
We have studied a strongly nonlinear backward stochastic partial differential equation (B-SPDE) through an approximation method and with machine learning (ML)-based Monte Carlo simulation. This equation is well-known and was previously derived from studies in finance. However, how to analyze and solve this equation has remained a problem for quite a long time. The main difficulty is due to the singularity of the B-SPDE since it is a strongly nonlinear one. Therefore, by introducing new truncation operators and integrating the machine learning technique into the platform of a convolutional neural network (CNN), we have developed an effective approximation method with a Monte Carlo simulation algorithm to tackle the well-known open problem. In doing so, the existence and uniqueness of a 2-tuple adapted strong solution to an approximation B-SPDE were proved. Meanwhile, the convergence of a newly designed simulation algorithm was established. Simulation examples and an application in finance were also provided.
Citation: Wanyang Dai. Simulating a strongly nonlinear backward stochastic partial differential equation via efficient approximation and machine learning[J]. AIMS Mathematics, 2024, 9(7): 18688-18711. doi: 10.3934/math.2024909
[1] | Min-Ku Lee, Jeong-Hoon Kim . Closed-form approximate solutions for stop-loss and Russian options with multiscale stochastic volatility. AIMS Mathematics, 2023, 8(10): 25164-25194. doi: 10.3934/math.20231284 |
[2] | Xiuxian Chen, Zhongyang Sun, Dan Zhu . Mean-variance investment and risk control strategies for a dynamic contagion process with diffusion. AIMS Mathematics, 2024, 9(11): 33062-33086. doi: 10.3934/math.20241580 |
[3] | Meijiao Wang, Qiuhong Shi, Maoning Tang, Qingxin Meng . Stochastic differential equations in infinite dimensional Hilbert space and its optimal control problem with Lévy processes. AIMS Mathematics, 2022, 7(2): 2427-2455. doi: 10.3934/math.2022137 |
[4] | Boubaker Smii . Representation of the solution of a nonlinear molecular beam epitaxy equation. AIMS Mathematics, 2024, 9(12): 36012-36030. doi: 10.3934/math.20241708 |
[5] | Jafar Biazar, Fereshteh Goldoust . Multi-dimensional Legendre wavelets approach on the Black-Scholes and Heston Cox Ingersoll Ross equations. AIMS Mathematics, 2019, 4(4): 1046-1064. doi: 10.3934/math.2019.4.1046 |
[6] | Abdelwahed Motwake, Aisha Hassan Abdalla Hashim, Marwa Obayya, Majdy M. Eltahir . Enhancing land cover classification in remote sensing imagery using an optimal deep learning model. AIMS Mathematics, 2024, 9(1): 140-159. doi: 10.3934/math.2024009 |
[7] | Xiaoming Wang, Muhammad W. Yasin, Nauman Ahmed, Muhammad Rafiq, Muhammad Abbas . Numerical approximations of stochastic Gray-Scott model with two novel schemes. AIMS Mathematics, 2023, 8(3): 5124-5147. doi: 10.3934/math.2023257 |
[8] | Mashael M Asiri, Abdelwahed Motwakel, Suhanda Drar . Robust sign language detection for hearing disabled persons by Improved Coyote Optimization Algorithm with deep learning. AIMS Mathematics, 2024, 9(6): 15911-15927. doi: 10.3934/math.2024769 |
[9] | Yao Fu, Sisi Zhou, Xin Li, Feng Rao . Multi-assets Asian rainbow options pricing with stochastic interest rates obeying the Vasicek model. AIMS Mathematics, 2023, 8(5): 10685-10710. doi: 10.3934/math.2023542 |
[10] | Wei Xue, Pengcheng Wan, Qiao Li, Ping Zhong, Gaohang Yu, Tao Tao . An online conjugate gradient algorithm for large-scale data analysis in machine learning. AIMS Mathematics, 2021, 6(2): 1515-1537. doi: 10.3934/math.2021092 |
We have studied a strongly nonlinear backward stochastic partial differential equation (B-SPDE) through an approximation method and with machine learning (ML)-based Monte Carlo simulation. This equation is well-known and was previously derived from studies in finance. However, how to analyze and solve this equation has remained a problem for quite a long time. The main difficulty is due to the singularity of the B-SPDE since it is a strongly nonlinear one. Therefore, by introducing new truncation operators and integrating the machine learning technique into the platform of a convolutional neural network (CNN), we have developed an effective approximation method with a Monte Carlo simulation algorithm to tackle the well-known open problem. In doing so, the existence and uniqueness of a 2-tuple adapted strong solution to an approximation B-SPDE were proved. Meanwhile, the convergence of a newly designed simulation algorithm was established. Simulation examples and an application in finance were also provided.
In this paper, we study a strongly nonlinear backward stochastic partial differential equation (B-SPDE) in (1.1) through an approximation method and with machine learning (ML)-based Monte Carlo simulation,
V(t,x)=H(T,x)−12∫Tt(Vx(s,x)+ˉVx(s,x))2Vxx(s,x)ds−∫TtˉV(s,x)dW(s), | (1.1) |
which has a given terminal random field H(T,x) at time T for (t,x)∈[0,T]×R and R=(−∞,∞) and which is driven by a Brownian motion W(⋅) over time interval [0,T].
The equation in (1.1) is well-known and was previously derived from studies in finance (see Musiela and Zariphopoulou [22], ∅ksendal et al. [24], etc.), where, the primary motivation to study the B-SPDE was to solve an optimal portfolio selection problem in an incomplete financial market modeled by a stochastic differential equation driven by multi-dimensional Brownian motion (see, e.g., Musiela and Zariphopoulou [22] and Kramkov and Sirbu [18]). Furthermore, the problem describes the evolution of a related value function and seeks to maximize an expected utility from terminal wealth over admissible strategies. To find an optimal investment policy for this stochastic optimization problem, the well-known dynamic programming principle can be applied to determine such an optimal control policy in a backward way with respect to the time parameter. More precisely, in this financial problem, the random field V(t,x) denotes the value function of an investor's target to maximize his expected utility of terminal wealth over admissible strategy set AT (see its definition in Subsection 2.2) with t∈[0,T], i.e.,
V(t,x)=supβ∈ATE[μT(Xβ(T))|Ft,X(t)=x] | (1.2) |
for a given trading horizon [0,T] and the investor's utility μT(⋅):R+→R at terminal time T (see more explanations in Subsection 2.2). Note that, in this case, we take H(T,⋅)=μT(⋅) in the equation of (1.1). Furthermore, Xβ(t) denotes the present value of the investor's aggregate investment at time t (see more details in Subsection 2.2). In existing studies (see, e.g., Cerny and Kallsen [2] and Dai [8]), the authors aim to find a so-called variance optimal martingale measure Q∗ such that the value function in (1.2) has a simple expression given by the following conditional expectation,
V(t,x)=EQ∗[H(T,x)|Ft,X(t)=x]. | (1.3) |
In this sense, V(t,x) can be decomposed into a macro-trend part and a micro-regulating (volatility) part as shown in (1.1) due to martingale representation theorem (see, e.g., ∅ksendal [23]). Roughly speaking, the random field ˉV(⋅,⋅) corresponds to the volatility rate and is the Malliavin derivative of V(t,x) with respect to the Wiener measure corresponding to the Brownian motion W (see more details in Dai [9]).
Due to some difficulties with solving the problem in (1.2) directly and as an alternative to the method presented in (1.3), the authors in Musiela and Zariphopoulou [22] derived the B-SPDE to solve the optimal investment problem. In this sense, the B-SPDE might be considered as the non-Markovian analogue of the traditional Hamilton-Jacobi-Bellman (HJB) equation in Markovian models or in its dual formulation. Once we obtain the solution to the B-SPDE, we can determine the optimal investment policy (see Subsection 2.2 for more details). However, how to analyze and solve Eq (1.1) in general has remained a problem for quite a long time. The main difficulty is due to the singularity of the B-SPDE since it is a strongly nonlinear one. To develop a method to solve this problem, we need to go over the work in Dai [9] and make a comparison between it and our current study.
