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

Influence of risk tolerance on long-term investments: A Malliavin calculus approach

Hyungbin Park

Department of Mathematical Sciences and RIMS, Seoul National University, 1, Gwanak-ro, Gwanak-gu, Seoul, Republic of Korea; [email protected], [email protected]
Abstract

This study investigates the influence of risk tolerance on the expected utility in the long run. We estimate the extent to which the expected utility of optimal portfolios is affected by small changes in the risk tolerance. For this purpose, we adopt the Malliavin calculus method and the Hansen–Scheinkman decomposition, through which the expected utility is expressed in terms of the eigenvalues and eigenfunctions of an operator. We conclude that the influence of risk aversion on the expected utility is determined by these eigenvalues and eigenfunctions in the long run.

1 Introduction

1.1 Overview

The risk–return trade-off is an important issue in finance. Investors assess returns against risk while considering investment strategies, and they choose strategies that maximize the returns while minimizing the risk. There are several ways to formulate the risk–return trade-off. One of the most commonly accepted forms is the utility function, which reflects the risk tolerance of an investor. We quantify the extent to which the expected utility is affected by small changes in the risk tolerance.

This study investigates long-term investment strategies. Under several market models, utility-maximizing portfolios and their expected utility are considered. For this purpose, we adopt the Malliavin calculus method and the Hansen–Scheinkman decomposition, through which the expected utility is expressed in terms of the eigenvalues and eigenfunctions of an operator. We conclude that the influence of risk tolerance on the long-term expected utility is determined by these eigenvalues and eigenfunctions.

The main objective of this study is to investigate the influence of small changes of the utility function on long-term investments. We recall the classical utility maximization problem

uT:=max𝔼[U(ΠT)]-u_{T}:=\max\mathbb{E}[U(\Pi_{T})]

over all admissible self-financing portfolios with a given initial capital for terminal time T.T. The constant relative risk aversion (CRRA) utility function

U(x)=xννU(x)=\frac{x^{\nu}}{\nu}

for ν<0\nu<0 is considered in this paper. The parameter ν\nu represents how an investor measures his/her degree of risk tolerance. The optimal portfolio Π^=Π^(ν)\hat{\Pi}=\hat{\Pi}^{(\nu)} depends on the parameter ν,\nu, thus

uT=𝔼[U(Π^T)]=𝔼[Π^Tν]/ν.-u_{T}=\mathbb{E}[U(\hat{\Pi}_{T})]=\mathbb{E}[\hat{\Pi}_{T}^{\nu}]/\nu\,.

We investigate the influence of small changes in the parameter ν\nu on the expected utility over a long time horizon. The influence of small changes in the parameter ν\nu in terms of its logarithmic value can be mathematically expressed as

νlnuT.\frac{\partial}{\partial\nu}\ln u_{T}\,.

We estimate the large-time behavior of this partial derivative as T.T\to\infty.

The main methodology for this analysis is a combination of the Hansen–Scheinkman decomposition and the Malliavin calculus technique presented in Sections 2 and 3. This approach is from Park (2018). First, we transform the expected utility into the expectation form

pT=𝔼ξ[e0Tq(Xs)𝑑s]p_{T}=\mathbb{E}_{\xi}[e^{-\int_{0}^{T}q(X_{s})\,ds}]

for some Markov diffusion process X=(Xt)0tTX=(X_{t})_{0\leq t\leq T} with X0=ξX_{0}=\xi and some measurable function q.q. The process XX with killing rate qq induces an infinitesimal generator. Using the Hansen–Scheinkman decomposition, we can find an eigenvalue λ\lambda and a positive eigenfunction ϕ\phi of the generator as well as a measurable function ff such that the expectation is written as

pT=ϕ(ξ)eλTf(T,ξ).p_{T}=\phi(\xi)e^{-\lambda T}f(T,\xi)\,.

The real number λ\lambda and the functions ϕ\phi and ff depend on the parameter ν.\nu. By differentiating with respect to ν,\nu, we have

νlnpT=νlnϕ(ξ)Tλν+νlnf(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=\frac{\partial}{\partial\nu}\ln\phi(\xi)-T\frac{\partial\lambda}{\partial\nu}+\frac{\partial}{\partial\nu}\ln f(T,\xi)\,.

Using the above-mentioned equation, we estimate the influence of risk tolerance on the long-term expected utility. If νlnf(T,ξ)\frac{\partial}{\partial\nu}\ln f(T,\xi) is bounded in TT on [0,),[0,\infty), then

|1TνlnpT+λν|cT\Big{|}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\Big{|}\leq\frac{c}{T}

for some positive constant c.c. This implies that the influence of risk tolerance is asymptotically equal to the partial derivative of λ-\lambda with respect to ν,\nu, which is the main conclusion of this study. The Malliavin calculus technique is used to verify that 1Tνlnf(T,ξ)\frac{1}{T}\frac{\partial}{\partial\nu}\ln f(T,\xi) is bounded in TT on [0,).[0,\infty). We cover several market models, namely the Ornstein–Uhlenbeck process, the CIR process, the 3/23/2 model, a quadratic drift model.

The remainder of this paper is organized as follows. The related literature is reviewed in Section 1.2. The utility maximization problem, the Hansen–Scheinkman decomposition and the Malliavin calculus method are explained as mathematical preliminaries in Section 2. The main ideas and arguments for investigating the influence of risk tolerance are discussed in Section 3. The influence of risk tolerance on utility-maximizing portfolios is illustrated in Section 4. Finally, our findings are summarized in Section 5. The technical details are presented in the appendices.

1.2 Related literature

Many authors have studied the stability of the optimal investment strategy with respect to the utility function. Jouini and Napp (2004) studied in a general complete financial market the stability of the optimal investment-consumption strategy with respect to the choice of the utility function. More precisely, for a given sequence of utility functions that converges pointwise, they proved the almost sure as well as the LpL^{p}-convergence (p1)(p\geq 1) of the optimal wealth and consumption at each date. Carassus and Rásonyi (2007) investigated the convergence of optimal strategies with respect to a sequence of utility functions. They also established the continuity of the utility indifference price with respect to changes in agents’ preferences. Nutz (2012) considered the economic problem of optimal consumption and investment with power utility. As the relative risk aversion tends to infinity or to one was proved, the convergence of the optimal consumption is obtained for general semimartingale models while the convergence of the optimal trading strategy is obtained for continuous models.

The dependence of the risk tolerance on the investment strategy has been studied many authors. Zariphopoulou and Zhou (2009) analyzed a portfolio choice problem when the local risk tolerance is time-dependent and asymptotically linear in wealth. This methodology allows the investment performance to be measured in terms of the risk tolerance and alternative market views. Mocha and Westray (2013) studied the sensitivity of the power utility maximization problem with respect to the investor’s relative risk aversion, the statistical probability measure, the investment constraints, and the market price of risk. Paravisini et al. (2017) estimated risk tolerance from investors’ financial decisions in a person-to-person lending platform. They developed a method that obtains a risk-tolerance parameter from each portfolio choice on the basis of the elasticity of risk tolerance to changes in wealth. Bi and Cai (2019) investigated the optimal investment–reinsurance strategies for an agent with state-dependent risk tolerance and value-at-risk constraints. They derived the closed-form expressions of the optimal strategies and discussed the impact of the risk tolerance. Delong (2019) considered agents whose risk tolerance consists of a constant risk tolerance and a small wealth-dependent risk tolerance. He investigated an exponential utility maximization problem for an agent who faces a stream of non-hedgeable claims.

Numerous studies have investigated the topic of long-term investment strategies. Fleming and Sheu (2000) considered an optimal investment model to maximize the long-term growth rate of the expected utility of wealth. The problem was reformulated as a risk‐sensitive control problem with an infinite time horizon. Hansen and Scheinkman (2009), Hansen (2012) and Hansen and Scheinkman (2012) exploited the Hansen–Scheinkman decomposition method and demonstrated a long-term risk–return trade-off. Guasoni and Robertson (2015) studied a class of static fund separation theorems that is valid for investors with a long time horizon and constant relative risk tolerance. Robertson and Xing (2015) investigated long-term portfolio choice problems by analyzing the large-time asymptotic behavior of solutions to semi-linear Cauchy problems.

Malliavin calculus has been studied in relation to various topics in quantitative finance. Fournié et al. (1999) investigated a probabilistic method for computations of Greeks in finance. This methodology is based on the integration-by-parts formula developed in Malliavin calculus. Benhamou (2003) showed that the Malliavin weight functions for Greeks must satisfy necessary and sufficient conditions expressed as conditional expectations. Alos et al. (2007) used Malliavin calculus techniques to obtain an expression for the short-time behavior of the at-the-money implied volatility skew for a generalization of the Bates model, where the volatility does not need to be a diffusion or a Markov process. Borovička et al. (2014) studied the shock elasticity, which reflects the sensitivity with respect to a perturbation over a time instant. They proposed a Malliavin calculus method to compute the shock elasticity. Alòs and Shiraya (2019) studied the difference between the fair strike of a volatility swap and the at-the-money implied volatility of a European call option. They used the Malliavin calculus approach to derive an exact expression for this difference. Park and Sturm (2019) employed the Malliavin calculus method to investigate the sensitivities of the long-term expected utility of optimal portfolios under incomplete markets.

As a closely related article, Park (2018) conducted a sensitivity analysis of long-term cash flows. However, the perturbation form is different from this paper. He studied the extent to which the price of the cash flow is affected by small perturbations of the underlying Markov diffusion. He considered the drift and volatility perturbations in the underlying process

dXtϵ=bϵ(Xtϵ)dt+σϵ(Xtϵ)dBt,X0ϵ=ξdX_{t}^{\epsilon}=b_{\epsilon}(X_{t}^{\epsilon})\,dt+\sigma_{\epsilon}(X_{t}^{\epsilon})\,dB_{t}\;,\,X_{0}^{\epsilon}=\xi

for the perturbation parameter ϵ\epsilon, and analyzed their influence to the cash flows. This paper, however, works with the investor’s risk tolerance and the perturbations in the underlying process is not considered.

2 Mathematical preliminary

In this section, we demonstrate the utility maximization problem and the two main methodologies exploited in the rest of this study: the Hansen–Scheinkman decomposition and the Malliavin calculus method.

2.1 Utility maximization problem

We consider the classical utility maximization problem in financial markets. Let (Ω,,(t)t0,𝐏)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},{\bf P}) be a filtered probability space having a one-dimensional Brownian motion (Zt)t0(Z_{t})_{t\geq 0} and a nn-dimensional Brownian motion WW with constant correlation ρ=(ρ1,,ρn)=dZ,Wt/dt.\rho=(\rho_{1},\cdots,\rho_{n})^{\top}=d\left\langle Z,\,W\right\rangle_{t}/dt. The measure 𝐏{\bf P} is referred to as the physical measure and the filtration (t)t0(\mathcal{F}_{t})_{t\geq 0} is the argumentation of the natural filtration of ZZ and W.W. We assume that the market has a state process (Xt)t0,(X_{t})_{t\geq 0}, which is a Markov diffusion satisfying

dXt=b(Xt)dt+σ(Xt)dBt,X0=ξdX_{t}=b(X_{t})\,dt+\sigma(X_{t})\,dB_{t}\,,\;X_{0}=\xi (2.1)

There are n+1n+1 assets S(0),S(1),,S(n)S^{(0)},S^{(1)},\ldots,S^{(n)} in the market, where S(0)S^{(0)} is a risk-free asset and S(1),,S(n)S^{(1)},\ldots,S^{(n)} are risky assets. The assets dynamics are given as

dSt(0)St(0)=r(Xt)dt,\displaystyle\frac{dS^{(0)}_{t}}{S^{(0)}_{t}}=r(X_{t})\,dt\,, (2.2)
dSt(i)St(i)=(r(Xt)+μi(Xt))dt+j=1nυij(Yt)dWt(j)1in.\displaystyle\dfrac{dS^{(i)}_{t}}{S^{(i)}_{t}}=(r(X_{t})+\mu_{i}(X_{t}))dt+\sum_{j=1}^{n}\upsilon_{ij}(Y_{t})\,dW^{(j)}_{t}\qquad 1\leq i\leq n\,.

Here, rr is the short interest rate function, μ=(μ1,,μn)\mu=(\mu_{1},\cdots,\mu_{n})^{\top} is the excess return function, and υ=(υij)1i,jn\upsilon=(\upsilon_{ij})_{1\leq i,j\leq n} is the volatility matrix functions.

A portfolio is a nn-dimensional process π=(πt)t0=(πt(1),,πt(n))t0\pi=(\pi_{t})_{t\geq 0}=(\pi^{(1)}_{t},\ldots,\pi^{(n)}_{t})_{t\geq 0}, which represents the proportions of wealth in each risky asset. The wealth process (Πt)t0=(Πtπ)t0(\Pi_{t})_{t\geq 0}=(\Pi_{t}^{\pi})_{t\geq 0} of π\pi with positive initial capital ω\omega satisfies

dΠtΠt=(r(Xt)+πtμ(Xt))dt+πtυ(Xt)dWt,Π0=ω>0.\frac{d\Pi_{t}}{\Pi_{t}}=(r(X_{t})+\pi_{t}\mu(X_{t}))\,dt+\pi_{t}\upsilon(X_{t})\,dW_{t},\quad\Pi_{0}=\omega>0. (2.3)

We assume that the portfolio process (πt)t0(\pi_{t})_{t\geq 0} is (t)t0(\mathcal{F}_{t})_{t\geq 0}-adapted and the integrations in Eq.(2.3) are well-defined, that is, 𝔼0t|(r(Xs)+πsμ(Xs))|+πsυ(Xs)2ds<\mathbb{E}\int_{0}^{t}|(r(X_{s})+\pi_{s}\mu(X_{s}))|+|\!|\pi_{s}\upsilon(X_{s})|\!|^{2}\,ds<\infty for all t0,t\geq 0, where |||||\!|\cdot|\!| is the usual Euclidean norm. The wealth process Πt>0\Pi_{t}>0 for all t0t\geq 0 a.s. since the initial wealth ω>0.\omega>0.

In this market, we consider an agent who wants to maximize the expected utility of the terminal wealth over all possible portfolios. More precisely, we are interested in

supπ𝔼𝐏[U(ΠTπ)]\sup_{\pi}\mathbb{E}^{\bf P}[U(\Pi_{T}^{\pi})]

over all possible portfolios π\pi for given positive initial capital ω.\omega. The utility function is assumed to be a power function of the form

U(x)=xν/νU(x)=x^{\nu}/\nu (2.4)

for ν<0.\nu<0.

This utility maximization problem can be solved by using stochastic control theory. Define the value function uTu_{T} as

uT=uT(t,ω,x)=supπ𝔼𝐏[U(ΠTπ)|Πt=ω,Xt=x].-u_{T}=-u_{T}(t,\omega,x)=\sup_{\pi}\mathbb{E}^{\bf P}[U(\Pi_{T}^{\pi})|\Pi_{t}=\omega,X_{t}=x]\,.

