^m_t-R^m_{\tau^m_{k-1}}\) into two terms; \[\begin{align} \Phi^m_1(t) &= \hat{Y}^{m}_t-\hat{Y}^{m}_{\tau^m_{k-1}}-(Y_t-Y_{\tau^m_{k-1}}), & \Phi^m_2(t) &= J_{\tau^m_{k-1}}I^m_{\tau^m_{k-1}}-J_tI^m_t. \end{align}\] We have \[\begin{align} \Phi^m_1(t)&=\big\{\sigma(\hat{Y}^{m}_{\tau^m_{k-1}})-\sigma(Y_{\tau^m_{k-1}})\big\}B_{\tau^m_{k-1},t} +\big\{((D\sigma)[\sigma])(\hat{Y}^{m}_{\tau^m_{k-1}})-((D\sigma)[\sigma])(Y_{\tau^m_{k-1}})\big\} {\mathbb{B}}_{\tau^m_{k-1},t}\\ &\qquad +\big\{b(\hat{Y}^{m}_{\tau^m_{k-1}})-b(Y_{\tau^m_{k-1}})\big\}(t-\tau^m_{k-1})+ c(\hat{Y}^{m}_{\tau^m_{k-1}})d^{m}_{\tau^m_{k-1},t}+\big\{\hat{\epsilon}^{m}_{\tau^m_{k-1},t}-\epsilon^m_{\tau^m_{k-1},t}\big\}, \end{align}\] which implies \[\begin{align} |\Phi^m_1(t)| &\leq C \big\{ |\hat{Y}^{m}_{\tau^m_{k-1}}-Y_{\tau^m_{k-1}}|\Delta_m^{H^-} +\hat{X}\Delta_m^{2H^-} +\hat{X}\Delta_m^{3H^-} \big\}\\ &\leq C \big\{ |J_{\tau^m_{k-1}}I^m_{\tau^m_{k-1}}+R^m_{\tau^m_{k-1}}|\Delta_m^{H^-} +\hat{X}\Delta_m^{2H^-} \big\}. \end{align}\] Here \(C\) is a constant depending on \(\sigma\), \(b\), \(c\) and \(C(B)\). From this we obtain \[\begin{align} (2^m)^{2H-\frac{1}{2}+\varepsilon} |\Phi^m_1(t)| &\leq C \big( 1+\|J\|_{H^-} \big) \|(2^m)^{2H-\frac{1}{2}}I^m|_{D_m}\|_{H^-} \Delta_m^{H^--\varepsilon}\\ &\qquad + C \big\{ (2^m)^{2H-\frac{1}{2}+\varepsilon} \max_k|R^m_{\tau^m_k}| \big\} \Delta_m^{H^-} + C \hat{X} \Delta_m^{\frac{1}{2}-2H+2H^--\varepsilon}\\ &=: CX_{m,1} \Delta_m^{H^--\varepsilon} +CX_{m,2} \Delta_m^{H^-} +C \hat{X} \Delta_m^{\frac{1}{2}-2H+2H^--\varepsilon}. \end{align}\] We have \(I^m_t=I^m_{\tau^m_{k-1}}\) \((\tau^m_{k-1}\le t\le \tau^m_k)\), which implies \[\begin{align} (2^m)^{2H-\frac{1}{2}+\varepsilon}|\Phi^m_2(t)| = (2^m)^{\varepsilon} |J_{\tau^m_{k-1}}-J_t| |(2^m)^{2H-\frac{1}{2}}I^m_{\tau^m_{k-1}}| \le X_{m,1} \Delta_m^{H^--\varepsilon}. \end{align}\] Noting that the right-hand sides in the two estimates are independent of \(k\), we have \[\begin{align} (2^m)^{2H-\frac{1}{2}+\varepsilon} \max_{1\leq k\leq 2^m} \max_{\tau^m_{k-1}\leq t\leq \tau^m_k} |R^m_t-R^m_{\tau^m_{k-1}}| &\leq (C+1)X_{m,1} \Delta_m^{H^--\varepsilon} +CX_{m,2} \Delta_m^{H^-}\\ &\phantom{\leq} \qquad +C \hat{X} \Delta_m^{\frac{1}{2}-2H+2H^--\varepsilon} \end{align}\] We see that \(\sup_m\{\|X_{m,1}\|_{L^p},\|X_{m,2}\|_{L^p}, \|\hat{X}\|_{L^p}\}<\infty\) for all \(p\ge 1\) which follows from Lemma 20, Remark 29, Condition 10 and Theorem 12. Hence noting \(3H^--1\leq H^-\) and \(3H^--1\leq \frac{1}{2}-2H+2H^-\), we complete the proof. ◻

Acknowledgment    The authors thank the anonymous referees for their helpful comments which improved the quality of this paper. This work was supported by JSPS KAKENHI Grant Numbers JP20H01804 and JP22K13932.