More precisely, in Dai [9], we developed a generic convolutional neural network (CNN)-based numerical scheme to simulate the 2-tuple adapted strong solution to a unified system of B-SPDEs driven by Brownian motions, which can be applied to many B-SPDE equations. Nevertheless, in proving the unique existence of the 2-tuple adapted strong solution to the unified system, we need to impose the so-called general local Lipschitz and linear growth conditions. Furthermore, in Dai [9], the generic numerical scheme was developed by a CNN through conditional expectation projection, which is a completely discrete and iterative algorithm in terms of both time and space. However, in estimating the mean-square error and proving the convergence for the CNN-based numerical scheme, we also need the general local Lipschitz and linear growth conditions. Moreover, the generic numerical scheme in Dai [9] does not integrate ML techniques into its computation algorithm.
Although the equation in (1.1) is a special case of our unified system of B-SPDEs in Dai [9], it is a strongly nonlinear one with singularity. Therefore, the B-SPDE in (1.1) does not satisfy the imposed general local Lipschitz and linear growth conditions in Dai [9], which implies that the developed method in Dai [9] cannot be directly applied to solve the equation in (1.1). Hence, by introducing new truncation operators and integrating the ML technique into the CNN platform in Dai [9], we develop an effective approximation method with a Monte Carlo simulation algorithm to tackle such a well-known open problem. Concerning a CNN, readers are also referred to Brizuela and Merchan [12], Dai [10], LeCun et al. [20], Vaswani et al. [27], and Yamashitza et al. [28] for more details. Note that the purpose of integrating the ML technique into our platform is to speed the convergence of our new algorithm designed in the current study (see Algorithm 3.1 and Figure 1 for more details).
To go further, we give some explanations about the notations used in the equation of (1.1). The notations Vx, Vxx, and ˉVx are the corresponding first-order and second-order partial derivatives of V and ˉV in terms of position parameter x∈R. Since the second partial derivative Vxx appears in the denominator of the second term on the right-hand side of (1.1), the equation in (1.1) is a strongly nonlinear one and exhibits singularity. In addition, since both V(t,x) and ˉV(t,x) are unknown random fields, this equation is a diophantine equation (see the related explanation in Dai [9]). Therefore, one major task in studying the equation in (1.1) is to try to obtain an adapted solution pair (V,ˉV) with respect to a filtration {Ft,t≥0} generated by the Brownian motion, i.e., Ft=σ(W(s),s≤t). Then, based on this paired solution, we can get its related further optimal financial investment strategy (see the discussions in Musiela and Zariphopoulou [22], ∅ksendal et al. [24], etc.).
Concerning the relationship between V and ˉV, we can interpret ˉV as a regulating process of V due to the martingale representation theorem (see, e.g., ∅ksendal [23]). Furthermore, under certain conditions, ˉV can be expressed as a functional of the Malliavin derivative of V (see Lemmas 4.6 and 4.7 in Dai [9] for more details). However, based on the Malliavin functional relationship between V and ˉV, it is difficult to design a direct computation algorithm to calculate (V,ˉV) numerically due to its complexity. Therefore, in Dai [9], this explicit relationship through Malliavin calculus is only used to prove the convergence of a more directly designed computation algorithm in solving (V,ˉV) numerically for B-SPDEs under the so-called generalized linear growth and Lipschitz conditions. Nevertheless, in the current study, our B-SPDE in (1.1) is a strongly nonlinear one that does not satisfies the generalized linear growth and Lipschitz conditions. Thus, our main focus in this paper is on the evolution of the study in Dai [9] to solve our current strongly nonlinear problem in terms of algorithm design, analysis, and implementation. The Malliavin functional relationship between V and ˉV will not be directly used in this study.
As introduced previously, to the best of our knowledge, the B-SPDE in (1.1) is still not well-solved. Hence, we here try to develop a numerical scheme with related theory to simulate this equation in an approximated way. More precisely, we consider an approximated analog of the equation in (1.1) as follows,
V(t,x)=H(T,x)−12∫Tt(ΦˉK(Vx(s,x)+ˉVx(s,x)))2Ψϵ,K(Vxx(s,x))ds−∫TtˉV(s,x)dW(s) | (1.4) |
for x∈D=[0,b] with b>0, where ΦˉK(⋅) is a truncation map corresponding to the first-order derivatives Vx(t,x) and ˉVx(t,x) (if any) for a small number ϵ>0 and a large number ˉK>0. In other words, for fx(t,x)=Vx(t,x)+ˉVx(t,x), we have that
ΦˉK(fx(t,x))≡{ˉKiffx(t,x)>ˉK,fx(t,x)if|fx(t,x)|≤ˉK,−ˉKiffx(t,x)<−ˉK, | (1.5) |
where |⋅| denotes the absolute value of a number ⋅. Furthermore, Ψϵ,K(⋅) is another truncation map corresponding to the second-order derivative Vxx(t,x) (if any) for a small number ϵ>0 and a large number K>0, i.e.,
Ψϵ,K(Vxx(t,x))≡{Vxx(t,x)ifϵ≤|Vxx(t,x)|≤K,ϵif0≤Vxx(t,x)<ϵ,−ϵif−ϵ<Vxx(t,x)<0,KifVxx(t,x)>K,−KifVxx(t,x)<−K. | (1.6) |
From (1.6), we can see that the absolute value of the truncation map Ψϵ,K(⋅) is always greater or equal to the positive number ϵ (i.e., |Ψϵ,K(⋅)|≥ϵ). Thus, the denominator appearing in (1.4) is always away from zero, which implies that the equation for the given constants ϵ>0, K>0, and ˉK>0 has the potential to be well-behaved. From (1.5) and (1.6), we can see that the truncation maps ΦˉK(⋅) and Ψϵ,K(⋅) are both bounded for the given constants ϵ>0, K>0, and ˉK>0. Furthermore, based on the truncated maps in (1.5) and (1.6) and the mentioned properties, we can prove that the equation in (1.4) satisfies the generalized local linear growth and generalized local Lipschitz conditions as given in Dai [9]. Hence, under a suitable terminal condition, there uniquely exists a {Ft}-adapted strong solution (Vϵ,KˉK(t,x),ˉVϵ,KˉK(t,x)) to the equation in (1.4), which implies that this equation in (1.4) is not a singular one for the given constants ϵ>0, K>0, and ˉK>0. Thus, according to the equation in (1.4), we can develop a Monte Carlo simulation algorithm by enhancing the one developed in Dai [9] through adding an additional machine learning (ML) loop (see Algorithm 3.1 in Section 3 for more details). Note that the design of using a sequence of structure-preserving B-SPDEs in (1.4) to approximate the one in (1.1) is actually motivated from the diffusion approximation for queueing networks where r is used to index a specific network. More exactly, the diffusion approximation is to approximate a target limit system through a sequence of physical systems when r tends to infinity (see, e.g., Dai [4,5,7]). Furthermore, during the approximation, each system keeps its structure unchanged.
In summary, the contributions of our paper can be stated in two folds: theoretic contributions and numerical contributions.
Concerning the theoretic contributions of our paper, we study the strongly nonlinear B-SPDE in (1.1) through an approximation method by introducing new truncation operators as in (1.5) and (1.6). The main reason for doing it in this way is due to the singularity of the B-SPDE in (1.1). To guarantee the meaningfulness of the approximation, the existence and uniqueness of a 2-tuple adapted strong solution to the approximation B-SPDE in (1.4) are proved. Furthermore, it is also proved that, if the B-SPDE in (1.1) has a solution pair (V,ˉV), the solution pair to the approximation B-SPDE in (1.4) will converge to (V,ˉV) as ϵ tends to zero and K,ˉK tend to the infinity in certain senses. In addition, based on the approximated B-SPDE in (1.4), some investment policies are analytically derived for a financial market with the aim to conduct some accuracy comparisons.