Following (Zariphopoulou, 2001, Proposition 2.1), we have

uT(t,ω,x)=ωνν(𝔼[eν(1ν+νρρ)1νtT(r+12(1ν)θθ)(Xu)𝑑u|Xt=x])1ν1ν+νρρu_{T}(t,\omega,x)=-\frac{\omega^{\nu}}{\nu}(\mathbb{E}^{\mathbb{P}}[e^{\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{1-\nu}\int_{t}^{T}(r+\frac{1}{2(1-\nu)}\theta^{\prime}\theta)(X_{u})\,du}|X_{t}=x])^{\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}}

where θ:=υ1μ,\theta:=\upsilon^{-1}\mu, b:=k+ν1νσρθb:=k+\frac{\nu}{1-\nu}\sigma\rho^{\prime}\theta and the \mathbb{P}-dynamics of XX is

dXt=b(Xt)dt+σ(Xt)dBtdX_{t}=b(X_{t})\,dt+\sigma(X_{t})\,dB_{t} (2.5)

with a \mathbb{P}-Brownian motion (Bt)t0.(B_{t})_{t\geq 0}. For t=0,t=0, it follows that

uT=uT(0,ω,ξ)=ωννqT1ν1ν+νρρu_{T}=u_{T}(0,\omega,\xi)=-\frac{\omega^{\nu}}{\nu}q_{T}^{\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}}

for

qT:=𝔼ξ[eν(1ν+νρρ)1ν0T(r+12(1ν)θθ)(Xu)𝑑u].q_{T}:=\mathbb{E}^{\mathbb{P}}_{\xi}[e^{\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{1-\nu}\int_{0}^{T}(r+\frac{1}{2(1-\nu)}\theta^{\prime}\theta)(X_{u})\,du}]\,. (2.6)

Here, we have used the notation 𝔼ξ[]=𝔼[|X0=ξ].\mathbb{E}^{\mathbb{P}}_{\xi}[\,\cdots]=\mathbb{E}^{\mathbb{P}}[\,\cdots|X_{0}=\xi]. In particular, when the short rate is a constant r,r,

uT=ωννerνTpT1ν1ν+νρρu_{T}=-\frac{\omega^{\nu}}{\nu}e^{r\nu T}p_{T}^{\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}} (2.7)

where

pT:=𝔼ξ[eqtT(θθ)(Xu)𝑑u]p_{T}:=\mathbb{E}_{\xi}^{{\mathbb{P}}}[e^{-q\int_{t}^{T}(\theta^{\prime}\theta)(X_{u})\,du}] (2.8)

and q:=ν(1ν+νρρ)2(1ν)2.q:=-\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{2(1-\nu)^{2}}.

In conclusion, the problem of analyzing the optimal expected utility boils down to the problem of analyzing the expectation qTq_{T} in Eq.(2.6) (or pTp_{T} in Eq.(2.8) if the short rate is constant) with the dynamics of XX given as Eq.(2.5). Later this will be used in Step I in Section 3.

2.2 Hansen–Scheinkman decomposition

We briefly review the Hansen–Scheinkman decomposition as a mathematical preliminary. Readers may refer to Hansen and Scheinkman (2009), Park (2018), and Qin and Linetsky (2016) for further details. Consider a filtered probability space (Ω,,(t)t0,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},{\mathbb{P}}) having a dd-dimensional Brownian motion B=(Bt(1),,Bt(d))t0.B=(B_{t}^{(1)},\cdots,B_{t}^{(d)})_{t\geq 0}^{\top}. The family (t)t0(\mathcal{F}_{t})_{t\geq 0} is the completed filtration generated by B.B.

We begin with the state space 𝒟\mathcal{D} and four functions b,σ,q,h.b,\sigma,q,h. Let 𝒟\mathcal{D} be an open and connected subset of d,\mathbb{R}^{d}, and let b:𝒟db:\mathcal{D}\to\mathbb{R}^{d} and σ:𝒟d×d\sigma:\mathcal{D}\to\mathbb{R}^{d}\times\mathbb{R}^{d} be continuously differentiable functions. The matrix σ\sigma is invertible. The function q:𝒟q:\mathcal{D}\to\mathbb{R} is continuous and the function h:𝒟h:\mathcal{D}\to\mathbb{R} is nonnegative continuously differentiable. For each ξ𝒟,\xi\in\mathcal{D}, assume that the SDE Eq.(2.1) has a unique strong solution XX on 𝒟.\mathcal{D}.

We can consider an infinitesimal generator and its eigenpair. Define the infinitesimal generator of the process XX with killing rate qq as

:=12i,j=1daij2xixj+i=1dbixiq,\mathcal{L}:=\frac{1}{2}\sum_{i,j=1}^{d}a_{ij}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}+\sum_{i=1}^{d}b_{i}\frac{\partial}{\partial x_{i}}-q, (2.9)

where a=σσ.a=\sigma\sigma^{\top}. For a real number λ\lambda and a positive C2C^{2}-function ϕ:𝒟,\phi:\mathcal{D}\to\mathbb{R}, we say that a pair (λ,ϕ)(\lambda,\phi) is an eigenpair of -\mathcal{L} if

ϕ=λϕon 𝒟.\mathcal{L}\phi=-\lambda\phi\quad\textnormal{on }\mathcal{D}\,. (2.10)

For T>0,T>0, by applying the Ito formula, we can show that for each eigenpair (λ,ϕ),(\lambda,\phi), a positive process

Mtϕ:=ϕ(Xt)ϕ(ξ)eλt0tq(Xs)𝑑s, 0tTM_{t}^{\phi}:=\frac{\phi(X_{t})}{\phi(\xi)}e^{\lambda t-\int_{0}^{t}q(X_{s})\,ds}\,,\;0\leq t\leq T (2.11)

is a local martingale. Assume that this process is a martingale. We can define a measure ^tϕ\hat{{\mathbb{P}}}_{t}^{\phi} on each t\mathcal{F}_{t} for 0tT0\leq t\leq T by

d^tϕd=Mtϕ.\frac{d\hat{{\mathbb{P}}}_{t}^{\phi}}{d{\mathbb{P}}\;}=M_{t}^{\phi}\,.

The family (^tϕ)0tT(\hat{{\mathbb{P}}}_{t}^{\phi})_{0\leq t\leq T} is consistent, i.e., ^tϕ|t=^tϕ\hat{{\mathbb{P}}}_{t^{\prime}}^{\phi}|_{\mathcal{F}_{t}}=\hat{{\mathbb{P}}}_{t}^{\phi} for all 0ttT.0\leq t\leq t^{\prime}\leq T. For fixed T>0,T>0, we use the notation ^ϕ\hat{\mathbb{P}}^{\phi} instead of ^Tϕ,\hat{{\mathbb{P}}}_{T}^{\phi}, suppressing T.T. This measure on T\mathcal{F}_{T} is called the eigen-measure with respect to ϕ.\phi. The process

B^tϕ=0t(σϕ/ϕ)(Xs)𝑑s+Bt,t0\hat{B}_{t}^{\phi}=-\int_{0}^{t}(\sigma^{\top}\nabla\phi/\phi)(X_{s})\,ds+B_{t}\,,\;t\geq 0

is a ^ϕ\hat{{\mathbb{P}}}^{\phi}-Brownian motion by the Girsanov theorem, and the process XX satisfies

dXt=(b+σσϕ/ϕ)(Xt)dt+σ(Xt)dB^tϕ.dX_{t}=(b+\sigma\sigma^{\top}\nabla\phi/\phi)(X_{t})\,dt+\sigma(X_{t})\,d\hat{B}_{t}^{\phi}\,.

Under these circumstances, consider the decomposition of the discount factor

e0tq(Xs)𝑑s=Mtϕeλtϕ(ξ)ϕ(Xt),t0,e^{-\int_{0}^{t}q(X_{s})\,ds}=M_{t}^{\phi}e^{-\lambda t}\frac{\phi(\xi)}{\phi(X_{t})}\,,\;t\geq 0, (2.12)

which comes from Eq.(2.11). This expression is called the Hansen–Scheinkman decomposition. The expectation pTp_{T} can be written as

pT=𝔼ξ[e0Tq(Xs)𝑑s]\displaystyle p_{T}=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-\int_{0}^{T}q(X_{s})\,ds}] =ϕ(ξ)eλT𝔼ξ[MTϕϕ(XT)]\displaystyle=\phi(\xi)\,e^{-\lambda T}\,\mathbb{E}_{\xi}^{\mathbb{P}}\Big{[}\frac{M_{T}^{\phi}}{\phi(X_{T})}\Big{]} (2.13)
=ϕ(ξ)eλT𝔼ξ^ϕ[1ϕ(XT)]\displaystyle=\phi(\xi)\,e^{-\lambda T}\,\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\phi}}\Big{[}\frac{1}{\phi(X_{T})}\Big{]}
=ϕ(ξ)eλTfϕ(T,ξ),\displaystyle=\phi(\xi)e^{-\lambda T}f_{\phi}(T,\xi),

where fϕ(t,x):=𝔼x^ϕ[1ϕ(Xt)]f_{\phi}(t,x):=\mathbb{E}_{x}^{{\hat{\mathbb{P}}}^{\phi}}[\frac{1}{\phi(X_{t})}] for t0t\geq 0 and x𝒟.x\in\mathcal{D}. The function fϕf_{\phi} is referred to as the remainder function. The decomposition in Eq.(2.13) is useful for the analysis of pT,p_{T}, because the expectation 𝔼ξ^ϕ[1ϕ(XT)]\mathbb{E}_{\xi}^{{\hat{\mathbb{P}}}^{\phi}}[\frac{1}{\phi(X_{T})}] depends on the final random variable XT,X_{T}, whereas the expression

pT=𝔼ξ[e0Tq(Xs)𝑑s]p_{T}=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-\int_{0}^{T}q(X_{s})\,ds}]

depends on the entire path of (Xt)0tT.(X_{t})_{0\leq t\leq T}. If we know the ^ϕ\hat{\mathbb{P}}^{\phi}-distribution of XTX_{T}, then we can analyze the expectation 𝔼ξ^ϕ[1ϕ(XT)]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\phi}}[\frac{1}{\phi(X_{T})}] directly. For notational simplicity, when the eigenpair is specified, we use the notations M,M, ^,\hat{{\mathbb{P}}}, (B^t)t0(\hat{B}_{t})_{t\geq 0}, and ff instead of Mϕ,M^{\phi}, ^ϕ,\hat{{\mathbb{P}}}^{\phi}, (B^tϕ)t0(\hat{B}_{t}^{\phi})_{t\geq 0}, and fϕ,f_{\phi}, respectively, suppressing ϕ.\phi.

We summarize this section in the following proposition.

Proposition 2.1.

Let 𝒟\mathcal{D} be an open and connected subset of d.\mathbb{R}^{d}. Assume the following conditions.

  1. (i)

    The functions b:𝒟db:\mathcal{D}\to\mathbb{R}^{d} and σ:𝒟d×d\sigma:\mathcal{D}\to\mathbb{R}^{d}\times\mathbb{R}^{d} are continuously differentiable functions, and the matrix σ\sigma is invertible. The function q:𝒟q:\mathcal{D}\to\mathbb{R} is continuous.

  2. (ii)

    For each ξ𝒟,\xi\in\mathcal{D}, the SDE (2.1) has a unique strong solution XX on 𝒟.\mathcal{D}.

If (λ,ϕ)(\lambda,\phi) is an eigenpair of the operator -\mathcal{L} in Eq.(2.9), then the process MϕM^{\phi} in Eq.(2.11) is a local martingale. If this process is a martingale, then

𝔼ξ[e0Tq(Xs)𝑑s]=ϕ(ξ)eλT𝔼ξ^[1ϕ(XT)]\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-\int_{0}^{T}q(X_{s})\,ds}]=\phi(\xi)\,e^{-\lambda T}\,\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\frac{1}{\phi(X_{T})}\Big{]}

where ^\hat{\mathbb{P}} is the eigen-measure with respect to ϕ.\phi.

2.3 Malliavin calculus

This section presents a brief review of Malliavin calculus. For further details, refer to Malliavin and Paul (2006) and Nualart and Nualart (2018). Let (Ω,,(t)0tT,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{0\leq t\leq T},\mathbb{P}) be a filtered probability space having a one-dimensional Brownian motion B.B. The filtration (t)0tT(\mathcal{F}_{t})_{0\leq t\leq T} is the natural filtration of B.B. Define the set of all cylindrical random variables as

𝒮:={f(0Th(1)(s)𝑑Bs,,0Th(n)(s)𝑑Bs):n,h(1),,h(n)L2[0,T],fCp(n)},\mathcal{S}:=\Big{\{}f\Big{(}\int_{0}^{T}h^{(1)}(s)\,dB_{s},\cdots,\int_{0}^{T}h^{(n)}(s)\,dB_{s}\Big{)}:n\in\mathbb{N},h^{(1)},\cdots,h^{(n)}\in L^{2}[0,T],f\in C_{p}^{\infty}(\mathbb{R}^{n})\Big{\}}\,,

where Cp(n)C_{p}^{\infty}(\mathbb{R}^{n}) is the set of all CC^{\infty} functions such that ff and all its partial derivatives have polynomial growth. The Malliavin derivative of F𝒮F\in\mathcal{S} is defined as a stochastic process DF=(DtF)0tTDF=(D_{t}F)_{0\leq t\leq T} given by

DtF=i=1nh(i)(t)fxi(0Th(1)(s)𝑑Bs,,0Th(n)(s)𝑑Bs).D_{t}F=\sum_{i=1}^{n}h^{(i)}(t)\frac{\partial f}{\partial x_{i}}\Big{(}\int_{0}^{T}h^{(1)}(s)\,dB_{s},\cdots,\int_{0}^{T}h^{(n)}(s)\,dB_{s}\Big{)}\,.

Then, DD is a linear operator from 𝒮L2(Ω)\mathcal{S}\subseteq L^{2}(\Omega) to L2([0,T]×Ω),L^{2}([0,T]\times\Omega), and it is known that DD is closable. The closure is also denoted as D.D. The domain of DD is the closure of 𝒮\mathcal{S} under the norm

F𝔻1,2:=FL2(Ω)+DFL2([0,T]×Ω)|\!|F|\!|_{\mathbb{D}^{1,2}}:=|\!|F|\!|_{L^{2}(\Omega)}+|\!|DF|\!|_{L^{2}([0,T]\times\Omega)}

and is denoted as 𝔻1,2.\mathbb{D}^{1,2}.

We will use the following propositions. Propositions 2.2, 2.3, and 2.5 are from (León et al., 2003, Lemma 2.1), (Protter, 2004, Theorem 39 on page 312), and (Park, 2018, Proposition A.1), respectively.

Proposition 2.2.

Let φ:\varphi:\mathbb{R}\to\mathbb{R} be a continuously differentiable function and F𝔻1,2.F\in\mathbb{D}^{1,2}. Then, φ(F)𝔻1,2\varphi(F)\in\mathbb{D}^{1,2} if and only if φ(F)L2(Ω)\varphi(F)\in L^{2}(\Omega) and φ(F)DFL2([0,T]×Ω),\varphi^{\prime}(F)DF\in L^{2}([0,T]\times\Omega), and in this case,

Dφ(F)=φ(F)DF.D\varphi(F)=\varphi^{\prime}(F)DF\,.
Proposition 2.3.

Let X=X(x)X=X^{(x)} be a Markov diffusion whose dynamics is given as

dXt=b(Xt)dt+σ(Xt)dBtdX_{t}=b(X_{t})\,dt+\sigma(X_{t})\,dB_{t}

with initial value X0=xX_{0}=x, where bb and σ\sigma are continuously differentiable functions with bounded derivatives. Then, the map xXt(x)x\mapsto X_{t}^{(x)} is continuously differentiable almost surely and the derivative process Yt:=xXt(x)Y_{t}:=\frac{\partial}{\partial x}X_{t}^{(x)} satisfies

dYt=b(Xt)Ytdt+σ(Xt)YtdBt,Y0=1,dY_{t}=b^{\prime}(X_{t})Y_{t}\,dt+\sigma^{\prime}(X_{t})Y_{t}\,dB_{t}\,,Y_{0}=1\,,

equivalently,

Yt=e0t(b(Xs)12σ2(Xs))𝑑s+0tσ(Xs)𝑑Bs.Y_{t}=e^{\int_{0}^{t}(b^{\prime}(X_{s})-\frac{1}{2}\sigma^{\prime 2}(X_{s}))\,ds+\int_{0}^{t}\sigma^{\prime}(X_{s})\,dB_{s}}\,.

Moreover, Xt𝔻1,2X_{t}\in\mathbb{D}^{1,2} for each 0tT0\leq t\leq T and its Malliavin derivative satisfies

DsXt=σ(Xs)YtYs𝕀{st}.D_{s}X_{t}=\sigma(X_{s})\frac{Y_{t}}{Y_{s}}\mathbb{I}_{\{s\leq t\}}.
Proposition 2.4.