Shigeki Aida
Graduate School of Mathematical Sciences,
The University of Tokyo,
Meguro-ku, Tokyo, 153-8914, Japan

E-mail address: aida@ms.u-tokyo.ac.jp

Nobuaki Naganuma
Faculty of Advanced Science and Technology,
Kumamoto University,
Kumamoto city, Kumamoto, 860-8555, Japan

E-mail address: naganuma@kumamoto-u.ac.jp

References↩︎

[1]
I. Nourdin, D. Nualart, and C. A. Tudor, Central and non-central limit theorems for weighted power variations of fractional Brownian motion. Ann. Inst. Henri Poincaré Probab. Stat. 46 (2010), 1055–1079.
[2]
Y. Liu and S. Tindel, Discrete rough paths and limit theorems. Ann. Inst. Henri Poincaré Probab. Stat. 56 (2020), no. 3, 1730–1774.
[3]
I. Nourdin and G. Peccati, Normal approximations with Malliavin calculus. From Stein’s method to universality. Cambridge Tracts in Mathematics, 192. Cambridge University Press, Cambridge, (2012).
[4]
S. Aida and N. Naganuma, Error analysis for approximations to one-dimensional SDEs via the perturbation method, Osaka J. Math. 57 (2020), no. 2, 381–424.
[5]
Y. Hu, Y. Liu, and D. Nualart, Rate of convergence and asymptotic error distribution of Euler approximation schemes for fractional diffusions. Ann. Appl. Probab. 26 (2016), no. 2, 1147–1207.
[6]
Y. Hu and Y. Liu, and D. Nualart, Crank-Nicolson scheme for stochastic differential equations driven by fractional Brownian motions. Ann. Appl. Probab. 31 (2021), no. 1, 39–83.
[7]
Y. Liu and S. Tindel, First-order Euler scheme for SDEs driven by fractional Brownian motions: the rough case. Ann. Appl. Probab. 29 (2019), no. 2, 758–826.
[8]
N. Naganuma, Asymptotic error distributions of the Crank-Nicholson scheme for SDEs driven by fractional Brownian motion. J. Theoret. Probab. 28 (2015), no. 3, 1082–1124.
[9]
K. Ueda, Error distribution for one-dimensional stochastic differential equations driven by fractional Brownian motion. J. Theoret. Probab. 38 (2025), no. 1, Paper No. 20, 61 pp.
[10]
A.M. Davie, Differential equations driven by rough paths: an approach via discrete approximations, Appl. Math. Res. Express. AMRX (2008), Art. ID abm009, 40 pp.
[11]
P. Friz and N. Victoir, Multidimensional Stochastic Processes as Rough Paths. Theory and Applications. Cambridge Studies in Advanced Mathematics, 120, Cambridge University Press (2010).
[12]
S. Aida and N. Naganuma, Hölder estimates and weak convergences of certain weighted sum processes, arXiv:2408.05255.
[13]
T. Cass, C. Litterer, T. Lyons, Integrability and tail estimates for Gaussian rough differential equations. Ann. Probab. 41 (2013), no. 4, 3026–3050.
[14]
P. Friz and M. Hairer, A Course on Rough Paths. With an Introduction to Regularity Structures. Second edition. Universitext, Springer, Cham, (2020).
[15]
T. Lyons and Z. Qian, System control and rough paths. Oxford Math. Monogr. Oxford Sci. Publ. Oxford University Press, Oxford, (2002).
[16]
P. Friz and N. Victoir, Differential equations driven by Gaussian signals. Ann. Inst. Henri Poincaré Probab. Stat. 46 (2010), no. 2, 369–413.
[17]
N. Naganuma, Exact convergence rate of the Wong-Zakai approximation to RDEs driven by Gaussian rough paths. Stochastics 88 (2016), no. 7, 1041–1059.