Concerning the numerical contributions of our paper, we develop a Monte Carlo simulation-based algorithm to numerically solve the approximation B-SPDE in (1.4). In doing so, we integrate an additional ML loop into the platform of a convolutional neural network (CNN) as studied in Dai [9]. Based on the equation in (1.4), we can prove the convergence of this newly designed algorithm in the case that the equation in (1.1) has an adapted strong solution. The basic idea to develop such an algorithm is to find a CNN denoted by U with an additional ML loop at each time point to approximate (Vϵ,KˉK,ˉVϵ,KˉK) (see Figure 1 with explanations in Subsection 3.1 for more details). In this CNN, we consider (Vϵ,KˉK,ˉVϵ,KˉK) as paired parameters to be trained at each node (t,x) (also called neuron (t,x)). The learning strategy for this CNN is to minimize the mean-squared error (loss) between (Vϵ,KˉK,ˉVϵ,KˉK) and U. Note that, in Dai [9], the CNN corresponding to U is expressed as a conditional expectation projection for a given data set. To compute the conditional expectation, the well-known tower law (see, e.g., Kallenberg [16]) is used to design the CNN U (see Subsection 3.1 for more details), which is a backward one. Similar to a multi-layer perceptron (MLP) as studied in Cybenko [3], Haykin [13], and Hornik et al. [14], our CNN U is a fully connected neural network. However, it is still computationally efficient since our B-SPDE in (1.4) is a Browinian motion driven one. Therefore, the computation of our conditional expectation does not need the additional multiple parameter training since the transition probability and density of a Brownian motion is given. As an alternative, our conditional expectation projection can also be presented by a linear or nonlinear regression model with parameters, which is closely related to the so-called Kolmogorov-Arnold network due to Kolmogorov's superposition theorem (see, e.g., Kolmogorov [17] and Braun and Griebel [1]) and the recent study in Liu et al. [21]. If we do in this way, we need to train additional parameters, which is time-consuming and may sacrifice some accuracy. Thus, in Dai [9] and in our current paper, we report our achievements based on our successfully implemented and tower-law-oriented CNN U. Furthermore, along this line, we refer the reader to these related papers Gonon et al. [11], Kratsios et al. [19], Peluchetti [25], Sirignano and Spiliopoulos [26], Vaswani et al. [27] for more details. Finally, concerning the numerical contributions of this study, we also provide simulation examples supported with an application in finance.
For the reader's convenience, the partial simulation results according to our developed simulation scheme are shown in Figure 2, where we present the simulation results concerning the paired solution (Vϵ,KˉK,ˉVϵ,KˉK) to the approximated B-SPDE in (1.4) with ϵ=1/100000, K=109, and ˉK=264. The computed values at time point tj0=tn0 with n0=10 correspond to the input terminal values at tn0+1=T. The "solution error check" titled in the third plot of the first column in Figure 2 is in terms of the difference between the two sides of the B-SPDE in (1.4) and is with respect to a particular sample path. From the simulation results displayed in the third plot of the first column, we can see that our algorithm is quite accurate. The three plots in the first row of Figure 2 display the simulated Vϵ,KˉK(tn0,x) together with its simulated first-order and second-order derivatives in terms of position parameter x∈{b/d,2b/d...,(d−dropnum)b/d} with b=1/6000, d=100, and dropnum=30. Although the graph in the third plot is non-smooth, it is close to a smooth line. Furthermore, the three plots in the second row of Figure 2 display the simulated ˉVϵ,KˉK(tn0,x) together with its simulated first-order and second-order derivatives. In addition, the second and third plots in the third row of Figure 2 display the simulated terminal value Vϵ,KˉK(tn0+1,x) together with its simulated first-order derivative. Concerning the simulations about the paired solution (Vϵ,KˉK(t,x),ˉVϵ,KˉK(t,x)), readers are referred to Section 4 for more details.
As pointed out previously, the primary interest in deriving the B-SPDE in (1.1) is to find the financial investment strategy together with myopic investment and excess hedging demand (see, e.g., Dai [6,8], Musiela and Zariphopoulou [22], and ∅ksendal et al. [24]). Therefore, based on the simulation study conducted in this paper and the related strategy study presented in Musiela and Zariphopoulou [22], we can also obtain the simulated financial control strategies as presented in Figure 3. In this figure, we display the simulated investment policy, myopic policy, and excess hedging demand at time point tj0 with j0=n0 and n0=10. These policies correspond to the formula derived in Musiela and Zariphopoulou [22]. The three graphs in the left column are corresponding to the simulated pathwise results. The three graphs in the right column correspond to the simulated results in the mean average sense with respect to the simulation iteration number Q. Theoretically, the myopic policy should continue to be constant. Our simulated results support this theoretical result. However, this theoretic result also further justifies the correctness of our algorithm and simulations. Concerning this financial policy simulation, readers are referred to Section 4 for more details.
The remainder of the paper is organized as follows. In Section 2, we present the unique existence theorem with proof of our approximated B-SPDE. Furthermore, in this section, we also present the application of our study in finance. In Section 3, we design our Monte Carlo simulation algorithm through CNN and ML. The convergence of our designed simulation algorithm is also proved. In Section 4, we present our numerical simulation case studies. Finally, in Section 5, we summarize our study conducted in this paper with conclusions.
This section consists of two subsections. The unique existence theorem is stated in Subsection 2.1. The financial application of our unique existence result is presented in Subsection 2.2.
Let C2(D,R) be the Banach space of all functions f having continuous derivatives up to the order 2 with the uniform norm,
‖f‖C2(D,R)=maxc∈{0,1,2}maxj∈{1,...,r(c)}supx∈D|f(c)j(x)| | (2.1) |
for each f∈C2(D,R). The r(c) in (2.1) for each c∈{0,1,2} is the total number of the partial derivatives of the cth order. Then, we can introduce our required measurable spaces used in this paper. First, we use L2F([0,T],C2(D;R)) to denote the set of all R-valued (also called C2(D;R)-valued) measurable random fields Z(t,x) satisfying
E[∫T0‖Z(t)‖2C2(D,R)dt]<∞, | (2.2) |
where the random field Z(t,x) is assumed to be adapted to {Ft,t∈[0,T]} for each x∈D and Z(t,x)∈C2(D,R) with a fixed t∈[0,T]). Second, we use L2F,1([0,T],C2(D,R)) to denote the corresponding set of predictable processes (see the definitions on pages 21 and 45 of Ikeda and Watanabe [15]). Third, we use L2FT(Ω,C2(D;R)) to denote the set of all R-valued, FT-measurable random fields ζ(x,ω) for each x∈D and sample point ω∈Ω, where ζ(x,ω)∈C2(D,R) for each ω∈Ω satisfies
‖ζ‖2L2FT(Ω,C2(D,R))≡E[‖ζ‖2C2(D,R)]<∞. | (2.3) |
Therefore, we can introduce our supporting space as follows,
ˉQ2F([0,T]×D)≡L2F([0,T],C2(D,R))×L2F,1([0,T],C2(D,R)). | (2.4) |
Finally, before introducing our unique existence theorem, we suppose that the terminal value in (1.1) (and hence in (1.4)) is given by
H(x)=h1(x)h2(W(T)), | (2.5) |
where both h1 and h2 are polynomials (interested readers are also referred to Dai [9] for more related explanation). Then, we can state our unique existence theorem as follows.
Theorem 2.1. Under the terminal condition in (2.5), there is a unique strong {Ft}-adapted solution pair (Vϵ,KˉK(t,x),ˉVϵ,KˉK(t,x)) to the B-SPDE in (1.4) within the space ˉQ2F([0,T]×D) for a small constant ϵ>0, a large number K>0, and a large number ˉK>0 with t∈[0,T]. Furthermore, we have that
supt∈[0,T]E[‖Vϵ,KˉK(t)‖2C2(D,R)]<∞. | (2.6) |
Proof. Corresponding to the B-SPDE in (1.4), we define a second-order partial differential operator Lϵ,KˉK as follows,
Lϵ,KˉK(t,x,V,ˉV)=(ΦˉK(Vx(t,x)+ˉVx(t,x)))2Ψϵ,K(Vxx(t,x)). | (2.7) |
Then, we can conclude that the operator Lϵ,KˉK defined in (2.7) satisfies the generalized local Lipschitz and linear growth conditions as introduced in Dai [9], i.e.,
|ΔLϵ,KˉK(s,x,(u,ˉu),(v,ˉv))|≤KD(‖u−v‖C2(D,R)+‖ˉu−ˉv‖C2(D,R)), | (2.8) |
|Lϵ,KˉK(s,x,(u,ˉu))|≤KD(‖u‖C2(D,R)+‖ˉu‖C2(D,R)) | (2.9) |
for each fixed (t,x)∈[0,T]×D, and any (u,ˉu), (v,ˉv) ∈C2(D,R)×C2(D,R), where (u,ˉu) and (v,ˉv) are two pairs corresponding to the equation in (1.4). Furthermore, KD is a nonnegative constant depending on D, ϵ, K, and ˉK. In addition, the operator ΔLϵ,KˉK is defined by
ΔLϵ,KˉK(s,x,(u,ˉu),(v,ˉv))≡Lϵ,KˉK(s,x,(u,ˉu))−Lϵ,KˉK(s,x,(v,ˉv)) | (2.10) |
for each given (t,x,(u,ˉu),(v,ˉv))).