(Integration by parts formula) Let F𝔻1,2F\in\mathbb{D}^{1,2} and (ht)0tT(h_{t})_{0\leq t\leq T} be a progressively measurable process with 𝔼0Ths2𝑑s<.\mathbb{E}\int_{0}^{T}h_{s}^{2}\,ds<\infty. Then

𝔼(F0Ths𝑑Bs)=𝔼(0T(DsF)hs𝑑s).\mathbb{E}\Big{(}F\int_{0}^{T}h_{s}\,dB_{s}\Big{)}=\mathbb{E}\Big{(}\int_{0}^{T}(D_{s}F)h_{s}\,ds\Big{)}\,.
Proposition 2.5.

Let b,b, b¯,\overline{b}, σ\sigma be continuously differentiable functions with σ>0\sigma>0 and let ff be a continuous function on an open interval 𝒟.\mathcal{D}\subseteq\mathbb{R}. Define bϵ=b+ϵb¯b_{\epsilon}=b+\epsilon\overline{b} for ϵI\epsilon\in I, where II is an open neighborhood of 0.0. Assume that the SDE

dXt(ϵ)=bϵ(Xt(ϵ))dt+σ(Xt(ϵ))dBt,X0(ϵ)=ξdX_{t}^{(\epsilon)}=b_{\epsilon}(X_{t}^{(\epsilon)})\,dt+\sigma(X_{t}^{(\epsilon)})\,dB_{t}\,,\;X_{0}^{(\epsilon)}=\xi

has a unique strong solution X(ϵ)X^{(\epsilon)} on 𝒟\mathcal{D} for all ξ\xi\in\mathbb{R} and ϵI.\epsilon\in I. Suppose that for T>0T>0, there exist positive constants ϵ0,\epsilon_{0}, ϵ1,\epsilon_{1}, p,p, qq with p2p\geq 2 and 1/p+1/q=11/p+1/q=1 such that

𝔼ξ[eϵ00Tb¯2(Xs)𝑑s]<,\displaystyle\mathbb{E}_{\xi}[e^{\epsilon_{0}\int_{0}^{T}\overline{b}^{2}(X_{s})\,ds}]<\infty\,, (2.14)
𝔼ξ[0T|b¯|p+ϵ1(Xs)𝑑s]<,\displaystyle\mathbb{E}_{\xi}\Big{[}\int_{0}^{T}|\overline{b}|^{p+\epsilon_{1}}(X_{s})\,ds\Big{]}<\infty\,, (2.15)
𝔼ξ[|f|q(XT)]<.\displaystyle\mathbb{E}_{\xi}[|f|^{q}(X_{T})]<\infty\,. (2.16)

Then, the expectation 𝔼[f(Xt(ϵ))]\mathbb{E}[f(X_{t}^{(\epsilon)})] is continuously differentiable in ϵ\epsilon and

ϵ|ϵ=0𝔼[f(XTϵ)]=𝔼[f(XT)0T(σ1b¯)(Xs)𝑑Bs],\displaystyle\frac{\partial}{\partial\epsilon}\Big{|}_{\epsilon=0}\mathbb{E}[f(X_{T}^{\epsilon})]=\mathbb{E}\Big{[}f(X_{T})\int_{0}^{T}(\sigma^{-1}\overline{b})(X_{s})\,dB_{s}\Big{]}, (2.17)

where X=X(0).X=X^{(0)}.

3 Main arguments

This study investigates the influence of risk tolerance on the optimal expected utility in the long run. It involves the following steps.

Step I. Transform the expected utility from the optimal investment strategy into the expectation form

pT=𝔼[e0Tq(Xs)𝑑s],p_{T}=\mathbb{E}^{\mathbb{P}}[e^{-\int_{0}^{T}q(X_{s})\,ds}],

where the \mathbb{P}-dynamics of XX is

dXt=b(Xt)dt+σ(Xt)dBt.dX_{t}=b(X_{t})\,dt+\sigma(X_{t})\,dB_{t}\,.

The drift function b()=b(;ν)b(\cdot)=b(\cdot;\nu) and the killing rate q()=q(;ν)q(\cdot)=q(\cdot;\nu) may depend on ν;\nu; however, the volatility function σ()\sigma(\cdot) does not depend on ν.\nu. This step was conducted in Section 2.1.

Step II. Through the Hansen–Scheinkman decomposition discussed in Section 2.2, the expectation can be expressed as

pT=ϕ(ξ)eλTf(T,ξ),p_{T}=\phi(\xi)e^{-\lambda T}f(T,\xi)\,,

where (λ,ϕ)(\lambda,\phi) is an eigenpair and f(T,ξ)=𝔼ξ^[1ϕ(XT)].f(T,\xi)=\mathbb{E}_{\xi}^{{\hat{\mathbb{P}}}}[\frac{1}{\phi(X_{T})}]. The ^\hat{\mathbb{P}}-dynamics of XX is

dXt=κ(Xt)dt+σ(Xt)dB^t,dX_{t}=\kappa(X_{t})\,dt+\sigma(X_{t})\,d\hat{B}_{t}\,,

where κ:=b+σ2ϕ/ϕ.\kappa:=b+\sigma^{2}\phi^{\prime}/\phi. It follows that

νlnpT=νlnϕ(ξ)Tλν+fν(T,ξ)f(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=\frac{\partial}{\partial\nu}\ln\phi(\xi)-T\frac{\partial\lambda}{\partial\nu}+\frac{f_{\nu}(T,\xi)}{f(T,\xi)}\,. (3.1)

In the remainder function f(T,ξ)=𝔼ξ^[1ϕ(XT)],f(T,\xi)=\mathbb{E}_{\xi}^{{\hat{\mathbb{P}}}}[\frac{1}{\phi(X_{T})}], observe that the drift function κ()=κ(;ν),\kappa(\cdot)=\kappa(\cdot;\nu), the eigenfunction ϕ()=ϕ(;ν)\phi(\cdot)=\phi(\cdot;\nu), and the measure ^=^ν\hat{\mathbb{P}}=\hat{\mathbb{P}}^{\nu} depend on ν.\nu. For convenience, define H(x;ν)=1ϕ(;ν).H(x;\nu)=\frac{1}{\phi(\cdot;\nu)}. Then,

fν(T,ξ)=𝔼ξ^ν[ν|ν=νH(XT;ν)]+ν|ν=ν𝔼ξ^ν[H(XT;ν)].f_{\nu}(T,\xi)=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}\frac{\partial}{\partial{\nu^{\prime}}}\Big{|}_{\nu^{\prime}=\nu}H(X_{T};\nu^{\prime})\Big{]}+\frac{\partial}{\partial\nu^{\prime}}\Big{|}_{\nu^{\prime}=\nu}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu^{\prime}}}[H(X_{T};\nu)]\,.

This is the key observation of this study. The perturbation of the risk aversion is transformed into perturbations of the drift function, payoff function, and eigenfunction.

Step III. We prove (case by case for each model) that the term fν(T,ξ)f(T,ξ)\frac{f_{\nu}(T,\xi)}{f(T,\xi)} is bounded in TT on [0,).[0,\infty). This is achieved as follows. First, show that the denominator f(T,ξ)f(T,\xi) converges to a positive constant as T.T\to\infty. More precisely, the process XX has an invariant distribution π\pi under the measure ^\hat{\mathbb{P}} and the remainder function

f(T,ξ)=𝔼x^ϕ[1ϕ(Xt)]1ϕ𝑑πf(T,\xi)=\mathbb{E}_{x}^{{\hat{\mathbb{P}}}^{\phi}}\Big{[}\frac{1}{\phi(X_{t})}\Big{]}\to\int\frac{1}{\phi}\,d\pi

as TT\to\infty with 1ϕL1(π).\frac{1}{\phi}\in L^{1}(\pi). Second, show that

𝔼ξ^ν[ν|ν=νH(XT;ν)]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}\frac{\partial}{\partial{\nu^{\prime}}}\Big{|}_{\nu^{\prime}=\nu}H(X_{T};\nu^{\prime})\Big{]}

is bounded in TT on [0.).[0.\infty). This can be easily checked by direct calculation; thus, we do not go into further detail here.

Finally, show that

ν|ν=ν𝔼ξ^ν[H(XT;ν)]\frac{\partial}{\partial\nu^{\prime}}\Big{|}_{\nu^{\prime}=\nu}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu^{\prime}}}[H(X_{T};\nu)]

is bounded in TT on [0,).[0,\infty). Observe that the perturbation parameter ν\nu is only in the drift term of the dynamics of X.X. We adopt the Malliavin calculus method to estimate this partial derivative. Assume that the map νκ(x;ν)\nu\mapsto\kappa(x;\nu) is continuously differentiable for each x,x, and denote the first-order approximation as κ¯(x,ν),\overline{\kappa}(x,\nu), i.e.,

κ(x;ν+ϵ)=κ(x;ν)+ϵκ¯(x;ν)+o(ϵ)\kappa(x;\nu+\epsilon)=\kappa(x;\nu)+\epsilon\overline{\kappa}(x;\nu)+o(\epsilon)

as ϵ0.\epsilon\to 0. Under suitable conditions (Propositions 2.2, 2.3, and 2.5),

ν|ν=ν𝔼ξ^ν[H(XT;ν)]\displaystyle\frac{\partial}{\partial\nu^{\prime}}\Big{|}_{\nu^{\prime}=\nu}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu^{\prime}}}[H(X_{T};\nu)] =𝔼ξ^ν[H(XT;ν)0Tκ¯(Xs;ν)σ(Xs)𝑑B^s]\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}H(X_{T};\nu)\int_{0}^{T}\frac{\overline{\kappa}(X_{s};\nu)}{\sigma(X_{s})}\,d\hat{B}_{s}\Big{]} (3.2)
=𝔼ξ^ν[0TDs(H(XT;ν))κ¯(Xs;ν)σ(Xs)𝑑s]\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}\int_{0}^{T}D_{s}(H(X_{T};\nu))\frac{\overline{\kappa}(X_{s};\nu)}{\sigma(X_{s})}\,ds\Big{]}
=𝔼ξ^ν[0THx(XT;ν)DsXTκ¯(Xs;ν)σ(Xs)𝑑s]\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}\int_{0}^{T}H_{x}(X_{T};\nu)D_{s}X_{T}\frac{\overline{\kappa}(X_{s};\nu)}{\sigma(X_{s})}\,ds\Big{]}
=𝔼ξ^ν[Hx(XT;ν)YT0Tκ¯(Xs;ν)Ys𝑑s],\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}^{\nu}}\Big{[}H_{x}(X_{T};\nu)Y_{T}\int_{0}^{T}\frac{\overline{\kappa}(X_{s};\nu)}{Y_{s}}\,ds\Big{]},

where YY is the first variation process of X.X. For all the models in Section 4, this probabilistic representation will be used to show that the partial derivative is bounded in TT on [0,).[0,\infty).

Step IV. Since fν(T,ξ)f(T,ξ)\frac{f_{\nu}(T,\xi)}{f(T,\xi)} is bounded in TT on [0,)[0,\infty) in Eq.(3.1), we finally obtain

|1TνlnpT+λν|cT\left|\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c}{T}

for some positive constant c.c. In particular,

limT1TνlnpT=λν.\lim_{T\to\infty}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}=-\frac{\partial\lambda}{\partial\nu}\,.

This implies that the influence of the risk tolerance parameter ν\nu on long-term investments is determined by the eigenvalue of the generator of the underlying Markov diffusion.

Remark 3.1.

The idea of deriving Eq.(3.2) in Step III is from (Borovička et al., 2014, Section 6.1). They presented the integration-by-parts formula in Malliavin calculus to compute the shock elasticity, and Eq.(3.2) comes from the same method.

4 Utility-maximizing portfolios

We cover several models, namely the Black–Scholes model, the Ornstein–Uhlenbeck (OU) process, the Cox–Ingersoll–Ross (CIR) model, the 3/23/2 model, and a quadratic drift model. Throughout this section, we assume the short rate is a constant rr (and thus, pTp_{T} in Eq.(2.8) will be used).

4.1 Black–Scholes model

As a motivating example, consider a constant proportion portfolio when the underlying market follows the Black–Scholes model. Assume that the short rate is a constant r0r\geq 0 and the stock price follows

dSt=μStdt+σStdZtdS_{t}=\mu S_{t}\,dt+\sigma S_{t}\,dZ_{t}

for μ,\mu\in\mathbb{R}, σ>0.\sigma>0. With the initial capital ω>0,\omega>0, it is known that the optimal expected utility is

uT:=maxπ𝔼𝐏[U(ΠTπ)]=ωννe(r+(μr)22(1ν)σ2)νT.\displaystyle-u_{T}:=\max_{\pi}\mathbb{E}^{\bf P}[U(\Pi_{T}^{\pi})]=\frac{\omega^{\nu}}{\nu}e^{(r+\frac{(\mu-r)^{2}}{2(1-\nu)\sigma^{2}})\nu T}. (4.1)

We aim to investigate the influence of the risk aversion on the long-term investment. We calculate the partial derivative

νlnuT=lnω1ν+(r+(μr)22(1ν)2σ2)T.\frac{\partial}{\partial\nu}\ln u_{T}=\ln\omega-\frac{1}{\nu}+(r+\frac{(\mu-r)^{2}}{2(1-\nu)^{2}\sigma^{2}})T\,.

The partial derivative νlnuT\frac{\partial}{\partial\nu}\ln u_{T} grows linearly as T,T\to\infty, and the linear growth rate is r+(μr)22(1ν)2σ2.r+\frac{(\mu-r)^{2}}{2(1-\nu)^{2}\sigma^{2}}.

4.2 OU process

Under the physical measure 𝐏,{\bf P}, let the state process XX follow the OU process

dXt=(bkXt)dt+σdZt,X0=ξdX_{t}=(b-kX_{t})\,dt+\sigma\,dZ_{t}\,,\;X_{0}=\xi (4.2)

for b,ξb,\xi\in\mathbb{R}, k,σ>0k,\sigma>0 and assume that θθ()=2\theta^{\prime}\theta(\cdot)=\cdot\,^{2} and ρθ()=ρ¯\rho^{\prime}\theta(\cdot)=\overline{\rho}\,\cdot for some constant ρ¯\overline{\rho}\in\mathbb{R} (this holds, for example, d=1d=1 and the state process is the market price of risk). Then, the \mathbb{P}-dynamics of XX given in Eq.(2.5) is

dXt=(baXt)dt+σdBt,X0=ξdX_{t}=(b-aX_{t})\,dt+\sigma\,dB_{t}\,,\;X_{0}=\xi (4.3)

and pT=𝔼[eq0TXu2𝑑u]p_{T}=\mathbb{E}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}^{2}\,du}], where a=kνσρ¯1νa=k-\frac{\nu\sigma\overline{\rho}}{1-\nu} and q=ν(1ν+νρρ)2(1ν)2.q=-\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{2(1-\nu)^{2}}.

We now apply the Hansen–Scheinkman decomposition stated in Section 2.2. Eq.(A.3) gives

pT=f(T,ξ)e12ηξ2ξeλT,\displaystyle p_{T}=f(T,\xi)e^{-\frac{1}{2}\eta\xi^{2}-\ell\xi}e^{-\lambda T}, (4.4)

where α=a2+2qσ2,\alpha=\sqrt{a^{2}+2q\sigma^{2}}, η=αaσ2,\eta=\frac{\alpha-a}{\sigma^{2}}, =bηα,\ell=\frac{b\eta}{\alpha}, λ=12σ22+b+12(αa)\lambda=-\frac{1}{2}\sigma^{2}\ell^{2}+b\ell+\frac{1}{2}(\alpha-a), and

f(t,x)=𝔼x^[e12ηXT2+XT], 0tT,xf(t,x)=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]\,,\;0\leq t\leq T\,,\;x\in\mathbb{R} (4.5)

is the remainder function. Under the measure ^,\hat{\mathbb{P}}, the process XX satisfies

dXt=(δαXt)dt+σdB^tdX_{t}=(\delta-\alpha X_{t})\,dt+\sigma\,d\hat{B}_{t}

for δ=bα.\delta=\frac{b}{\alpha}. It follows that

νlnpT=12ξ2ηνξνTλν+fν(T,ξ)f(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=-\frac{1}{2}\xi^{2}\frac{\partial\eta}{\partial\nu}-\xi\frac{\partial\ell}{\partial\nu}-T\frac{\partial\lambda}{\partial\nu}+\frac{f_{\nu}(T,\xi)}{f(T,\xi)}\,. (4.6)

It suffices to investigate the term fν(T,ξ)f_{\nu}(T,\xi) since the other terms on the right-hand side of Eq.(4.6) are easy to estimate. Since only the parameters η,\eta, ,\ell, α,\alpha, and δ\delta depend on ν\nu in the remainder function f(T,ξ)f(T,\xi), using the chain rule, we know that

fν(T,ξ)=fη(T,ξ)ην+f(T,ξ)ν+fα(T,ξ)αν+fδ(T,ξ)δν.f_{\nu}(T,\xi)=f_{\eta}(T,\xi)\frac{\partial\eta}{\partial\nu}+f_{\ell}(T,\xi)\frac{\partial\ell}{\partial\nu}+f_{\alpha}(T,\xi)\frac{\partial\alpha}{\partial\nu}+f_{\delta}(T,\xi)\frac{\partial\delta}{\partial\nu}\,.

By Proposition A.1, the function f(T,ξ)f(T,\xi) converges to a positive constant as TT\to\infty, and four partial derivatives are bounded in T.T. Therefore,

|1TlnpT+λ|cT and |1TνlnpT+λν|cT\Big{|}\frac{1}{T}\ln p_{T}+\lambda\Big{|}\leq\frac{c}{T}\;\textnormal{ and }\;\Big{|}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\Big{|}\leq\frac{c}{T}

for some positive constant c.c. By using Eq.(2.7), we finally conclude that

|1TνlnuTrρρλ(1ν+ρρν)2+1ν1ν+νρρλν|cT\left|\frac{1}{T}\frac{\partial}{\partial\nu}\ln u_{T}-r-\frac{\rho^{\prime}\rho\lambda}{(1-\nu+\rho^{\prime}\rho\nu)^{2}}+\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c^{\prime}}{T}

for some positive constant cc^{\prime} and

λν\displaystyle\frac{\partial\lambda}{\partial\nu} =(bσ2)ν+12αν12aν\displaystyle=(b-\sigma^{2}\ell)\frac{\partial\ell}{\partial\nu}+\frac{1}{2}\frac{\partial\alpha}{\partial\nu}-\frac{1}{2}\frac{\partial a}{\partial\nu} (4.7)
=(b(bσ2)(a2α3σ1ασ)+σ2(aα1))ρ¯(1ν)2(ab(bσ2)α2σ2+12)(1ν+2ρρν2(1ν)3)\displaystyle=-\Big{(}b(b-\sigma^{2}\ell)\Big{(}\frac{a^{2}}{\alpha^{3}\sigma}-\frac{1}{\alpha\sigma}\Big{)}+\frac{\sigma}{2}\Big{(}\frac{a}{\alpha}-1\Big{)}\Big{)}\frac{\overline{\rho}}{(1-\nu)^{2}}-\Big{(}\frac{ab(b-\sigma^{2}\ell)}{\alpha^{2}\sigma^{2}}+\frac{1}{2}\Big{)}\Big{(}\frac{1-\nu+2\rho^{\prime}\rho\nu}{2(1-\nu)^{3}}\Big{)}

which is obtained by direct calculation.

Remark 4.1.

The optimal expected utility has explicit solutions when the market price of risk is an affine model such as the OU process or the CIR model. However, limTνlnpT\lim_{T\to\infty}\frac{\partial}{\partial\nu}\ln p_{T} is extremely complicated and challenging to calculate from the explicit solutions. In this study, we adopt the Hansen–Scheinkman decomposition and Malliavin calculus so that it is much simpler to calculate the long-term sensitivity using our approach rather than using the explicit solutions.

4.3 CIR model

Under the physical measure 𝐏,{\bf P}, let the state process XX follow the CIR model

dXt=(bkXt)dt+σXtdZt,X0=ξdX_{t}=(b-kX_{t})\,dt+\sigma\sqrt{X_{t}}\,dZ_{t}\,,\;X_{0}=\xi (4.8)

for k,σ,ξ>0,k,\sigma,\xi>0, b>σ2/2b>\sigma^{2}/2 and assume that θθ()=\theta^{\prime}\theta(\cdot)=\cdot and ρθ()=ρ¯\rho^{\prime}\theta(\cdot)=\overline{\rho}\sqrt{\,\cdot\,} for some constant ρ¯\overline{\rho}\in\mathbb{R} (this holds, for example, d=1d=1 and the market price of risk is the square root of the state process). Then, the \mathbb{P}-dynamics of XX given in Eq.(2.5) is

dXt=(baXt)dt+σXtdBt,X0=ξdX_{t}=(b-aX_{t})\,dt+\sigma\sqrt{X_{t}}\,dB_{t}\,,\;X_{0}=\xi (4.9)

and pT=𝔼[eq0TXu𝑑u]p_{T}=\mathbb{E}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}\,du}], where a=kνσρ¯1νa=k-\frac{\nu\sigma\overline{\rho}}{1-\nu} and q=ν(1ν+νρρ)2(1ν)2.q=-\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{2(1-\nu)^{2}}.

We now apply the Hansen–Scheinkman decomposition. Eq.(B.4) gives

pT=f(T,ξ)eηξeλT,\displaystyle p_{T}=f(T,\xi)e^{-\eta\xi}e^{-\lambda T}, (4.10)

where α:=a2+2qσ2,\alpha:=\sqrt{a^{2}+2q\sigma^{2}}, η:=αaσ2,\eta:=\frac{\alpha-a}{\sigma^{2}}, λ:=bη\lambda:=b\eta, and f(t,x):=𝔼x^[eηXT]f(t,x):=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}] is the remainder function. The ^\hat{\mathbb{P}}-dynamics of XX is

dXt=(bαXt)dt+σXtdB^t.dX_{t}=(b-\alpha X_{t})\,dt+\sigma\sqrt{X_{t}}\,d\hat{B}_{t}\,.

It follows that

νlnpT=ξηνTλν+fν(T,ξ)f(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=-\xi\frac{\partial\eta}{\partial\nu}-T\frac{\partial\lambda}{\partial\nu}+\frac{f_{\nu}(T,\xi)}{f(T,\xi)}\,. (4.11)

It suffices to investigate the term fν(T,ξ)f_{\nu}(T,\xi) since the other terms on the right-hand side of the above-mentioned equality are easy to estimate. Since only the parameters η\eta and α\alpha depend on ν\nu in the remainder function, using the chain rule, we know that

fν(T,ξ)=fη(T,ξ)ην+fα(T,ξ)αν.f_{\nu}(T,\xi)=f_{\eta}(T,\xi)\frac{\partial\eta}{\partial\nu}+f_{\alpha}(T,\xi)\frac{\partial\alpha}{\partial\nu}\,.

By Proposition B.1, the function f(T,ξ)f(T,\xi) converges to a positive constant as TT\to\infty, and two partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in T.T. Therefore,

|1TlnpT+λ|cT and |1TνlnpT+λν|cT\Big{|}\frac{1}{T}\ln p_{T}+\lambda\Big{|}\leq\frac{c}{T}\;\textnormal{ and }\;\Big{|}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\Big{|}\leq\frac{c}{T}

for some positive constant c.c. By using Eq.(2.7), we finally conclude that

|1TνlnuTrρρλ(1ν+ρρν)2+1ν1ν+νρρλν|cT\left|\frac{1}{T}\frac{\partial}{\partial\nu}\ln u_{T}-r-\frac{\rho^{\prime}\rho\lambda}{(1-\nu+\rho^{\prime}\rho\nu)^{2}}+\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c^{\prime}}{T}

for some positive constant cc^{\prime} and

λν=bσ((aα1)aν+σ2αqν)=(1aα)bρ¯(1ν)2bσ(1ν+2ρρν)2α(1ν)3\frac{\partial\lambda}{\partial\nu}=\frac{b}{\sigma}\Big{(}\Big{(}\frac{a}{\alpha}-1\Big{)}\frac{\partial a}{\partial\nu}+\frac{\sigma^{2}}{\alpha}\frac{\partial q}{\partial\nu}\Big{)}=\Big{(}1-\frac{a}{\alpha}\Big{)}\frac{b\overline{\rho}}{(1-\nu)^{2}}-\frac{b\sigma(1-\nu+2\rho^{\prime}\rho\nu)}{2\alpha(1-\nu)^{3}}

which is obtained by direct calculation.

4.4 3/23/2 model

Under the physical measure 𝐏,{\bf P}, let the state process XX follow the 3/23/2 model

dXt=(bkXt)Xtdt+σXt3/2dZt,X0=ξdX_{t}=(b-kX_{t})X_{t}\,dt+\sigma X_{t}^{3/2}\,dZ_{t}\,,\;X_{0}=\xi (4.12)

for b,k,σ,ξ>0b,k,\sigma,\xi>0 and assume that θθ()=\theta^{\prime}\theta(\cdot)=\cdot and ρθ()=ρ¯\rho^{\prime}\theta(\cdot)=\overline{\rho}\sqrt{\,\cdot\,} for some constant ρ¯\overline{\rho}\in\mathbb{R} (this holds, for example, d=1d=1 and the market price of risk is the square root of the state process). Then, the \mathbb{P}-dynamics of XX given in Eq.(2.5) is

dXt=(baXt)dt+σXt3/2dBt,X0=ξdX_{t}=(b-aX_{t})\,dt+\sigma X_{t}^{3/2}\,dB_{t}\,,\;X_{0}=\xi (4.13)

and pT=𝔼[eq0TXu𝑑u]p_{T}=\mathbb{E}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}\,du}], where a=kνσρ¯1νa=k-\frac{\nu\sigma\overline{\rho}}{1-\nu} and q=ν(1ν+νρρ)2(1ν)2.q=-\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{2(1-\nu)^{2}}.

We now apply the Hansen–Scheinkman decomposition. Eq.(C.4) gives

pT=f(T,ξ)ξηeλT,\displaystyle p_{T}=f(T,\xi)\xi^{-\eta}e^{-\lambda T}, (4.14)

where η:=(a+σ2/2)2+2qσ2(a+σ2/2)σ2,\eta:=\frac{\sqrt{(a+\sigma^{2}/2)^{2}+2q\sigma^{2}}-(a+\sigma^{2}/2)}{\sigma^{2}}, λ:=bη\lambda:=b\eta, and f(t,x)=𝔼x^[Xtη]f(t,x)=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[X_{t}^{\eta}] is the remainder function. The ^\hat{\mathbb{P}}-dynamics of XX is

dXt=(bαXt)Xtdt+σXt3/2dB^t\displaystyle dX_{t}=(b-\alpha X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\hat{B}_{t} (4.15)

for α:=a+σ2η.\alpha:=a+\sigma^{2}\eta. It follows that

νlnpT=lnξηνTλν+fν(T,ξ)f(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=-\ln\xi\frac{\partial\eta}{\partial\nu}-T\frac{\partial\lambda}{\partial\nu}+\frac{f_{\nu}(T,\xi)}{f(T,\xi)}\,. (4.16)

It suffices to investigate the term fν(T,ξ)f_{\nu}(T,\xi) since the other terms on the right-hand side of the above-mentioned equality are easy to estimate. Since only the parameters η\eta and α\alpha depend on ν\nu in the remainder function, using the chain rule, we know that

fν(T,ξ)=fη(T,ξ)ην+fα(T,ξ)αν.f_{\nu}(T,\xi)=f_{\eta}(T,\xi)\frac{\partial\eta}{\partial\nu}+f_{\alpha}(T,\xi)\frac{\partial\alpha}{\partial\nu}\,.

The function f(T,ξ)f(T,\xi) converges to a positive constant as T,T\to\infty, and two partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT by Proposition C.1. Therefore,

|1TlnpT+λ|cT and |1TνlnpT+λν|cT\Big{|}\frac{1}{T}\ln p_{T}+\lambda\Big{|}\leq\frac{c}{T}\;\textnormal{ and }\;\Big{|}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\Big{|}\leq\frac{c}{T}

for some positive constant c.c. By using Eq.(2.7), we finally conclude that

|1TνlnuTrρρλ(1ν+ρρν)2+1ν1ν+νρρλν|cT\left|\frac{1}{T}\frac{\partial}{\partial\nu}\ln u_{T}-r-\frac{\rho^{\prime}\rho\lambda}{(1-\nu+\rho^{\prime}\rho\nu)^{2}}+\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c^{\prime}}{T}

for some positive constant cc^{\prime} and

λν\displaystyle\frac{\partial\lambda}{\partial\nu} =b(a+σ2/2)2+2qσ2(ηaν+qν)\displaystyle=\frac{b}{\sqrt{(a+\sigma^{2}/2)^{2}+2q\sigma^{2}}}\Big{(}-\eta\frac{\partial a}{\partial\nu}+\frac{\partial q}{\partial\nu}\Big{)} (4.17)
=b(1ν)2(a+σ2/2)2+2qσ2(ησρ¯1ν+2ρρν2(1ν)).\displaystyle=\frac{b}{(1-\nu)^{2}\sqrt{(a+\sigma^{2}/2)^{2}+2q\sigma^{2}}}\Big{(}\eta\sigma\overline{\rho}-\frac{1-\nu+2\rho^{\prime}\rho\nu}{2(1-\nu)}\Big{)}\,.

4.5 Quadratic drift model

Under the physical measure 𝐏,{\bf P}, let the state process XX follow a quadratic drift model

dXt=(bkXt2)dt+σXtdZt,X0=ξdX_{t}=(b-kX_{t}^{2})\,dt+\sigma X_{t}\,dZ_{t}\,,\;X_{0}=\xi\, (4.18)

for b,k,ξ>0,b,k,\xi>0, σ0\sigma\neq 0 and assume that θθ()=2\theta^{\prime}\theta(\cdot)=\cdot\,^{2} and ρθ()=ρ¯\rho^{\prime}\theta(\cdot)=\overline{\rho}\,\cdot for some constant ρ¯\overline{\rho}\in\mathbb{R} (this holds, for example, d=1d=1 and the state process is the market price of risk). There is a unique strong solution to the above SDE (Carr and Willems, 2019, Proposition 2.1). Then, under the measure ,\mathbb{P}, the process XX satisfies

dXt=(baXt2)dt+σXtdBt,X0=ξdX_{t}=(b-aX_{t}^{2})\,dt+\sigma X_{t}\,dB_{t}\,,\;X_{0}=\xi (4.19)

and pT=𝔼[eq0TXu2𝑑u]p_{T}=\mathbb{E}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}^{2}\,du}], where a=kνσρ¯1νa=k-\frac{\nu\sigma\overline{\rho}}{1-\nu} and q=ν(1ν+νρρ)2(1ν)2.q=-\frac{\nu(1-\nu+\nu\rho^{\prime}\rho)}{2(1-\nu)^{2}}.