In fact, for any two pairs of (U,ˉU) and (V,ˉV) in the space ˉQ2F([0,T]×D), it follows from the definition in (2.7) that we can prove the claim in (2.8) as follows,
|ΔLϵ,KˉK(s,x,(U,ˉU),(V,ˉV))|≤|(ΦˉK(Ux(s,x)+ˉUx(s,x)))2Ψϵ,K(Uxx(s,x))−(ΦˉK(Vx(s,x)+ˉVx(s,x)))2Ψϵ,K(Vxx(s,x))|=|Ψϵ,K(Vxx(s,x))(ΦˉK(Ux(s,x)+ˉUx(s,x)))2Ψϵ,K(Uxx(s,x))Ψϵ,K(Vxx(s,x))−Ψϵ,K(Uxx(s,x))(ΦˉK(Vx(s,x)+ˉVx(s,x)))2Ψϵ,K(Uxx(s,x))Ψϵ,K(Vxx(s,x))|≤1ϵ2|Ψϵ,K(Vxx(s,x))((ΦˉK(Ux(s,x)+ˉUx(s,x)))2−(ΦˉK(Vx(s,x)+ˉVx(s,x)))2)|+1ϵ2|(Ψϵ,K(Uxx(s,x))−Ψϵ,K(Vxx(s,x)))(ΦˉK(Vx(s,x)+ˉVx(s,x)))2|≤K1D(|Ψϵ,K(Uxx(s,x))−Ψϵ,K(Vxx(s,x))|+|ΦˉK(Ux(s,x)+ˉUx(s,x))−ΦˉK(Vx(s,x)+ˉVx(s,x))|)≤KD(‖u−v‖C2(D,R)+‖ˉu−ˉv‖C2(D,R)), | (2.11) |
where K1D is some positive constant depending on ϵ, K, ˉK, and D. Furthermore, KD is a positive constant depending on K1D.
Similarly, by applying the definition in (2.7), we can prove the claim in (2.9) as follows,
|LϵK(s,x,(U,ˉU))|≤1ϵ(ΦˉK(Ux(t,x)+ˉUx(t,x)))2≤KD(‖u‖C2(D,R)+‖ˉu‖C2(D,R)), |
where the constant KD in (2.11) and (2.12) can be chosen such that the inequalities in (2.11) and (2.12) are both true.
Finally, it follows from (2.11), (2.12), and the discussion in Dai [9] that the claims in our main theorem are true. Hence, we complete the proof of Theorem 2.1.
Consider a financial market consisting of one risky asset and one riskless asset. The risky asset is a stock whose price dynamics is driven by the Brownian motion W(⋅), i.e.,
dS(t)=S(t)(u(t)dt+σ(t)dW(t)) | (2.12) |
with initial price S(0)>0. Furthermore, u(⋅) and σ(⋅) are {Ft}-progressive measurable processes with values in R=(−∞,∞). The riskless asset is with the price process R(t) with an interest rate r(t), i.e.,
dR(t)=r(t)R(t)dt. | (2.13) |
Then, it follows from the discussion in Musiela and Zariphopoulou [22] that the present value Xβ(t)=β0(t)+β1(t) of the aggregate investment concerning the riskless investment strategy β0(t) and the risky investment strategy β1(t) is given by
dXβ(t)=β(t)(u(t)−r(t))dt+σ(t)β(t)dW(t), | (2.14) |
where β(t)=β1(t) is the discounted strategy in the following admissibility set,
A={β:β(t)is self-financing and{Ft}-progressively measurablesatisfyingE[∫t0|σ(s)β(s)|2]<∞,Xβ(t)≥0,t≥0}. | (2.15) |
Now, for a given trading horizon [0,T] and an investor's utility μT(⋅):R+→R at terminal time T, which is supposed to be an increasing and convex function of his wealth, we can represent the risk-seeking attitude of an investor (the risk-avoiding case corresponding to a concave function can be similarly discussed, see, e.g., Musiela and Zariphopoulou [22] for a reference). The investor's target is to maximize the expected utility of terminal wealth over the admissible strategy set AT corresponding to the one in (2.15) with t∈[0,T], i.e., to solve the optimization problem presented in (1.2). Then, if we take u(t)=σ(t)≡1 and r(t)≡0, respectively, in (2.12) and (2.13), the optimal equation corresponding to the optimization problem in (1.2) is given by the B-SPDE in (1.1) with H(T,x)=μT(x) (concerning the justification of this claim, readers are referred to Musiela and Zariphopoulou [22] for more details). In this case, the optimal feedback investment strategy is given by
β∗(t)=−Vx(t,x)Vxx(t,x)−ˉVx(t,x)Vxx(t,x), | (2.16) |
where the first term (will be denoted by β∗,m(t)) on the right-hand-side of (2.16) is called a myopic investment strategy resembling the investment policy followed by an investor in a financial market where the investment opportunity set keeps constant through time. Furthermore, the second term (will be denoted by β∗,h(t)) on the right-hand-side of (2.16) is referred as the excess hedging demand denoting the additional investment due to the volatility ˉV(t,x) of the performance process V(t,x). Thus, corresponding to the paired solution (Vϵ,KˉK(t,x),ˉVϵ,KˉK(t,x)) to the equation in (1.4), the approximated optimal investment strategies can be denoted by β∗ϵ(t), β∗,mϵ(t), and β∗,hϵ(t) (see, e.g., the simulation results in Figures 3–5).
This section consists of two subsections concerning the design of a Monte Carlo simulation algorithm and proving its convergence with error bound estimation.
In this subsection, we develop a Monte Carlo simulation algorithm based on both CNN and machine learning (see Figure 1) to simulate the 2-tuple adapted strong solution to the B-SPDE in (1.4). More precisely, we consider a partition π for the product region of [0,T]×D with D=[0,b] as follows,
π:0=t0<t1<⋅⋅⋅<tn0=Twithn0∈{0,1,...},0=x0<x1<⋅⋅⋅<xn1=bwithn1∈{0,1,...}, | (3.1) |
where tj0 for j0∈{0,1,2,...,n0} and xj1 for j1∈{0,1,...,n1} are the points of divisions over the time interval [0,T] and the space interval D. Then, for all jl∈{1,...,nl} with l∈{0,1}, we take
Δt,πj0=tj0−tj0−1, | (3.2) |
Δπ1=xj1−xj1−1=bn1, | (3.3) |
ΔπWj0=W(tj0)−W(tj0−1). | (3.4) |
Furthermore, let
|π|≡maxj0∈{1,...,n0}{Δt,πj0,Δπ1}, | (3.5) |
Dj1≡[xj1−1,xj1), | (3.6) |
X≡{xj1:j1∈{0,1,...,n1}}. | (3.7) |
Now, we use the forward and the backward difference techniques to approximate the partial derivatives appearing in (1.4). More precisely, for each f∈{Vϵ,KˉK,ˉVϵ,KˉK}, x∈X, and each integer c∈{1,2}, we define the cth-quotient of differences, which corresponds to the cth-order derivative of f along the x direction, as follows,
f(c)π(t,x)≡{f(c−1)π(t,x+Δπ1)−f(c−1)π(t,x)Δπ1ifx=xj1andj1<n1,−f(c−1)π(t,x−Δπ1)−f(c−1)π(t,x)Δπ1ifx=xj1andj1=n1, | (3.8) |
where we adopt the convention that f(0)π=fπ. Furthermore, to simplify the notations, we use {Vϵ,ˉVϵ} to denote {Vϵ,KˉK,ˉVϵ,KˉK}, and we define
Lϵ,KˉK(t,x,Vϵπ(t,x))≡Lϵ,KˉK(t,x,(Vϵπ(t,x),Vϵ,(1)π(t,x),Vϵ,(2)π(t,x)),(ˉVϵπ(t,x),ˉVϵ,(1)π(t,x))) | (3.9) |
for each x∈X. Moreover, we use Lϵ,Kπ,ˉK to denote the fully discretized version of Lϵ,KˉK. Then, we can present the following Monte Carlo simulation algorithm.
Algorithm 3.1. This algorithm consists of three parts: Part I, Part II, and Part III:
Part I. This part is an iterative one in terms of {(Vϵ(tj0,x),ˉVϵ(tj0,x)): x∈X} with j0 decreasing from n0 to 1 in a backward way,
Vϵπ(tn0,x)=Hπ(x), | (3.10) |
ˉVϵπ(tn0,x)=0, | (3.11) |
Vϵπ(tj0−1,x)=E[Vϵπ(tj0,x)+Lϵ,Kπ,ˉK(tj0,x,Vϵπ(tj0,x))Δt,πj0|Ftj0−1], | (3.12) |
ˉVϵπ(tj0−1,x)=1Δπj0E[Vϵπ(tj0,x)ΔπWj0|Ftj0−1]+E[Lϵ,Kπ,ˉK(tj0,x,Vϵπ(tj0,x))ΔπWj0|Ftj0−1]. | (3.13) |
Part II. This part is a machine learning loop at time tj0−1 in order to minimize the difference concerning the values on both sides of the equation in (1.4) along each sample path, i.e.,
Vϵ,k+1π(tj0−1,x)=Vϵ,kπ(tj0−1,x)−α∇G(Vϵ,kπ(tj0−1,x)) | (3.14) |
for each k∈{0,1,2,...} with Vϵ,0π(tj0−1,x)=Vϵπ(tj0−1,x), where α is a given learning rate and ∇G(⋅) is the stochastic gradient of an optimization problem, i.e.,
minVϵ,kπ(tj0−1,x)∈RG(Vϵ,kπ(tj0−1,x)) |
with its objective function G(Vϵ,kπ(tj0−1,x)) given by
(Vϵ,kπ(tj0−1,x)−(Vϵπ(tj0,x)−12Lϵ,Kπ,ˉK(tj0−1,x,Vϵ,kπ(tj0−1,x))Δt,πj0)+ˉVϵπ(tj0−1,x)ΔπWj0)2. |
The machine learning loop in this part has a stopping rule as follows.
Choose a number (k+1) to stop the iteration,then take the new Vϵπ(tj0−1,x) to be Vϵ,k+1π(tj0−1,x). | (3.15) |
Part III. This part is to compute the numerical derivatives at each time tj0−1, i.e.,
Compute Vϵ,(c)π(tj0−1,x) and ˉVϵ,(c)π(tj0−1,x)for each x∈X with c∈{1,2} via the formula in (3.8). | (3.16) |
Concerning the architecture, the loss, and the learning strategy corresponding to Algorithm 3.1, the associated flow chart is shown in Figure 1. In this flow chart, we present a backward CNN supported with a ML loop.
As displayed in the graph on the left-hand side of Figure 1, this CNN denoted by U has (n0+1)-layers arranged in a backward way, which corresponds to Part I of Algorithm 3.1. In this CNN, a node (also called a neuron) is indexed by (tj,xi) with j∈{n0,n0−1,...,1,0} and i∈{0,1,...,n1}, where, in our Algorithm 3.1, we take j=j0 and i=j1. Associated with each node (tj,xi), the paired solution (Vϵ(tj,xi),ˉVϵ(tj,xi)) is considered as a pair of parameters to be trained or estimated. The training process is arranged in a backward way as follows,
(Vϵ(tn0,xi),ˉVϵ(tn0,xi))→(Vϵ(tn0−1,xi),ˉVϵ(tn0−1,xi))→⋅⋅⋅→(Vϵ(t0,xi),ˉVϵ(t0,xi)) |
for each i∈{0,1,...,n1}. The design rational for this CNN is the tower law for expectation (see, e.g., Theorem 5.1 (vii) on page 81 of Kallenberg [16]) and more explanations concerning the design rational for a CNN can be found in Dai [9]. The learning strategy for this CNN is to minimize the mean-squared error (loss) between (Vϵ,ˉVϵ) and U.
As displayed in the graph on the right-hand side of Figure 1, the supporting ML loop corresponds to Part II of Algorithm 3.1. The purpose of adding this ML loop into Algorithm 3.1 is to speed up the convergence of the algorithm in Dai [9]. Actually, in Dai [9], the algorithm is designed only through the backward CNN corresponding to Part I of Algorithm 3.1. Interestingly, our numerical implementations presented in this paper indicate that the convergence of Algorithm 3.1 is indeed faster after adding this ML loop. Concerning the supporting ML loop, the iteration is only designed for Vϵπ as shown in Part II of Algorithm 3.1. In this iteration, we keep ˉVϵπ the same for all k. In this ML loop, the learning strategy corresponds to solve an optimization problem as presented in Part II of Algorithm 3.1, which is to minimize the difference (loss) of the two sides of the B-SPDE in (1.4) along each sample path in the squared error sense. In the learning iteration of (3.14), the learning rate α is designed to satisfy the condition αK1D<1 for a constant K1D (that is presented in (3.22) of this paper). It is worthwhile to point out that, even after we add this ML loop, we still can prove the convergence of this newly designed Algorithm 3.1, which is presented in the next subsubsection.
Finally, the associated simulation examples based on Algorithm 3.1 will be provided in Section 4. Instead, in the next subsection, we first conduct the convergence analysis with error bound estimation.
In this subsection, we focus on the discussion concerning the convergence and error bound estimation of Algorithm 3.1. More precisely, for each t∈[0,T], x∈X, and j0∈{n0,n0−1,...,1}, we define
ΔVϵ,KˉK(t,x)=Vϵ,KˉK(t,x)−Vϵπ(t,x), | (3.17) |
ΔˉVϵ,KˉK(t,x)=ˉVϵ,KˉK(t,x)−ˉVϵπ(t,x),Vϵπ(t,x)=Vϵπ(tj0−1,x),t∈[tj0−1,tj0),ˉVϵπ(t,x)=ˉVϵπ(tj0−1,x),t∈[tj0−1,tj0). | (3.18) |
Furthermore, let
ξ(k)=k∑i0=11+k∑i0=1i0∑i1=11+k∑i0=1i0∑i1=1i1∑i2=11+⋅⋅⋅+k∑i0=1i0∑i1=1i1∑i2=1⋅⋅⋅2∑im−1=11. | (3.19) |
In addition, it follows from the generalized local Lipschitz and linear growth conditions in (2.8) and (2.9) that
|Δ(∇G)(s,x,u,v)|≤K1D‖u−v‖C2(D,R), | (3.20) |
|∇G(s,x,u))|≤K1D‖u‖C2(D,R) | (3.21) |
for some positive constant K1D, where (t,x)∈[0,T]×D and u,v∈C2(D,R)×C2(D,R). Therefore, we can introduce the following assumption concerning the learning rate α,
αK1D<1. | (3.22) |
Finally, let ‖⋅‖ be the largest absolute value corresponding to each used function for all x∈X. Then, we can present our algorithm convergence theorem with error bound estimation as follows.