We now apply the Hansen–Scheinkman decomposition. Eq.(D.2) gives

pT=f(T,ξ)eηξeλT,\displaystyle p_{T}=f(T,\xi)e^{-\eta\xi}e^{-\lambda T}, (4.20)

where α:=a2+2qσ2,\alpha:=\sqrt{a^{2}+2q\sigma^{2}}, η:=αaσ2,\eta:=\frac{\alpha-a}{\sigma^{2}}, λ:=bη\lambda:=b\eta, and f(T,ξ):=𝔼ξ^[eηXT]f(T,\xi):=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}] is the remainder function. The ^\hat{\mathbb{P}}-dynamics of XX is

dXt=(bαXt2)dt+σXtdB^t.\displaystyle dX_{t}=(b-\alpha X_{t}^{2})\,dt+\sigma X_{t}\,d\hat{B}_{t}\,. (4.21)

It follows that

νlnpT=lnξηνTλν+fν(T,ξ)f(T,ξ).\frac{\partial}{\partial\nu}\ln p_{T}=-\ln\xi\frac{\partial\eta}{\partial\nu}-T\frac{\partial\lambda}{\partial\nu}+\frac{f_{\nu}(T,\xi)}{f(T,\xi)}\,. (4.22)

It suffices to investigate the term fν(T,ξ)f_{\nu}(T,\xi) since the other terms on the right-hand side of the above-mentioned equality are easy to estimate. Since only the parameters η\eta and α\alpha depend on ν\nu in the remainder function, using the chain rule, we know that

fν(T,ξ)=fη(T,ξ)ην+fα(T,ξ)αν.f_{\nu}(T,\xi)=f_{\eta}(T,\xi)\frac{\partial\eta}{\partial\nu}+f_{\alpha}(T,\xi)\frac{\partial\alpha}{\partial\nu}\,.

The function f(T,ξ)f(T,\xi) converges to a positive constant as T,T\to\infty, and two partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT by Proposition D.1. Therefore,

|1TlnpT+λ|cT and |1TνlnpT+λν|cT\Big{|}\frac{1}{T}\ln p_{T}+\lambda\Big{|}\leq\frac{c}{T}\;\textnormal{ and }\;\Big{|}\frac{1}{T}\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\Big{|}\leq\frac{c}{T}

for some positive constant c.c. By using Eq.(2.7), we finally conclude that

|1TνlnuTrρρλ(1ν+ρρν)2+1ν1ν+νρρλν|cT\left|\frac{1}{T}\frac{\partial}{\partial\nu}\ln u_{T}-r-\frac{\rho^{\prime}\rho\lambda}{(1-\nu+\rho^{\prime}\rho\nu)^{2}}+\frac{1-\nu}{1-\nu+\nu\rho^{\prime}\rho}\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c^{\prime}}{T}

for some positive constant cc^{\prime} and

λν\displaystyle\frac{\partial\lambda}{\partial\nu} =bα(ηaν+qν)\displaystyle=\frac{b}{\alpha}\Big{(}-\eta\frac{\partial a}{\partial\nu}+\frac{\partial q}{\partial\nu}\Big{)} (4.23)
=bα(1ν)2(ησρ¯1ν+2ρρν2(1ν)).\displaystyle=\frac{b}{\alpha(1-\nu)^{2}}\Big{(}\eta\sigma\overline{\rho}-\frac{1-\nu+2\rho^{\prime}\rho\nu}{2(1-\nu)}\Big{)}\,.

Figure 1 displays comparative analysis between the four models. Two graphs show the partial derivative λν\frac{\partial\lambda}{\partial\nu} as a function of ν\nu and μ,\mu, respectively. The model parameters are given as b=0.16,b=0.16, σ=0.8,\sigma=0.8, k=2k=2 (for the first graph), ν=2\nu=-2 (for the second graph) and ρ=0.5,\rho=-0.5,

Refer to caption
Refer to caption
Figure 1: Comparative analysis between the different models.

5 Conclusion

This study investigated the influence of risk tolerance on the expected utility in the long run. We focused on the power utility function of the form

U(x)=xνU(x)=-x^{\nu}

for ν<0\nu<0, where the parameter ν\nu represents how an investor measures the degree of his/her risk tolerance. We considered utility-maximizing portfolios and demonstrated the influence of small changes in the parameter ν\nu on the expected utility of the portfolios in the long run.

The main methodology for this analysis involved a combination of the Hansen–Scheinkman decomposition and the Malliavin calculus technique. First, we transformed the expected utility into the expectation form

pT=𝔼ξ[e0Tq(Xs)𝑑sh(XT)]p_{T}=\mathbb{E}_{\xi}[e^{-\int_{0}^{T}q(X_{s})\,ds}h(X_{T})]

for some Markov diffusion process X=(Xt)0tTX=(X_{t})_{0\leq t\leq T} with X0=ξX_{0}=\xi and some measurable functions qq and h.h. Using the Hansen–Scheinkman decomposition, the expectation pTp_{T} was written as

pT=ϕ(ξ)eλTf(T,ξ)p_{T}=\phi(\xi)e^{-\lambda T}f(T,\xi)

for a real number λ,\lambda, a positive function ϕ\phi, and a measurable function f,f, which depend on the parameter ν.\nu.

The influence of risk tolerance on the long-term expected utility was obtained from the above-mentioned Hansen–Scheinkman decomposition. Under the condition that 1Tνlnf(T,ξ)\frac{1}{T}\frac{\partial}{\partial\nu}\ln f(T,\xi) is bounded in TT on [0,),[0,\infty), we showed that

|νlnpT+λν|cT\left|\frac{\partial}{\partial\nu}\ln p_{T}+\frac{\partial\lambda}{\partial\nu}\right|\leq\frac{c}{T}

for some positive constant c.c. The influence of risk tolerance is asymptotically equal to the partial derivative of λ-\lambda with respect to ν,\nu, which is the main conclusion of this study. To verify that 1Tνlnf(T,ξ)\frac{1}{T}\frac{\partial}{\partial\nu}\ln f(T,\xi) is bounded in TT on [0,),[0,\infty), the Malliavin calculus method was used under several market models, namely the Ornstein–Uhlenbeck process, the CIR process, the 3/23/2 model, and a quadratic drift model.

Acknowledgments.
Hyungbin Park was supported by the National Research Foundation of Korea (NRF) grants funded by the Ministry of Science and ICT (No. 2017R1A5A1015626, No. 2018R1C1B5085491 and No. 2021R1C1C1011675) and the Ministry of Education (No. 2019R1A6A1A10073437) through the Basic Science Research Program.

Appendix A OU process with quadratic killing rate

Assume that a process XX satisfies

dXt=(baXt)dt+σdBt,X0=ξ,dX_{t}=(b-aX_{t})\,dt+\sigma\,dB_{t}\,,\;X_{0}=\xi, (A.1)

where b,b\in\mathbb{R}, a,σ>0.a,\sigma>0. For q>a22σ2,q>-\frac{a^{2}}{2\sigma^{2}}, consider the expectation

pT:=𝔼ξ[eq0TXu2𝑑u].p_{T}:=\mathbb{E}_{\xi}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}^{2}\,du}]\,.

The corresponding operator is

𝒫Th(x)=𝔼x[eq0TXu2𝑑uh(XT)],\mathcal{P}_{T}h(x)=\mathbb{E}_{x}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}^{2}\,du}h(X_{T})]\,,

and it can be shown that

(λ,ϕ(x)):=(12σ22+b+12(αa),e12ηx2x)(\lambda,\phi(x)):=\Big{(}-\frac{1}{2}\sigma^{2}\ell^{2}+b\ell+\frac{1}{2}(\alpha-a),e^{-\frac{1}{2}\eta x^{2}-\ell x}\Big{)}

is an eigenpair, where

α:=a2+2qσ2,η:=αaσ2,:=bηα.\alpha:=\sqrt{a^{2}+2q\sigma^{2}}\,,\;\eta:=\frac{\alpha-a}{\sigma^{2}}\,,\;\ell:=\frac{b\eta}{\alpha}\,.

The eigen-measure ^\hat{\mathbb{P}} is defined on T\mathcal{F}_{T} as

d^d=e12σ20T(ηXs+)2𝑑sσ0T(ηXs+)𝑑Bs.\frac{d\hat{\mathbb{P}}}{d\mathbb{P}}=e^{-\frac{1}{2}\sigma^{2}\int_{0}^{T}(\eta X_{s}+\ell)^{2}\,ds-\sigma\int_{0}^{T}(\eta X_{s}+\ell)\,dB_{s}}\,. (A.2)

The expectation pTp_{T} can be expressed as

pT\displaystyle p_{T} =𝔼ξ[eq0TXs2𝑑s]\displaystyle=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{s}^{2}\,ds}] (A.3)
=𝔼ξ^[e12ηXT2+XT]e12ηξ2ξeλT=f(T,ξ)e12ηξ2ξeλT,\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]e^{-\frac{1}{2}\eta\xi^{2}-\ell\xi}e^{-\lambda T}=f(T,\xi)e^{-\frac{1}{2}\eta\xi^{2}-\ell\xi}e^{-\lambda T},

where

f(t,x)=𝔼x^[e12ηXT2+XT], 0tT,xf(t,x)=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]\,,\;0\leq t\leq T\,,\;x\in\mathbb{R} (A.4)

is the remainder function. The process

B^t=Bt+σ0t(ηXs+)𝑑s, 0tT\hat{B}_{t}=B_{t}+\sigma\int_{0}^{t}(\eta X_{s}+\ell)\,ds\,,\;0\leq t\leq T

is a ^\hat{\mathbb{P}}-Brownian motion and XX follows

dXt=(δαXt)dt+σdB^tdX_{t}=(\delta-\alpha X_{t})\,dt+\sigma\,d\hat{B}_{t}\,

for δ:=bα.\delta:=\frac{b}{\alpha}.

We study the large-time asymptotic behavior of the sensitivity of the remainder function f.f.

Proposition A.1.

Suppose that XX follows

dXt=(δαXt)dt+σdB^t,X0=ξdX_{t}=(\delta-\alpha X_{t})\,dt+\sigma\,d\hat{B}_{t}\,,\;X_{0}=\xi (A.5)

for δ,ξ\delta,\xi\in\mathbb{R} and α,σ>0.\alpha,\sigma>0. Define

f(T,ξ)=𝔼ξ^[e12ηXT2+XT],T0,f(T,\xi)=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]\,,\;T\geq 0\,,

for η<2ασ2\eta<\frac{2\alpha}{\sigma^{2}} and .\ell\in\mathbb{R}. Then,

f(T,ξ)e12ηx2+x𝑑π(x)f(T,\xi)\to\int e^{\frac{1}{2}\eta x^{2}+\ell x}\,d\pi(x) (A.6)

as TT\to\infty, where π\pi is the invariant distribution of X.X. The partial derivatives fη(T,ξ),f_{\eta}(T,\xi), f(T,ξ),f_{\ell}(T,\xi), fα(T,ξ),f_{\alpha}(T,\xi), and fδ(T,ξ)f_{\delta}(T,\xi) are bounded in TT on [0,).[0,\infty).

Proof.

Observe that the density function of XTX_{T} with X0=ξX_{0}=\xi is

z(x;T):=1ΣT2πe12(xmT)2ΣT2,z(x;T):=\frac{1}{\Sigma_{T}\sqrt{2\pi}}e^{-\frac{1}{2}\frac{(x-m_{T})^{2}}{\Sigma_{T}^{2}}}\,,

where mT:=ξeαT+δα(1eαT)m_{T}:=\xi e^{-\alpha T}+\frac{\delta}{\alpha}(1-e^{-\alpha T}) is the mean and ΣT2:=σ22α(1e2αT)\Sigma_{T}^{2}:=\frac{\sigma^{2}}{2\alpha}(1-e^{-2\alpha T}) is the variance. Then, it is clear that

f(T,ξ)=𝔼ξ^[e12ηXT2+XT]=e12ηx2+xz(x;T)𝑑xe12ηx2+x𝑑π(x)f(T,\xi)=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]=\int_{\mathbb{R}}e^{\frac{1}{2}\eta x^{2}+\ell x}z(x;T)\,dx\to\int_{\mathbb{R}}e^{\frac{1}{2}\eta x^{2}+\ell x}\,d\pi(x)

as TT\to\infty, where

dπ(x):=1πσ2/αe(xδ/α)2σ2/αdx.d\pi(x):=\frac{1}{\sqrt{\pi\sigma^{2}/\alpha}}e^{-\frac{(x-\delta/\alpha)^{2}}{\sigma^{2}/\alpha}}\,dx\,.

This proves Eq.(A.6).

We now show that fη(T,ξ)f_{\eta}(T,\xi) is bounded in TT on [0,).[0,\infty). This is direct from

fη(T,ξ)\displaystyle f_{\eta}(T,\xi) =η𝔼ξ^[e12ηXT2+XT]\displaystyle=\frac{\partial}{\partial\eta}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}] (A.7)
=𝔼ξ^[η(e12ηXT2+XT)]\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\frac{\partial}{\partial\eta}\Big{(}e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}\Big{)}\Big{]}
=12𝔼ξ^[XT2e12ηXT2+XT]12y2e12ηy2+yz(y;)𝑑y\displaystyle=\frac{1}{2}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{T}^{2}e^{\frac{1}{2}\eta X_{T}^{2}+\ell X_{T}}]\to\frac{1}{2}\int_{\mathbb{R}}y^{2}e^{\frac{1}{2}\eta y^{2}+\ell y}z(y;\infty)\,dy

since the limit is a finite number. Using the same method, we can show that f(T,ξ)f_{\ell}(T,\xi) is bounded in TT on [0,).[0,\infty).

We show that fα(T,ξ)f_{\alpha}(T,\xi) is bounded in TT on [0,).[0,\infty). Define H(x)=e12ηx2+xH(x)=e^{\frac{1}{2}\eta x^{2}+\ell x} for notational simplicity. By Propositions 2.2, 2.4, and 2.5, we have

fα(T,ξ)=α𝔼ξ^[H(XT)]\displaystyle f_{\alpha}(T,\xi)=\frac{\partial}{\partial\alpha}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H(X_{T})] =1σ𝔼ξ^[H(XT)0TXs𝑑B^s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}H(X_{T})\int_{0}^{T}X_{s}\,d\hat{B}_{s}\Big{]} (A.8)
=1σ𝔼ξ^[0TDs(H(XT))Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}D_{s}(H(X_{T}))X_{s}\,ds\Big{]}
=1σ𝔼ξ^[0TH(XT)(DsXT)Xs𝑑s].\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})(D_{s}X_{T})X_{s}\,ds\Big{]}\,.

Considering that the Malliavin derivative of XTX_{T} is DsXT=σea(Ts)D_{s}X_{T}=\sigma e^{-a(T-s)} for sT,s\leq T, we have

|fδ(T,ξ)|\displaystyle|f_{\delta}(T,\xi)| eaT𝔼ξ^[0T|H(XT)|Xseas𝑑s]=eaT0T𝔼ξ^[|H(XT)|Xs]eas𝑑s.\displaystyle\leq e^{-aT}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|X_{s}e^{as}\,ds\Big{]}=e^{-aT}\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})|X_{s}]e^{as}\,ds\,. (A.9)

Choose u>1u>1 such that 12ηu<ασ2\frac{1}{2}\eta u<\frac{\alpha}{\sigma^{2}} and v>1v>1 such that 1/u+1/v=1.1/u+1/v=1. Then,

|fδ(T,ξ)|eaT0T𝔼ξ^[|H(XT)|Xs]eas𝑑seaT0T𝔼ξ^[|H(XT)|u]1/u𝔼ξ^[Xsv]1/veas𝑑s.|f_{\delta}(T,\xi)|\leq e^{-aT}\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})|X_{s}]e^{as}\,ds\leq e^{-aT}\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})|^{u}]^{1/u}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{v}]^{1/v}e^{as}\,ds\,.

Using H(x)=e12ηx2+x(ηx+)H^{\prime}(x)=e^{\frac{1}{2}\eta x^{2}+\ell x}(\eta x+\ell) and the density function of XT,X_{T}, it is easy to check that 𝔼ξ^[|H(XT)|u]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})|^{u}] and 𝔼ξ^[Xsv]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{v}] are bounded in TT and s,s, respectively. This gives the desired result. Using the same method, we can show that fδ(T,ξ)f_{\delta}(T,\xi) is bounded in TT on [0,).[0,\infty).