Theorem 3.1. For Algorithm 3.1 with the condition in (3.22) and a given iteration number k in (3.14), there is a nonnegative constant C depending only on the terminal time T, the Lipschitz constant KD in (2.8) and (2.9), and the supremum (a constant) in (2.6) such that the following mean-square error estimation is true,
supt∈[0,T](E[‖ΔVϵ,KˉK(t)‖2]+E[‖ΔˉVϵ,KˉK(t)‖2])≤C|π| | (3.23) |
for all sufficiently small |π|. In addition, consider a sequence of increasing sets Xˉk with ˉk∈{1,2,...} (i.e., X1⊂X2⊂⋅⋅⋅). Suppose that the corresponding maximal mesh gauge |π|ˉk along ˉk∈{1,2,...} satisfies
|π|ˉk→0,∞∑ˉk=1(|π|ˉk)13<∞. |
Then, for a given X∈{X1,X2,...}, we have the a.s. convergence for Algorithm 3.1 as ˉk→∞,
supt∈[0,T](‖ΔVϵ,KˉK(t)‖+‖ΔˉVϵ,KˉK(t)‖)→0a.s. | (3.24) |
Proof. First, for the machine learning loop given by (3.14) in Part II of Algorithm 3.1, a given integer k∈{0,1,2,...}, and an index j0∈{n0,n0−1,...,1}, it follows from the facts in (3.20) and (3.21) that
‖Vϵ,k+1π(tj0−1)−Vϵ,0π(tj0−1)‖≤αk∑i0=1‖∇G(Vϵ,i0π(tj0−1))‖≤αk∑i0=1(i0∑i1=1‖∇G(Vϵ,i1π(tj0−1))−∇G(Vϵ,i1−1π(tj0−1))‖+‖∇G(Vϵ,0π(tj0−1))‖)≤α|π|K1Dk∑i0=1(i0∑i1=1‖Vϵ,i1π(tj0−1)−Vϵ,i1−1π(tj0−1)‖C2(X,R)+‖Vϵ,0(tj0−1)‖C2(X,R))≤α|π|K1Dk∑i0=1(i0∑i1=1α|π|K1D(i1∑i2=1‖Vϵ,i2π(tj0−1)−Vϵ,i2−1π(tj0−1)‖C2(X,R)+‖Vϵ,0π(tj0−1)‖C2(X,R))+‖Vϵ,0π(tj0−1)‖C2(X,R))⋅⋅⋅⋅⋅⋅⋅⋅⋅≤α|π|K1Dξ(k)‖Vϵ,0π(tj0−1)‖C2(X,R), | (3.25) |
where ξ(k) is given in (3.19). Thus, for the given iteration number k and t∈[tj0−1,tj0), it follows from (3.25) and the discussion in Dai [9] that
E[‖ΔVϵ,KˉK(t)‖2]=E[‖Vϵ,KˉK(t)−Vϵ,k+1π(t)‖2]≤2E[‖Vϵ,KˉK(t,x)−Vϵ,0π(t)‖2]+2E[‖Vϵ,0π(t)−Vϵ,k+1π(t)‖2]≤C1|π|+(α|π|K1Dξ(k))2E[‖Vϵ,0π(tj0−1)‖2C2(X,R)]≤C1|π|+2(α|π|K1Dξ(k))2(E[‖Vϵ,0π(t)−Vϵ,KˉK(t)‖2C2(X,R)]+E[‖Vϵ,KˉK(t)‖2C2(X,R)])≤C1|π|+2(α|π|K1Dξ(k))2(C1|π|+supt∈[0,T]E[‖Vϵ,KˉK(t)‖2C2(D,R)]), | (3.26) |
where C1 is some constant depending on T and KD. Since Vϵ,0π(tj0−1,x) for each j0∈{n0,n0−1,...,1} satisfies the equation in (3.12), it follows from Theorem 2.1 and (3.26) that the claim in (3.23) is true. Furthermore, the claim in (3.24) can be similarly proved by applying the discussion in Dai [9]. Hence, we finish the proof of Theorem 3.1.
Now, if the B-SPDE in (1.1) has a {Ft}-adapted solution pair (V(t,x),ˉV(t,x)), then Vxx(t,x) can not be zero a.s. at all points {(t,x)∈[0,T]×D}. Therefore, the numerical procedure in Algorithm 3.1 can also be applied to solve the equation in (1.1), e.g.,
Vπ(tn0,x)=Hπ(x), | (3.27) |
ˉVπ(tn0,x)=0, | (3.28) |
Vπ(tj0−1,x)=E[Vπ(tj0,x)+Lπ(tj0,x,Vπ(tj0,x))Δt,πj0|Ftj0−1], | (3.29) |
ˉVπ(tj0−1,x)=1Δπj0E[Vπ(tj0,x)ΔπWj0|Ftj0−1]+E[Lπ(tj0,x,Vπ(tj0,x))ΔπWj0|Ftj0−1], | (3.30) |
where Lπ in (3.29) and (3.30) is the discrete version of the following partial differential operator,
L(t,x,V,ˉV)=(Vx(t,x)+ˉVx(t,x))2Vxx(t,x). | (3.31) |
Then, we have the following corollary.
Corollary 3.2. Under the conditions in (3.22) with the Lipschitz constant KD in (2.8)−(2.9), if there is a pair of {Ft}-adapted solutions (V(t,x),ˉV(t,x)) in the space ˉQ2F([0,T]×D) to the B-SPDE in (1.1), then we have that
(Vϵπ(tj0,x),ˉVϵπ(tj0,x))→(Vπ(tj0,x),ˉVπ(tj0,x))a.s. | (3.32) |
E[|(Vϵπ(tj0,x),ˉVϵπ(tj0,x))−(Vπ(tj0,x),ˉVπ(tj0,x))|]→0 | (3.33) |
as ˉK→∞ first, K→∞ second, ϵ→0 third for each x∈X and each j0∈{n0,n0−1,...,1} in a backward way, and |π|ˉk→0 last along ˉk∈{1,2,...}.
Proof. We prove the claim in (3.32) and (3.33) by induction in terms of j0∈{n0,n0−1,...,2,1}. First, we consider the case that j0=n0−1, and it follows from (3.12) and (3.29) that
|Vπ(tn0−1,x)−Vϵπ(tn0−1,x)|≤E[((Vx(tn0,x))2|Vxx(tn0,x)|+ˉK2|Vxx(tn0,x)|)I{ϵ≤|Vxx(tn0,x)|≤K}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+ˉK2ϵ)I{|Vxx(tn0,x)|<ϵ}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+ˉK2K)I{|Vxx(tn0,x)|>K}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2ϵ)I{|Vxx(tn0,x)|<ϵ}I{|Vx(tn0,x)|≤ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2K)I{|Vxx(tn0,x)|>K}I{|Vx(tn0,x)|≤ˉK}Δt,πn0|Ftn0−1]≤2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{ϵ≤|Vxx(tn0,x)|≤K}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{|Vxx(tn0,x)|<ϵ}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2K)I{|Vxx(tn0,x)|>K}I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{|Vxx(tn0,x)|<ϵ}I{|Vx(tn0,x)|≤ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2K)I{|Vxx(tn0,x)|>K}I{|Vx(tn0,x)|≤ˉK}Δt,πn0|Ftn0−1]≤2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2K)I{|Vx(tn0,x)|>ˉK}Δt,πn0|Ftn0−1]+2E[(Vx(tn0,x))2|Vxx(tn0,x)|I{|Vxx(tn0,x)|<ϵ}Δt,πn0|Ftn0−1]+E[((Vx(tn0,x))2|Vxx(tn0,x)|+(Vx(tn0,x))2K)I{|Vxx(tn0,x)|>K}Δt,πn0|Ftn0−1]. | (3.34) |
Then, by the monotone (and conditional monotone) convergence theorem, we first let ˉK→∞, second let K→∞, and third let ϵ→0, and we can conclude that
|Vϵπ(tn0−1,x)−Vπ(tn0−1,x)|→0a.s., | (3.35) |
E[|Vϵπ(tn0−1,x)−Vπ(tn0−1,x)|]→0. | (3.36) |
Second, we consider the case that j0=n0−2. Let Vϵ,Kπ,ˉK(tn0−1,x) be the corresponding value computed through (3.29). Then, it follows from (3.12) and (3.29) that
|Vπ(tn0−2,x)−Vϵπ(tn0−2,x)|≤E[|ΔVϵ,KˉK(tn0−1,x)||Ftn0−2]+E[|Vπ(tn0−1,x)−Vϵ,Kπ,ˉK(tn0−1,x)||Ftn0−2]+E[((Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|+ˉK2|Vxx(tn0−1,x)|)I{ϵ≤|Vxx(tn0−1,x)|≤K}I{|Vx(tn0−1,x)+ˉVx(tn0−1,x)|>ˉK}Δt,πn0−1|Ftn0−2]+E[((Vx(tn0,x)+ˉVx(tn0−1,x))2|Vxx(tn0,x)|+ˉK2ϵ)I{|Vxx(tn0,x)|<ϵ}I{|Vx(tn0−1,x)+ˉVx(tn0−1,x)|>ˉK}Δt,πj0|Ftn0−2]+E[((Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|+ˉK2K)I{|Vxx(tn0−1,x)|>K}I{|Vx(tn0−1,x)+ˉVx(tn0−1,x)|>ˉK}Δt,πn0−1|Ftn0−2]+E[((Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|+(Vx(tn0−1,x)+ˉVx(tn0−1,x))2ϵ)I{|Vxx(tn0−1,x)|<ϵ}I{|Vx(tn0−1,x)+ˉVx(tn0−1,x)|≤ˉK}Δt,πn0−1|Ftn0−2]+E[((Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|+(Vx(tn0−1,x)+ˉVx(tn0−1,x))2K)I{|Vxx(tn0−1,x)|>K}I{|Vx(tn0−1,x)+ˉVx(tn0−1,x)|≤ˉK}Δt,πn0−1|Ftn0−2]≤E[|ΔVϵ,KˉK(tn0−1,x)||Ftn0−2]+E[|Vπ(tn0−1,x)−Vϵ,Kπ,ˉK(tn0−1,x)||Ftn0−2]+4E[(Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|I{|Vx(tn0−1,x)|+ˉVx(tn0−1,x)>ˉK}Δt,πn0−1|Ftn0−2]+2E[(Vx(tn0−1,x)+ˉVx(tn0−1,x))2(1|Vxx(tn0−1,x)|+1K)I{|Vxx(tn0−1,x)|>K}Δt,πn0−1|Ftn0−2]+2E[(Vx(tn0−1,x)+ˉVx(tn0−1,x))2|Vxx(tn0−1,x)|I{|Vxx(tn0−1,x)|<ϵ}Δt,πn0−1|Ftn0−2]. | (3.37) |
Then, by the monotone convergence theorem, we first let ˉK→∞, second let K→∞, and third let ϵ→0, and it follows from (3.35) and (3.36) that
|Vϵπ(tn0−2,x)−Vπ(tn0−2,x)| | (3.38) |
→E[|ΔVϵ,KˉK(tn0−1,x)||Ftn0−2]+E[|Vπ(tn0−1,x)−Vϵ,Kπ,ˉK(tn0−1,x)||Ftn0−2]a.s.,E[|Vϵπ(tn0−2,x)−Vπ(tn0−2,x)|]→E[|ΔVϵ,KˉK(tn0−1,x)|]+E[|Vπ(tn0−1,x)−Vϵ,Kπ,ˉK(tn0−1,x)|]. | (3.39) |
Finally, a similar argument can be applied to the analysis for (ˉVπ(tn0−1,x)−ˉVϵπ(tn0−1,x)). Then, it follows from the induction, dominated convergence theorem, and Theorem 3.1 that the claims in (3.32) and (3.33). Hence, we have finished the proof of Corollary 3.2.