Appendix B CIR model

Let XX be the CIR model given as

dXt=(baXt)dt+σXtdBt,X0=ξ,dX_{t}=(b-aX_{t})\,dt+\sigma\sqrt{X_{t}}\,dB_{t}\,,\;X_{0}=\xi, (B.1)

where a,σ,ξ>0a,\sigma,\xi>0 and 2bσ2.2b\geq\sigma^{2}. For

q>a22σ2,q>-\frac{a^{2}}{2\sigma^{2}}\,, (B.2)

consider the expectation

pT:=𝔼ξ[eq0TXu𝑑u].p_{T}:=\mathbb{E}_{\xi}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}\,du}]\,.

The corresponding operator is

𝒫Th(x)=𝔼x[eq0TXu𝑑uh(XT)].\mathcal{P}_{T}h(x)=\mathbb{E}_{x}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}\,du}h(X_{T})]\,.

Following Qin and Linetsky (2016), we know that

(λ,ϕ(x)):=(bη,eηx)(\lambda,\phi(x)):=(b\eta,e^{-\eta x})

is an eigenpair of this operator, where

α:=a2+2qσ2,η:=αaσ2.\alpha:=\sqrt{a^{2}+2q\sigma^{2}}\,,\;\eta:=\frac{\alpha-a}{\sigma^{2}}\,.

The eigen-measure ^\hat{\mathbb{P}} is defined on T\mathcal{F}_{T} as

d^d=e12σ2η20TXs𝑑sση0TXs𝑑Bs.\frac{d\hat{\mathbb{P}}}{d\mathbb{P}}=e^{-\frac{1}{2}\sigma^{2}\eta^{2}\int_{0}^{T}X_{s}\,ds-\sigma\eta\int_{0}^{T}\sqrt{X_{s}}\,dB_{s}}. (B.3)

It is easy to check that a local martingale (e12σ2η20tXs𝑑sση0tXs𝑑Bs)0tT(e^{-\frac{1}{2}\sigma^{2}\eta^{2}\int_{0}^{t}X_{s}\,ds-\sigma\eta\int_{0}^{t}\sqrt{X_{s}}\,dB_{s}})_{0\leq t\leq T} is a martingale. The expectation pTp_{T} can be expressed as

pT\displaystyle p_{T} =𝔼ξ[eq0TXs𝑑s]=𝔼ξ^[eηXT]eηξeλT=f(T,ξ)eηξeλT,\displaystyle=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{s}\,ds}]=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}]e^{-\eta\xi}e^{-\lambda T}=f(T,\xi)e^{-\eta\xi}e^{-\lambda T}, (B.4)

where

f(t,x)=𝔼x^[eηXt], 0tT,x>0f(t,x)=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[e^{\eta X_{t}}]\,,\;0\leq t\leq T\,,\;x>0 (B.5)

is the remainder function. Note that the expectation f(t,x)<f(t,x)<\infty by Lemma B.2 because η=αaσ2<2ασ2.\eta=\frac{\alpha-a}{\sigma^{2}}<\frac{2\alpha}{\sigma^{2}}. The process

B^t=Bt+ση0tXs𝑑s, 0tT\hat{B}_{t}=B_{t}+\sigma\eta\int_{0}^{t}\sqrt{X_{s}}\,ds\,,\;0\leq t\leq T

is a ^\hat{\mathbb{P}}-Brownian motion and XX follows

dXt=(bαXt)dt+σXtdB^t.dX_{t}=(b-\alpha X_{t})\,dt+\sigma\sqrt{X_{t}}\,d\hat{B}_{t}\,.

We study the large-time asymptotic behavior of the remainder function ff and its sensitivity with respect to the parameters α\alpha and η.\eta.

Proposition B.1.

Suppose that XX follows

dXt=(bαXt)dt+σXtdB^t,X0=ξdX_{t}=(b-\alpha X_{t})\,dt+\sigma\sqrt{X_{t}}\,d\hat{B}_{t}\,,\;X_{0}=\xi (B.6)

for α,σ,ξ>0\alpha,\sigma,\xi>0 and 2bσ2.2b\geq\sigma^{2}. Define

f(T,ξ)=𝔼ξ^[eηXT],T0,f(T,\xi)=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}]\,,\;T\geq 0\,,

for η<2ασ2.\eta<\frac{2\alpha}{\sigma^{2}}. Then,

f(T,ξ)eηx𝑑π(x)f(T,\xi)\to\int e^{\eta x}\,d\pi(x) (B.7)

as TT\to\infty, where π\pi is the invariant distribution of X.X. The partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT on [0,).[0,\infty).

Proof.

It is easy to prove Eq.(B.7) by considering the density function of X;X; thus, we omit the proof. Consider the partial derivative fη(T,ξ).f_{\eta}(T,\xi). Choose any γ\gamma with η=αaσ2<γ<ασ2,\eta=\frac{\alpha-a}{\sigma^{2}}<\gamma<\frac{\alpha}{\sigma^{2}}, then there is a positive constant cγc_{\gamma} such that eηxxcγeγxe^{\eta x}x\leq c_{\gamma}e^{\gamma x} for x>0.x>0. Observe that

fη(T,ξ)=η𝔼ξ^[eηXT]=𝔼ξ^[ηeηXT]=𝔼ξ^[eηXTXT];f_{\eta}(T,\xi)=\frac{\partial}{\partial\eta}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}]=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\frac{\partial}{\partial\eta}e^{\eta X_{T}}\Big{]}=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}X_{T}]\,;

thus,

|fη(T,ξ)|𝔼ξ^[eηXTXT]cγ𝔼ξ^[eγXT].|f_{\eta}(T,\xi)|\leq\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}X_{T}]\leq c_{\gamma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\gamma X_{T}}]\,.

By Lemma B.2, the expectation 𝔼ξ^[eγXT]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\gamma X_{T}}] is bounded in TT on [0,)[0,\infty), and this gives the desired result.

Now, we show that the partial derivative fα(T,ξ)f_{\alpha}(T,\xi) is bounded in TT on [0,).[0,\infty). Define H(x)=eηxH(x)=e^{\eta x} for notational simplicity. By Propositions 2.2, 2.4, and 2.5, we have

fα(T,ξ)=α𝔼ξ^[H(XT)]\displaystyle f_{\alpha}(T,\xi)=\frac{\partial}{\partial\alpha}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H(X_{T})] =1σ𝔼ξ^[H(XT)0TXs𝑑B^s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}H(X_{T})\int_{0}^{T}\sqrt{X_{s}}\,d\hat{B}_{s}\Big{]} (B.8)
=1σ𝔼ξ^[0TDs(H(XT))Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}D_{s}(H(X_{T}))\sqrt{X_{s}}\,ds\Big{]}
=1σ𝔼ξ^[0TH(XT)(DsXT)Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})(D_{s}X_{T})\sqrt{X_{s}}\,ds\Big{]}
=𝔼ξ^[0TH(XT)esT(α2(b2σ28)1Xu)𝑑uXTXs𝑑s].\displaystyle=-\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})e^{\int_{s}^{T}\big{(}-\frac{\alpha}{2}-(\frac{b}{2}-\frac{\sigma^{2}}{8})\frac{1}{X_{u}}\big{)}\,du}\sqrt{X_{T}}\sqrt{X_{s}}\,ds\Big{]}\,.

The last equality is from

DsXT=σesT(α2(b2σ28)1Xu)𝑑uXT,D_{s}X_{T}=\sigma e^{\int_{s}^{T}\big{(}-\frac{\alpha}{2}-(\frac{b}{2}-\frac{\sigma^{2}}{8})\frac{1}{X_{u}}\big{)}\,du}\sqrt{X_{T}}\,,

which is obtained by DsXT=σXsYTYsD_{s}X_{T}=\sigma\sqrt{X_{s}}\frac{Y_{T}}{Y_{s}} for the first variation process YY of X.X. This can be obtained from the work of Alòs and Ewald (2008) (note that Proposition 2.3 cannot be applied here because the coefficients in Eq.(B.6) do not have bounded derivatives). Then,

|fα(T,ξ)|\displaystyle|f_{\alpha}(T,\xi)| 𝔼ξ^[0T|H(XT)|esT(α2(b2σ28)1Xu)𝑑uXTXs𝑑s]\displaystyle\leq\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|e^{\int_{s}^{T}\big{(}-\frac{\alpha}{2}-(\frac{b}{2}-\frac{\sigma^{2}}{8})\frac{1}{X_{u}}\big{)}\,du}\sqrt{X_{T}}\sqrt{X_{s}}\,ds\Big{]} (B.9)
𝔼ξ^[0T|H(XT)|eα2(Ts)XTXsds](b2σ28>0)\displaystyle\leq\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|e^{-\frac{\alpha}{2}(T-s)}\sqrt{X_{T}}\sqrt{X_{s}}\,ds\Big{]}\quad\Big{(}\because\frac{b}{2}-\frac{\sigma^{2}}{8}>0\Big{)}
0T𝔼ξ^[|H(XT)|XTXs]eα2(Ts)𝑑s\displaystyle\leq\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})|\sqrt{X_{T}}\sqrt{X_{s}}]e^{-\frac{\alpha}{2}(T-s)}\,ds
0T𝔼ξ^[(H(XT))2XT]1/2𝔼ξ^[Xs]1/2eα2(Ts)𝑑s.\displaystyle\leq\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[(H^{\prime}(X_{T}))^{2}X_{T}]^{1/2}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}]^{1/2}e^{-\frac{\alpha}{2}(T-s)}\,ds.

Since (H(XT))2XT=η2XTe2ηXT(H^{\prime}(X_{T}))^{2}X_{T}=\eta^{2}X_{T}e^{2\eta X_{T}} and 2η<2ασ2,2\eta<\frac{2\alpha}{\sigma^{2}}, by Lemma B.2, the expectation 𝔼ξ^[(H(XT))2XT]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[(H^{\prime}(X_{T}))^{2}X_{T}] is bounded in TT on [0,).[0,\infty). It is clear that 𝔼ξ^[Xs]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}] is bounded in ss on [0,).[0,\infty). Thus,

|fα(T,ξ)|c0Teα2(Ts)𝑑s2cα\displaystyle|f_{\alpha}(T,\xi)|\leq c\int_{0}^{T}e^{-\frac{\alpha}{2}(T-s)}\,ds\leq\frac{2c}{\alpha} (B.10)

for some positive constant c,c, which gives the desired result. ∎

Lemma B.2.

Let XX be a solution of

dXt=(bαXt)dt+σXtdBt,X0=x,dX_{t}=(b-\alpha X_{t})\,dt+\sigma\sqrt{X_{t}}\,d{B}_{t}\,,\;X_{0}=x\,,

where α,σ,x>0\alpha,\sigma,x>0 and 2b>σ2.2b>\sigma^{2}. For γ<2α/σ2\gamma<2\alpha/\sigma^{2}, we have

𝔼x[eγXT]=(11γc(T))2b/σ2eγ1γc(T)eαTx,\mathbb{E}_{x}^{\mathbb{P}}[e^{\gamma X_{T}}]=\Big{(}\frac{1}{1-\gamma c(T)}\Big{)}^{2b/\sigma^{2}}e^{\frac{\gamma}{1-\gamma c(T)}e^{-\alpha T}x}\,,

where c(T):=σ22α(1eαT).c(T):=\frac{\sigma^{2}}{2\alpha}(1-e^{-\alpha T}).

Proof.

See Corollary 6.3.4.4 in Jeanblanc et al. (2009), where the proof is given for γ<0\gamma<0; the same proof holds for γ<2α/σ2.\gamma<2\alpha/\sigma^{2}.

Appendix C 3/23/2 model

Consider the 3/23/2 model

dXt=(baXt)Xtdt+σXt3/2dBt,X0=ξdX_{t}=(b-aX_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,dB_{t}\,,\;X_{0}=\xi

where b,σ,ξ>0b,\sigma,\xi>0, and a>σ22.a>-\frac{\sigma^{2}}{2}. For

q>12σ2(a+σ22)2+σ28,q>-\frac{1}{2\sigma^{2}}\Big{(}a+\frac{\sigma^{2}}{2}\Big{)}^{2}+\frac{\sigma^{2}}{8}\,, (C.1)

define

η:=(a+σ2/2)2+2qσ2(a+σ2/2)σ2.\eta:=\frac{\sqrt{(a+\sigma^{2}/2)^{2}+2q\sigma^{2}}-(a+\sigma^{2}/2)}{\sigma^{2}}\,.

Then, it is easy to check that α:=a+σ2η>0.\alpha:=a+\sigma^{2}\eta>0.

We apply the Hansen–Scheinkman decomposition to estimate the expectation

pT=𝔼[eq0TXs𝑑s].p_{T}=\mathbb{E}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{s}\,ds}]\,.

The corresponding operator is

𝒫Th(x)=𝔼x[eq0TXu𝑑uh(XT)],\mathcal{P}_{T}h(x)=\mathbb{E}_{x}^{{\mathbb{P}}}[e^{-q\int_{0}^{T}X_{u}\,du}h(X_{T})]\,,

and it can be shown that (λ,ϕ(x)):=(bη,xη)(\lambda,\phi(x)):=(b\eta,x^{-\eta}) is an eigenpair. The eigen-measure ^\hat{\mathbb{P}} is defined on T\mathcal{F}_{T} as

d^d=e12σ2η20TXs𝑑sση0TXs𝑑Bs.\frac{d\hat{\mathbb{P}}}{d\mathbb{P}}=e^{-\frac{1}{2}\sigma^{2}\eta^{2}\int_{0}^{T}X_{s}\,ds-\sigma\eta\int_{0}^{T}\sqrt{X_{s}}\,dB_{s}}\,. (C.2)

The process

B^t=ση0tXs𝑑s+Bt, 0tT\hat{B}_{t}=\sigma\eta\int_{0}^{t}\sqrt{X_{s}}\,ds+B_{t}\,,\;0\leq t\leq T

is a Brownian motion by the Girsanov theorem, and XX follows

dXt=(bαXt)Xtdt+σXt3/2dB^t\displaystyle dX_{t}=(b-\alpha X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\hat{B}_{t} (C.3)

for α=a+σ2η.\alpha=a+\sigma^{2}\eta.

The expectation pTp_{T} can be expressed as

pT\displaystyle p_{T} =𝔼ξ[eq0TXs𝑑s]=𝔼ξ^[XTη]ξηeλT=f(T,ξ)ξηeλT,\displaystyle=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{s}\,ds}]=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{T}^{\eta}]\xi^{-\eta}e^{-\lambda T}=f(T,\xi)\xi^{-\eta}e^{-\lambda T}, (C.4)

where

f(t,x)=𝔼x^[Xtη], 0tT,x>0f(t,x)=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[X_{t}^{\eta}]\,,\;0\leq t\leq T\,,x>0 (C.5)

is the remainder function. Considering that 1/X1/X is a CIR model and hh has linear growth, the function f(T,ξ)f(T,\xi) converges to a constant as T.T\to\infty.

We study the large-time asymptotic behavior of the remainder function ff and its sensitivity with respect to the parameters α\alpha and η.\eta.

Proposition C.1.

Suppose that XX follows

dXt=(bαXt)Xtdt+σXt3/2dB^t,X0=ξdX_{t}=(b-\alpha X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\hat{B}_{t}\,,\;X_{0}=\xi (C.6)

for b,α,ξ>0,b,\alpha,\xi>0, σ0.\sigma\neq 0. Define

f(T,ξ)=𝔼ξ^[XTη],T0,f(T,\xi)=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{T}^{\eta}]\,,\;T\geq 0\,,

for η<2ασ2+1.\eta<\frac{2\alpha}{\sigma^{2}}+1. Then,

f(T,ξ)xη𝑑π(x)f(T,\xi)\to\int x^{\eta}\,d\pi(x) (C.7)

as TT\to\infty, where π\pi is the invariant distribution of X.X. The partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT on [0,).[0,\infty).

Proof.