In this section, we present simulation examples to show the effectiveness of Algorithm 3.1. More precisely, we first simulate the paired solution to the approximated B-SPDE in (1.4). Since there is no analytic solution available to the B-SPDE, we conduct numerical comparison concerning the difference between two sides of our simulated B-SPDE to show the correctness of our computed solution. All of the results concerning the simulated solution are presented in Figures 2, 6, and 8. Second, based on the simulated solution of the B-SPDE, we also conduct further simulations concerning the financial investment strategies derived in (2.16). All of the simulated results are displayed in Figures 3–5. Note that some predicted properties in theory concerning the financial investment strategies are found in the simulation results, which further justifies the correctness of our simulated solution to the B-SPDE in (1.4).
In all of the simulations, we use the notation T to denote the terminal time, a positive integer n to denote the number of equally divided subintervals over [0,T], a positive integer hhh to represent the highest order of partial derivatives corresponding to the implemented equation, a positive integer Q to be the total number of normally distributed random numbers, a positive number b to be the size for the space parameter, and a positive integer d to be the number of equally divided subintervals over [0,b]. Furthermore, we use a positive number BMD to denote the upper bound of an interval such that our driving Brownian motion W∈[−BMD,BMD] and use a positive integer bmdp to denote the number of equally divided subintervals over [0,BMD]. In addition, we use an integer (n0+1) to denote the number of layers of backward network as in the left part of Figure 1 and use an integer (k+1) to denote the number of layers of reinforcement iterations as in the right part of Figure 1.
Note that the main purpose to impose the upper bounds K and ˉK in (1.4) is for the equation in (1.4) to satisfy the general local Lipschitz and linear growth conditions. In our simulations, we take them to be sufficiently large in order that our simulation results are stable and acceptable. Actually, we take ˉK to be the largest number allowed by our used computer, i.e., ˉK=264. The computed first-order derivatives of (Vϵ,KˉK,ˉVϵ,KˉK) are much less than this upper bound ˉK. In this sense, our simulations are stable and acceptable. Concerning K, we have used different numbers for tests. Our simulation results presented in this paper correspond to K=109. If we used K=107, the simulation results were not affected by this change. For safety, we used the simulation results corresponding to K=109 in this paper.
However, since we are handling a B-SPDE in (1.1) with singularity, the choice of ϵ indeed has a significant impact on our simulation results. After careful tests, we chose to present the simulation results corresponding to ϵ=1/100000 in this paper. If we chose ϵ>1/100000 up to ϵ=1/100, the corresponding simulation results were still reasonable and acceptable. Nevertheless, if we chose ϵ<1/100000 significantly, our simulation results were not acceptable. Concerning the choice of the learning rate α, we chose α=1/2 for good simulation results. The simulation results corresponding to α around 1/2 were also acceptable. Nevertheless, if we chose α=1, the simulation results are not acceptable.
In Figure 2, we present the simulation results of the paired solution (Vϵ,KˉK,ˉVϵ,KˉK) to the approximated B-SPDE in (1.4) with ϵ=1/100000, K=109, and ˉK=264. The computed values at time point tj0=tn0 with n0=10 correspond to the input terminal values at tn0+1=T. The terminal random field H(x) in (1.4) is taken to be the form
H(x)=C(1+W2(T))(x+5.560000)2, | (4.1) |
where C is a positive constant and is taken to be terminalcoefficient=2∗100 as in the explanation for Figure 2. Note that the "solution error check" titled in the third plot of the first column in Figure 2 is in terms of the difference (denoted by "Err") between the two sides of the B-SPDE in (1.4) and is with respect to a particular sample path. In applying Algorithm 3.1 with the terminal time at tn0+1=T to solve the equation in (1.4), the corresponding Err at time tn0 has the following expression,
Err(tn0,x)=V(tn0,x)−H(x)+12(T−tn0)ΦˉK(Vx(tn0,x)+ˉVx(tn0,x))Ψϵ,K(Vxx(tn0,x)+(W(T)−W(tn0))ˉV(tn0,x). |
From the simulation results displayed in the third plot of the first column, we can see that our algorithm is quite accurate. Furthermore, the three plots in the first row of Figure 2 display the simulated Vϵ,KˉK(tn0,x) together with its simulated first-order and second-order derivatives in terms of position parameter x∈{b/d,2b/d...,(d−dropnum)b/d} with d=100 and dropnum=30. Although the graph in the third plot is non-smooth, it is close to a smooth line. In addition, the three plots in the second row of Figure 2 display the simulated ˉVϵ,KˉK(tn0,x) together with its simulated first-order and second-order derivatives. Finally, the second and third plots in the third row of Figure 2 display the simulated terminal value Vϵ,KˉK(tn0+1,x) together with its simulated first-order derivative. In Figure 6, we display the similar plots as in Figure 2 but with the simulated results at time point tn0−8=t3.
In Figure 7, we display the first case of the solution evolving as the time index j0 decreases from n0+1 to 3 with n0=10. More precisely, the three graphs in the magenta color display the dynamic evolving of Vϵ,KˉK(tj0,x) together with their first-order and second-order derivatives. The graph in the blue color displays the solution evolving of ˉVϵ,KˉK(tj0,x). Similarly, in Figure 8, we display the second case of the solution evolving as the time index j0 decreases from n0+1 to 3 with n0=10. More precisely, the three graphs in the blue color display the dynamic evolving of ˉVϵ,KˉK(tj0,x) together with their first-order and second-order derivatives. The graph in the magenta color displays the solution evolving of Vϵ,KˉK(tj0,x).
In Figure 3, we display the simulated investment policy, myopic policy, and excess hedging demand at time point tj0 with j0=n0 and n0=10. These policies correspond to the formula in (2.16) and its related explanations. The three graphs in the left column correspond to the simulated pathwise results. The three graphs in the right column correspond to the simulated results in the mean average sense with respect to the simulation iteration number Q. Theoretically, the myopic policy should continue to be constant. Our simulated results support this theoretical result. However, this theoretic result also further justifies the correctness of our algorithm and simulations. Similarly, in Figure 4, we display the simulated investment policy, myopic policy, and excess hedging demand at time point tn0−8=t3 with n0=10. Finally, in Figure 5, we display the simulated dynamical evolutions of investment policy, myopic policy, and excess hedging demand with respect to time parameter t∈{T,T(n−1)/n,...,T(n−8)/n}.
In this paper, we studied a strongly nonlinear B-SPDE through an approximation method. This equation is well-known and was previously derived from studies in finance. However, how to analyze and solve this equation has continued to an open problem for quite a long time. Therefore, by applying our previously established theory and numerical scheme together with CNN and ML, we have developed an effective approximation method with a Monte Carlo simulation algorithm to tackle the well-known open problem. In doing so, the existence and uniqueness of the 2-tuple adapted strong solution to an approximation B-SPDE were proved. Meanwhile, the convergence of a newly designed simulation algorithm was established. Simulation examples and applications in finance were also provided.
The author declares he has not used Artificial Intelligence (AI) tools in the creation of this article.
The author acknowledges that the project was funded by the National Natural Science Foundation of China with Grant No. 11771006. The author would like to thank the editors and the reviewers for their helpful comments and suggestions to revise this paper.
The author declares that he has no competing interests.
[1] |
J. Braun, M. Griebel, On a constructive proof of Kolmogorov's superposition thoerem, Constr. Approx., 35 (2009), 653–675. https://doi.org/10.1007/s00365-009-9054-2 doi: 10.1007/s00365-009-9054-2
![]() |
[2] |
A. Cĕrný, J. Kallsen. On the structure of general mean-variance hedging strategies, Ann. Appl. Probab., 35 (2007), 1479–1531. https://doi.org/10.1214/009117906000000872 doi: 10.1214/009117906000000872
![]() |
[3] |
G. Cybenko, Approximation by superpositions of a sigmoidal function, Math. Control Signal System, 1 (1989), 303–314. https://doi.org/10.1007/BF02551274 doi: 10.1007/BF02551274
![]() |
[4] | W. Dai, Brownian approximations for queueing networks with finite buffers: modeling, heavy traffic analysis and numerical implementations, Ph.D thesis, Georgia Institute of Technology, 1996. |
[5] |
J. G. Dai, W. Dai, A heavy traffic limit theorem for a class of open queueing networks with finite buffers, Queueing Syst., 32 (1999), 5–40. https://doi.org/10.1023/A:1019178802391 doi: 10.1023/A:1019178802391
![]() |
[6] |
W. Dai, Mean-variance portfolio selection based on a generalized BNS stochastic volatility model, Int. J. Comput. Math., 88 (2011), 3521–3534. https://doi.org/10.1080/00207160.2011.606904 doi: 10.1080/00207160.2011.606904
![]() |
[7] |
W. Dai, Optimal rate scheduling via utility-maximization for J-user MIMO Markov fading wireless channels with cooperation, Oper. Res., 61 (2013), 1450–1462. https://doi.org/10.1287/opre.2013.1224 doi: 10.1287/opre.2013.1224
![]() |
[8] |
W. Dai, Mean-variance hedging based on an incomplete market with external risk factors of non-Gaussian OU processes, Math. Probl. Eng., 2015 (2015), 625289. https://doi.org/10.1155/2015/625289 doi: 10.1155/2015/625289
![]() |
[9] |
W. Dai, Convolutional neural network based simulation and analysis for backward stochastic partial differential equations, Comput. Math. Appl., 119 (2022), 21–58. https://doi.org/10.1016/j.camwa.2022.05.019 doi: 10.1016/j.camwa.2022.05.019
![]() |
[10] |
W. Dai, Optimal policy computing for blockchain based smart contracts via federated learning, Oper. Res. Int. J., 22 (2022), 5817–5844. https://doi.org/10.1007/s12351-022-00723-z doi: 10.1007/s12351-022-00723-z
![]() |
[11] |
L. Gonon, L. Grigoryeva, J. P. Ortega, Approximation bounds for random neural networks and reservoir systems, Ann. Appl. Probab., 33 (2023), 28–69. https://doi.org/10.1214/22-AAP1806 doi: 10.1214/22-AAP1806
![]() |
[12] | R. Gozalo-Brizuela, E. C. Garrido-Merchan, ChatGPT is not all you need. A state of the art review of large generative AI models, preprint paper, 2023. https://doi.org/10.48550/arXiv.2301.04655 |
[13] | S. Haykin, Neural networks: A Comprehensive Foundation, New Jersey: Prentice Hall PTR, 1994. |
[14] |
K. Hornik, M. Stinchcombe, H. White, Multilayer feedforward networks are universal approximators, Neur. Networks, 2 (1989), 359–366. https://doi.org/10.1016/0893-6080(89)90020-8 doi: 10.1016/0893-6080(89)90020-8
![]() |
[15] | N. Ikeda, S. Watanabe, Stochastic Differential Equations and Diffusion Processes, 2 Eds., Kodansha: North-Holland, 1989. |
[16] | O. Kallenberg, Foundation of Modern Probability, Berlin: Springer, 1997. |
[17] | A. N. Kolmogorov, On the representation of continuous functions of several variables as superpositions of continuous functions of a smaller number of variables, Dokl. Akad. Nauk, 108 (1956). |
[18] |
D. Kramkov, M. Sirbu, On the two times differentiability of the value function in the problem of optimal investment in incomplete markets, Ann. Appl. Probab., 16 (2006), 1352–1384. https://doi.org/10.1214/105051606000000259 doi: 10.1214/105051606000000259
![]() |
[19] | A. Kratsios, V. Debarnot, I. Dokmannić, Small transformers compute universal metric embeddings, J. Mach. Learning Res., 24 (2023), 1–48. |
[20] |
Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, et al., Backpropagation applied to handwritten zip code recognition, Neur. Comput., 1 (1989), 541–551. https://doi.org/10.1162/neco.1989.1.4.541 doi: 10.1162/neco.1989.1.4.541
![]() |
[21] | Z. Liu, Y. Wang, S. Vaidya, F. Ruehle, J. Halverson, M. Solja˘cić, et al., KAN: Kolmogorov-Arnold networks, preprint paper, 2024. https://arXiv.org/pdf/2404.19756 |
[22] | M. Musiela, T. Zariphopoulou. Stochastic partial differential equations and portfolio choice, In: Contemporary Quantitative Finance, Berlin: Springer, 2009. https://doi.org/10.1007/978-3-642-03479-4_11 |
[23] | B. ∅ksendal, Stochastic Differential Equations, 6 Eds, New York: Springer, 2005. |
[24] | B. ∅ksendal, A. Sulem, T. Zhang, A stochastic HJB equation for optimal control of forward-backward SDEs, In: The Fascination of Probability, Statistics and their Applications, Berlin: Springer, 2016. |
[25] | S. Peluchetti, Diffusion bridge mixture transports, Schr¨odinger bridge problems and generative modeling, J. Mach. Learning Res., 24 (2023), 1–51. |
[26] |
J. Sirignano, K. Spiliopoulos, Dgm: a deep learning algorithm for solving partial differential equations, J. Comput. Phys., 375 (2018), 1339–1364. https://doi.org/10.1016/j.jcp.2018.08.029 doi: 10.1016/j.jcp.2018.08.029
![]() |
[27] | A. Vaswani, N. Shazeer, N. Parmar, J. Uszkoreit, L. Jones, A. N. Gomez, et al., Attention is all you need, Adv. Neur. Informa. Proc. Syst., 30 (2017), 5998–6008. |
[28] |
R. Yamashitza, M. Nishio, R. K. G. Do, Togashi, Convolutional neural networks: an overview and application in radiology, Insights into Imaging, 9 (2018), 611–629. https://doi.org/10.1007/s13244-018-0639-9 doi: 10.1007/s13244-018-0639-9
![]() |
1. | Wanyang Dai, Stochastic Differential Games and a Unified Forward–Backward Coupled Stochastic Partial Differential Equation with Lévy Jumps, 2024, 12, 2227-7390, 2891, 10.3390/math12182891 | |
2. | Wanyang Dai, Gene mutation estimations via mutual information and Ewens sampling based CNN & machine learning algorithms, 2025, 0266-4763, 1, 10.1080/02664763.2025.2460076 |