It is easy to prove Eq.(C.7) by considering the density function of X.X. We prove that two partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT on [0,).[0,\infty). First, consider the partial derivative fη(T,ξ).f_{\eta}(T,\xi). Observe that

fη(T,ξ)=η𝔼ξ^[XTη]=𝔼ξ^[ηXTη]=𝔼ξ^[XTηlnXT].f_{\eta}(T,\xi)=\frac{\partial}{\partial\eta}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{T}^{\eta}]=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\frac{\partial}{\partial\eta}X_{T}^{\eta}\Big{]}=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{T}^{\eta}\ln X_{T}]\,.

There are constants c>0,c>0, γ<2ασ2+2\gamma<\frac{2\alpha}{\sigma^{2}}+2, and mm\in\mathbb{N} such that

xηlnxcxγ,x1x^{\eta}\ln x\leq cx^{\gamma}\,,\;x\geq 1

and

xη|lnx|cxm, 0<x<1.x^{\eta}|\ln x|\leq cx^{-m}\,,\;0<x<1\,.

Thus, by Lemma C.2, the partial derivative fη(T,ξ)f_{\eta}(T,\xi) is bounded in TT on [0,).[0,\infty).

Now we show that the partial derivative fα(T,ξ)f_{\alpha}(T,\xi) is bounded in TT on [0,).[0,\infty). Define H(x)=xηH(x)=x^{\eta} for notational simplicity. By Propositions 2.2, 2.4, and 2.5, we have

fα(T,ξ)=α𝔼ξ^[H(XT)]\displaystyle f_{\alpha}(T,\xi)=\frac{\partial}{\partial\alpha}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H(X_{T})] =1σ𝔼ξ^[H(XT)0TXs𝑑B^s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}H(X_{T})\int_{0}^{T}\sqrt{X_{s}}\,d\hat{B}_{s}\Big{]} (C.8)
=1σ𝔼ξ^[0TDs(H(XT))Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}D_{s}(H(X_{T}))\sqrt{X_{s}}\,ds\Big{]}
=1σ𝔼ξ^[0TH(XT)(DsXT)Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})(D_{s}X_{T})\sqrt{X_{s}}\,ds\Big{]}
=𝔼ξ^[0TH(XT)XT3/2eb2(Ts)e(α2+3σ28)sTXu𝑑uXs𝑑s].\displaystyle=-\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})X_{T}^{3/2}e^{-\frac{b}{2}(T-s)}e^{-(\frac{\alpha}{2}+\frac{3\sigma^{2}}{8})\int_{s}^{T}X_{u}\,du}\sqrt{X_{s}}\,ds\Big{]}\,.

For the last equality, we used that the Malliavin derivative of XTX_{T} is

DsXT=σeb2(Ts)e(α2+3σ28)sTXu𝑑uXT3/2.\displaystyle D_{s}X_{T}=\sigma e^{-\frac{b}{2}(T-s)}e^{-(\frac{\alpha}{2}+\frac{3\sigma^{2}}{8})\int_{s}^{T}X_{u}\,du}X_{T}^{3/2}\,. (C.9)

From

XT=Xseb2(Ts)(α2+σ24)sTXu𝑑u+σ2sTXu1/2𝑑B^u,\displaystyle\sqrt{X_{T}}=\sqrt{X_{s}}e^{\frac{b}{2}(T-s)-(\frac{\alpha}{2}+\frac{\sigma^{2}}{4})\int_{s}^{T}X_{u}\,du+\frac{\sigma}{2}\int_{s}^{T}X_{u}^{1/2}\,d\hat{B}_{u}}\,, (C.10)

we have

fα(T,ξ)\displaystyle f_{\alpha}(T,\xi) =𝔼ξ^[0TH(XT)XTe(α+5σ28)sTXu𝑑u+σ2sTXu1/2𝑑B^uXs𝑑s]\displaystyle=-\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})X_{T}e^{-(\alpha+\frac{5\sigma^{2}}{8})\int_{s}^{T}X_{u}\,du+\frac{\sigma}{2}\int_{s}^{T}X_{u}^{1/2}\,d\hat{B}_{u}}X_{s}\,ds\Big{]} (C.11)
=𝔼ξ^[0T𝔼ξ^[H(XT)XTe(α+5σ28)sTXu𝑑u+σ2sTXu1/2𝑑B^u|Xs]Xs2𝑑s]\displaystyle=-\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H^{\prime}(X_{T})X_{T}e^{-(\alpha+\frac{5\sigma^{2}}{8})\int_{s}^{T}X_{u}\,du+\frac{\sigma}{2}\int_{s}^{T}X_{u}^{1/2}\,d\hat{B}_{u}}|X_{s}]X_{s}^{2}\,ds\Big{]}
=𝔼ξ^[0Tg(Ts,Xs)Xs2𝑑s],\displaystyle=-\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}g(T-s,X_{s})X_{s}^{2}\,ds\Big{]},

where gg is defined as

g(t,x)\displaystyle g(t,x) =𝔼x^[H(Xt)Xte(α+5σ28)0tXu𝑑u+σ20tXu1/2𝑑B^u]\displaystyle=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[H^{\prime}(X_{t})X_{t}e^{-(\alpha+\frac{5\sigma^{2}}{8})\int_{0}^{t}X_{u}\,du+\frac{\sigma}{2}\int_{0}^{t}X_{u}^{1/2}\,d\hat{B}_{u}}] (C.12)
=𝔼ξ^[H(Xt+s)Xt+se(α+5σ28)st+sXu𝑑u+σ2st+sXu1/2𝑑B^u|Xs=x].\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H^{\prime}(X_{t+s})X_{t+s}e^{-(\alpha+\frac{5\sigma^{2}}{8})\int_{s}^{t+s}X_{u}\,du+\frac{\sigma}{2}\int_{s}^{t+s}X_{u}^{1/2}\,d\hat{B}_{u}}|X_{s}=x]\,.

The second equality is from the Markov property.

We aim to estimate this function g.g. Define a measure ~\tilde{\mathbb{P}} on T\mathcal{F}_{T} as

d~d^=eσ280TXu𝑑u+σ20TXu1/2𝑑B^u.\frac{d\tilde{\mathbb{P}}}{d\hat{\mathbb{P}}}=e^{-\frac{\sigma^{2}}{8}\int_{0}^{T}X_{u}\,du+\frac{\sigma}{2}\int_{0}^{T}X_{u}^{1/2}\,d\hat{B}_{u}}\,.

It is easy to show that a local martingale (eσ280tXu𝑑u+σ20tXu1/2𝑑B^u)0tT(e^{-\frac{\sigma^{2}}{8}\int_{0}^{t}X_{u}\,du+\frac{\sigma}{2}\int_{0}^{t}X_{u}^{1/2}\,d\hat{B}_{u}})_{0\leq t\leq T} is a martingale. The process XX satisfies

dXt=(b(α12σ2)Xt)Xtdt+σXt3/2dB~t,dX_{t}=(b-(\alpha-\frac{1}{2}\sigma^{2})X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\tilde{B}_{t}\,,

where (B~t)0tT(\tilde{B}_{t})_{0\leq t\leq T} is a ~\tilde{\mathbb{P}}-Brownian motion. Note that since the mean-reversion speed α12σ2>12σ2,\alpha-\frac{1}{2}\sigma^{2}>-\frac{1}{2}\sigma^{2}, the process stays positive under the measure ~.\tilde{\mathbb{P}}. It follows that

g(t,x)=𝔼x~[H(Xt)Xte(α+σ22)0tXu𝑑u].g(t,x)=\mathbb{E}_{x}^{\tilde{\mathbb{P}}}[H^{\prime}(X_{t})X_{t}e^{-(\alpha+\frac{\sigma^{2}}{2})\int_{0}^{t}X_{u}\,du}]\,.

We apply the Hansen–Scheinkman decomposition here. Consider the operator

h𝔼x~[h(Xt)e(α+σ22)0tXu𝑑u].h\mapsto\mathbb{E}_{x}^{\tilde{\mathbb{P}}}[h(X_{t})e^{-(\alpha+\frac{\sigma^{2}}{2})\int_{0}^{t}X_{u}\,du}]\,.

It can be shown that (λ~,ϕ~(x)):=(b,x1)(\tilde{\lambda},\tilde{\phi}(x)):=(b,x^{-1}) is an eigenpair, and let ¯\overline{\mathbb{P}} be the corresponding eigen-measure. The ¯\overline{\mathbb{P}}-dynamics of XX is

dXt=(b(α+12σ2)Xt)Xtdt+σXt3/2dB¯t,dX_{t}=(b-(\alpha+\frac{1}{2}\sigma^{2})X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\overline{B}_{t}\,,

where (B¯t)0tT(\overline{B}_{t})_{0\leq t\leq T} is a ¯\overline{\mathbb{P}}-Brownian motion. Then,

g(t,x)=𝔼x~[H(Xt)Xte(α+σ22)0tXu𝑑u]=𝔼x¯[H(Xt)Xt2]ebtx1.g(t,x)=\mathbb{E}_{x}^{\tilde{\mathbb{P}}}[H^{\prime}(X_{t})X_{t}e^{-(\alpha+\frac{\sigma^{2}}{2})\int_{0}^{t}X_{u}\,du}]=\mathbb{E}_{x}^{\overline{\mathbb{P}}}[H^{\prime}(X_{t})X_{t}^{2}]e^{-bt}x^{-1}\,.

Since |H(x)|x2=ηxη+1|H^{\prime}(x)|x^{2}=\eta x^{\eta+1} and η+12ασ2+3\eta+1\leq\frac{2\alpha}{\sigma^{2}}+3 holds, by Lemma C.2 (with α\alpha replaced by α+12σ2\alpha+\frac{1}{2}\sigma^{2}), the expectation 𝔼x¯[|H(Xt)|Xt2]\mathbb{E}_{x}^{\overline{\mathbb{P}}}[|H^{\prime}(X_{t})|X_{t}^{2}] is uniformly bounded in (t,x)(t,x) on [0,)×(0,).[0,\infty)\times(0,\infty). Thus,

|g(t,x)|cebtx1|g(t,x)|\leq c^{\prime}e^{-bt}x^{-1}

for some positive constant c,c^{\prime}, which is independent of tt and x.x. Eq.(C.11) gives

|fα(T,ξ)|\displaystyle|f_{\alpha}(T,\xi)| 𝔼ξ^[0T|g(Ts,Xs)|eb(Ts)Xs2𝑑s]c0Teb(Ts)𝔼ξ^[Xs2]𝑑s.\displaystyle\leq\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|g(T-s,X_{s})|e^{-b(T-s)}X_{s}^{2}\,ds\Big{]}\leq c^{\prime}\int_{0}^{T}e^{-b(T-s)}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{2}]\,ds\,. (C.13)

By Lemma C.2, the expectation 𝔼ξ^[Xs2]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{2}] is bounded in ss on [0,).[0,\infty). This gives the desired result. ∎

Lemma C.2.

Let XX be a solution of

dXt=(bαXt)Xtdt+σXt3/2dB^t,X0=ξdX_{t}=(b-\alpha X_{t})X_{t}\,dt+\sigma{X_{t}}^{3/2}\,d\hat{B}_{t}\,,\;X_{0}=\xi

for b,σ,ξ>0b,\sigma,\xi>0, and ασ22.\alpha\geq-\frac{\sigma^{2}}{2}. Then, for A<2ασ2+2,A<\frac{2\alpha}{\sigma^{2}}+2,

H(T,ξ):=𝔼ξ(XTA)\displaystyle H(T,\xi):=\mathbb{E}_{\xi}(X_{T}^{A}) =Γ(2ασ2+2A)Γ(2ασ2+2)(2bσ211ebT)AF(A,2ασ2+2,2bσ21(ebT1)ξ),\displaystyle=\frac{\Gamma(\frac{2\alpha}{\sigma^{2}}+2-A)}{\Gamma(\frac{2\alpha}{\sigma^{2}}+2)}\Big{(}\frac{2b}{\sigma^{2}}\frac{1}{1-e^{-bT}}\Big{)}^{A}F\Big{(}A,\frac{2\alpha}{\sigma^{2}}+2,-\frac{2b}{\sigma^{2}}\frac{1}{(e^{bT}-1)\xi}\Big{)}\,, (C.14)

where FF is the confluent hypergeometric function. The function H(T,ξ)H(T,\xi) converges to

Γ(2ασ2+2A)Γ(2ασ2+2)(2bσ2)A\frac{\Gamma(\frac{2\alpha}{\sigma^{2}}+2-A)}{\Gamma(\frac{2\alpha}{\sigma^{2}}+2)}\Big{(}\frac{2b}{\sigma^{2}}\Big{)}^{A}

as T.T\to\infty. Moreover, if 0<A<2ασ2+2,0<A<\frac{2\alpha}{\sigma^{2}}+2, then the function HH is uniformly bounded on the domain [0,)×(0,).[0,\infty)\times(0,\infty).

See (Park, 2019, Lemma B.1) for the proof.

Appendix D Quadratic drift model

Assume that XX follows

dXt=(baXt2)dt+σXtdBt,X0=ξdX_{t}=(b-aX_{t}^{2})\,dt+\sigma X_{t}\,dB_{t}\,,\;X_{0}=\xi (D.1)

for b,a,σ,ξ>0.b,a,\sigma,\xi>0. This SDE has a unique strong solution and the solution stays positive by (Carr and Willems, 2019, Proposition 2.1). For q>a22σ2,q>-\frac{a^{2}}{2\sigma^{2}}, we define

pT:=𝔼ξ[eq0TXu2𝑑u].p_{T}:=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{u}^{2}\,du}]\,.

Consider the generator

(ϕ)(x):=12σ2x2ϕ′′(x)+(bax2)ϕ(x)qx2ϕ(x).(\mathcal{L}\phi)(x):=\frac{1}{2}\sigma^{2}x^{2}\phi^{\prime\prime}(x)+(b-ax^{2})\phi^{\prime}(x)-qx^{2}\phi(x)\,.

It can be shown that (λ,ϕ):=(bη,eηx)(\lambda,\phi):=(b\eta,e^{-\eta x}) is an eigenpair, where

α:=a2+2qσ2,η:=αaσ2.\alpha:=\sqrt{a^{2}+2q\sigma^{2}}\,,\;\eta:=\frac{\alpha-a}{\sigma^{2}}\,.

Let ^\hat{\mathbb{P}} be the eigen-measure on T\mathcal{F}_{T} defined as

d^d=e12σ2η20TXs2𝑑sση0TXs𝑑Bs.\frac{d\hat{\mathbb{P}}}{d\mathbb{P}}=e^{-\frac{1}{2}\sigma^{2}\eta^{2}\int_{0}^{T}X_{s}^{2}\,ds-\sigma\eta\int_{0}^{T}X_{s}\,dB_{s}}\,.

Then, the ^\hat{\mathbb{P}}-dynamics of XX is

dXt=(bαXt2)dt+σXtdB^t,X0=ξdX_{t}=(b-\alpha X_{t}^{2})\,dt+\sigma X_{t}\,d\hat{B}_{t}\,,\;X_{0}=\xi

for a ^\hat{\mathbb{P}}-Brownian motion (B^t)0tT.(\hat{B}_{t})_{0\leq t\leq T}. It can be easily checked that the local martingale (e12σ2η20tXs2𝑑sση0tXs𝑑Bs)0tT(e^{-\frac{1}{2}\sigma^{2}\eta^{2}\int_{0}^{t}X_{s}^{2}\,ds-\sigma\eta\int_{0}^{t}X_{s}\,dB_{s}})_{0\leq t\leq T} is a martingale. It follows that

pT:=𝔼ξ[eq0TXs2𝑑s]=f(T,ξ)eηξeλT,p_{T}:=\mathbb{E}_{\xi}^{\mathbb{P}}[e^{-q\int_{0}^{T}X_{s}^{2}\,ds}]=f(T,\xi)e^{-\eta\xi}e^{-\lambda T}, (D.2)

where

f(t,x):=𝔼x^[eηXt].f(t,x):=\mathbb{E}_{x}^{\hat{\mathbb{P}}}[e^{\eta X_{t}}]\,.

The invariant measure of XX under the measure ^\hat{\mathbb{P}} is

dπ(x)=1σ2x2e2bσ21x2ασ2xdxd\pi(x)=\frac{1}{\sigma^{2}x^{2}}e^{-\frac{2b}{\sigma^{2}}\frac{1}{x}-\frac{2\alpha}{\sigma^{2}}x}\,dx

up to positive constant multiples. For further details on the invariant measure, readers may refer to (Löcherbach, 2015, Proposition 3.2) or (Kallenberg, 2006, Lemma 20.19).

Proposition D.1.

Suppose that XX follows

dXt=(bαXt2)dt+σXtdB^t,X0=ξdX_{t}=(b-\alpha X_{t}^{2})\,dt+\sigma X_{t}\,d\hat{B}_{t}\,,\;X_{0}=\xi (D.3)

for b,α,ξ>0,b,\alpha,\xi>0, σ0.\sigma\neq 0. Define

f(T,ξ):=𝔼ξ^[eηXT],f(T,\xi):=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[e^{\eta X_{T}}]\,,

for η<ασ2.\eta<\frac{\alpha}{\sigma^{2}}. Then,

f(T,ξ)eηx𝑑π(x)f(T,\xi)\to\int e^{\eta x}\,d\pi(x) (D.4)

as TT\to\infty, where π\pi is the invariant distribution of X.X. The partial derivatives fη(T,ξ)f_{\eta}(T,\xi) and fα(T,ξ)f_{\alpha}(T,\xi) are bounded in TT on [0,).[0,\infty).

Proof.

Since eηxL2(m),e^{\eta x}\in L^{2}(m), by Proposition (D.2), we have Eq.(D.4). By the same method in Eq.(A.7), it follows that fη(T,ξ)f_{\eta}(T,\xi) is bounded in T.T. We now show that the partial derivative fα(T,ξ)f_{\alpha}(T,\xi) is bounded in TT on [0,).[0,\infty). Define H(x)=eηxH(x)=e^{\eta x} for notational simplicity. We have

fα(T,ξ)=α𝔼ξ^[H(XT)]\displaystyle f_{\alpha}(T,\xi)=\frac{\partial}{\partial\alpha}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[H(X_{T})] =1σ𝔼ξ^[H(XT)0TXs𝑑B^s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}H(X_{T})\int_{0}^{T}X_{s}\,d\hat{B}_{s}\Big{]} (D.5)
=1σ𝔼ξ^[0TDs(H(XT))Xs𝑑s]\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}D_{s}(H(X_{T}))X_{s}\,ds\Big{]}
=1σ𝔼ξ^[0TH(XT)(DsXT)Xs𝑑s].\displaystyle=-\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}H^{\prime}(X_{T})(D_{s}X_{T})X_{s}\,ds\Big{]}\,.

By Proposition D.3, the Malliavin derivative of XTX_{T} is

DsXT=σXse2αsTXs𝑑s12σ2(Ts)+σ(B^TB^s)=σXTesT(bXs+αXs)𝑑s, 0sT.D_{s}X_{T}=\sigma X_{s}e^{-2\alpha\int_{s}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-s)+\sigma(\hat{B}_{T}-\hat{B}_{s})}=\sigma X_{T}e^{-\int_{s}^{T}(\frac{b}{X_{s}}+\alpha X_{s})\,ds}\,,\;0\leq s\leq T\,.

The last inequality is from XT/Xs=esT(bXsαXs)𝑑s12σ2(Ts)+σ(B^TB^s).X_{T}/X_{s}=e^{\int_{s}^{T}(\frac{b}{X_{s}}-\alpha X_{s})\,ds-\frac{1}{2}\sigma^{2}(T-s)+\sigma(\hat{B}_{T}-\hat{B}_{s})}. Thus,

|fα(T,ξ)|\displaystyle|f_{\alpha}(T,\xi)| 1σ𝔼ξ^[0T|H(XT)|(DsXT)Xs𝑑s]\displaystyle\leq\frac{1}{\sigma}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|(D_{s}X_{T})X_{s}\,ds\Big{]} (D.6)
=𝔼ξ^[0T|H(XT)|XTesT(bXs+αXs)𝑑sXs𝑑s]\displaystyle=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|X_{T}e^{-\int_{s}^{T}(\frac{b}{X_{s}}+\alpha X_{s})\,ds}X_{s}\,ds\Big{]}
𝔼ξ^[0T|H(XT)|XTe2αb(Ts)Xsds](bXs+αXs2αb).\displaystyle\leq\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}\Big{[}\int_{0}^{T}|H^{\prime}(X_{T})|X_{T}e^{-2\sqrt{\alpha b}(T-s)}X_{s}\,ds\Big{]}\quad\Big{(}\because\frac{b}{X_{s}}+\alpha X_{s}\geq 2\sqrt{\alpha b}\Big{)}\,.

Choose constants pp and qq such that 1<p<ααa1<p<\frac{\alpha}{\alpha-a} and 1/p+1/q=1.1/p+1/q=1. Then,

|fα(T,ξ)|\displaystyle|f_{\alpha}(T,\xi)| 0T𝔼ξ^[|H(XT)XT|p]1/p𝔼ξ^[Xsq]1/qe2αb(Ts)ds].\displaystyle\leq\int_{0}^{T}\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})X_{T}|^{p}]^{1/p}\,\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{q}]^{1/q}e^{-2\sqrt{\alpha b}(T-s)}\,ds\Big{]}\,. (D.7)

Observe that |H(x)x|p=ηpxpepηxL2(m)|H^{\prime}(x)x|^{p}=\eta^{p}x^{p}e^{p\eta x}\in L^{2}(m) since pη<ασ2.p\eta<\frac{\alpha}{\sigma^{2}}. By Proposition D.2, the expectation 𝔼ξ^[|H(XT)XT|p]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[|H^{\prime}(X_{T})X_{T}|^{p}] converges to |H(x)x|pm(dx)\int|H(x)x|^{p}\,m(dx) as T.T\to\infty. In particular, the expectation is bounded in TT on [0,).[0,\infty). Similarly, the expectation 𝔼ξ^[Xsq]\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[X_{s}^{q}] is also bounded in ss on [0,).[0,\infty). Since the right-hand side of Eq.(D.7) is bounded in TT on [0,),[0,\infty), we obtain the desired result. ∎

Proposition D.2.

Let gL2(m).g\in L^{2}(m). Then,

𝔼ξ^[g(XT)](0,)g(x)m(dx)\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[g(X_{T})]\to\int_{(0,\infty)}g(x)\,m(dx)

as T.T\to\infty.

Proof.

Consider the eigenvalue–eigenfunction problem ϕ=λϕ\mathcal{L}\phi=-\lambda\phi of the second-order differential operator

ϕ:=12σ2x2ϕ′′+(bαx2)ϕ\mathcal{L}\phi:=\frac{1}{2}\sigma^{2}x^{2}\phi^{\prime\prime}+(b-\alpha x^{2})\phi^{\prime}

densely defined on the space L2(m).L^{2}(m). We first show that the spectral gap is positive. By (Fulton et al., 2005, Theorem 12 (ii)), \mathcal{L} is bounded below since ϕ=1L2(m)\phi=1\in L^{2}(m) is non-oscillatory. By (Fulton et al., 2005, Theorem 14), it suffices to show that the essential spectrum is empty. We can write the equation ϕ=λϕ\mathcal{L}\phi=-\lambda\phi in the divergence form as

(p(x)ϕ(x))=λw(x)ϕ(x),\displaystyle-(p(x)\phi^{\prime}(x))^{\prime}=\lambda w(x)\phi(x), (D.8)

where

p(x)=e2bσ21x2ασ2x,w(x)=2σ2x2e2bσ21x2ασ2x.p(x)=e^{-\frac{2b}{\sigma^{2}}\frac{1}{x}-\frac{2\alpha}{\sigma^{2}}x}\,,\;w(x)=\frac{2}{\sigma^{2}x^{2}}e^{-\frac{2b}{\sigma^{2}}\frac{1}{x}-\frac{2\alpha}{\sigma^{2}}x}\,.

Using the Liouville transformation (for example, (Everitt, 2005, Section 7)), Eq.(D.8) becomes

Y(X)+Q(X)Y(X)=λY(X),-Y^{\prime}(X)+Q(X)Y(X)=\lambda Y(X)\,,

where

Q(X)=(p(x)/w3(x))1/4(p(x)((p(x)w(x))1/4))Q(X)=-(p(x)/w^{3}(x))^{1/4}(p(x)((p(x)w(x))^{-1/4})^{\prime})^{\prime}

and X=2σlnx.X=\frac{\sqrt{2}}{\sigma}\ln x. By direct calculation, Q(X)x=eσ2XQ(X)\simeq x=e^{\frac{\sigma}{\sqrt{2}}X} as X;X\to\infty; more precisely, limXQ(X)eσ2X\lim_{X\to\infty}\frac{Q(X)}{e^{\frac{\sigma}{\sqrt{2}}X}} exists and is a positive constant. In particular, Q(X)Q(X)\to\infty as XX\to\infty, and this implies that the essential spectrum is empty from (Ćurgus and Read, 2002, Corollary 4.2). Finally, since the spectral gap is positive, by (Qin and Linetsky, 2016, Theorem 5.2 (iii)), we have

f(T,ξ):=𝔼ξ^[g(XT)](0,)g(x)m(dx)f(T,\xi):=\mathbb{E}_{\xi}^{\hat{\mathbb{P}}}[g(X_{T})]\to\int_{(0,\infty)}g(x)\,m(dx)

for gL2(m).g\in L^{2}(m). This completes the proof. ∎

Proposition D.3.

Let (Xt)t0(X_{t})_{t\geq 0} be the solution of

dXt=(bαXt2)dt+σXtdB^t.dX_{t}=(b-\alpha X_{t}^{2})\,dt+\sigma X_{t}\,d\hat{B}_{t}\,.

Then, for T>0,T>0, the random variable XTX_{T} is in 𝔻1,2\mathbb{D}^{1,2}, and the Malliavin derivative is

DtXT=σXte2αtTXs𝑑s12σ2(Tt)+σ(B^TB^t)D_{t}X_{T}=\sigma X_{t}e^{-2\alpha\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}

for 0tT.0\leq t\leq T.

Proof.

Choose b~,a~>0\tilde{b},\tilde{a}>0 such that bax2<b~a~xb-ax^{2}<\tilde{b}-\tilde{a}x for all x>0.x>0. Let X~\tilde{X} be the solution of the SDE

dX~t=(b~a~X~t)dt+σX~dB^t.d\tilde{X}_{t}=(\tilde{b}-\tilde{a}\tilde{X}_{t})\,dt+\sigma\tilde{X}\,d\hat{B}_{t}\,.

This SDE has a unique strong solution since the coefficients are Lipschitz, and the solution X~\tilde{X} stays positive by (Zhao, 2009, Eq.(0.2)). For N,N\in\mathbb{N}, let ΛN\Lambda_{N} be a continuously differentiable function satisfying

ΛN(x)={bax2if xNb~a~xif xb~b+aN2a~+1\Lambda_{N}(x)=\left\{\begin{aligned} &b-ax^{2}&&\textnormal{if }x\leq N\\ &\tilde{b}-\tilde{a}x&&\textnormal{if }x\geq\frac{\tilde{b}-b+aN^{2}}{\tilde{a}}+1\end{aligned}\right. (D.9)

as well as ΛN(x)0\Lambda_{N}^{\prime}(x)\leq 0 for all x>0.x>0. Since ΛN\Lambda_{N} is a Lipschitz function, the SDE

dXt(N)=ΛN(Xt(N))dt+σXt(N)dB^tdX_{t}^{(N)}=\Lambda_{N}(X_{t}^{(N)})\,dt+\sigma X_{t}^{(N)}\,d\hat{B}_{t} (D.10)

has a unique strong solution (Xt(N))t0.(X_{t}^{(N)})_{t\geq 0}. By the comparison theorem (for example, see (Karatzas and Shreve, 1991, Proposition 2.18 in Chapter 5)), we know that XtXt(N)X~tX_{t}\leq X_{t}^{(N)}\leq\tilde{X}_{t} a.s.

Since the Malliavin derivative is a closed operator, it suffices to show that XT(N)XTX_{T}^{(N)}\to X_{T} in L2L^{2} as NN\to\infty and DtXT(N)σXte2αtTXs𝑑s12σ2(Tt)+σ(B^TB^t)D_{t}X_{T}^{(N)}\to\sigma X_{t}e^{-2\alpha\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})} in L2L^{2} as N.N\to\infty. We prove these in the following two stops. The first step is to show that the random variable XT(N)XTX_{T}^{(N)}\to X_{T} in L2L^{2} as N.N\to\infty. We use the dominated convergence theorem to prove this. For each N,N\in\mathbb{N}, define a stopping time τN:=inf{t0|XtN}.\tau_{N}:=\inf\{t\geq 0\,|\,X_{t}\geq N\}. Then, it is clear that (τN)N(\tau_{N})_{N\in\mathbb{N}} is nondecreasing and limτN=\lim\tau_{N}=\infty a.s. Let XτNX^{\tau_{N}} denote the stopped process of XX at τN\tau_{N}. Then, we have

XT(N)=XTτN,TτNX_{T}^{(N)}=X_{T}^{\tau_{N}},\;T\leq\tau_{N}

from the definition of ΛN.\Lambda_{N}. Letting N,N\to\infty, it follows that limNXT(N)=limNXTτN=XT\lim_{N\to\infty}X_{T}^{(N)}=\lim_{N\to\infty}X_{T}^{\tau_{N}}=X_{T} a.s. Observe that (XT(N)XT)2(|XT(N)|+|XT|)24X~T2(X_{T}^{(N)}-X_{T})^{2}\leq(|X_{T}^{(N)}|+|X_{T}|)^{2}\leq 4\tilde{X}_{T}^{2} and 𝔼[X~T2]<\mathbb{E}[\tilde{X}_{T}^{2}]<\infty by (Zhao, 2009, Corollay 2.2). The dominated convergence theorem implies that 𝔼[(XT(N)XT)2]0\mathbb{E}[(X_{T}^{(N)}-X_{T})^{2}]\to 0 as N.N\to\infty.

The second step is to show that

DtXT(N)=σXt(N)etTΛN(Xs(N))𝑑s12σ2(Tt)+σ(B^TB^t)σXte2atTXs𝑑s12σ2(Tt)+σ(B^TB^t) in L2D_{t}X_{T}^{(N)}=\sigma X_{t}^{(N)}e^{\int_{t}^{T}\Lambda_{N}^{\prime}(X_{s}^{(N)})\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\to\sigma X_{t}e^{-2a\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\;\textnormal{ in }L^{2}

as N.N\to\infty. Since the coefficients of SDE (D.10) are continuously differentiable with bounded derivatives, the solution Xt(N)𝔻1,2X_{t}^{(N)}\in\mathbb{D}^{1,2} for each tt and

DtXT(N)=σXt(N)etTΛN(Xs(N))𝑑s12σ2(Tt)+σ(B^TB^t)D_{t}X_{T}^{(N)}=\sigma X_{t}^{(N)}e^{\int_{t}^{T}\Lambda_{N}^{\prime}(X_{s}^{(N)})\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}

by (Fournié et al., 1999, Property P2). Observe that Xt(N)XtX_{t}^{(N)}\to X_{t} a.s. for each tt and ΛN(x)2ax\Lambda_{N}^{\prime}(x)\to-2ax for all x>0x>0 as N.N\to\infty. Thus,

σXt(N)etTΛN(Xs(N))𝑑s12σ2(Tt)+σ(B^TB^t)σXte2atTXs𝑑s12σ2(Tt)+σ(B^TB^t) a.s.\sigma X_{t}^{(N)}e^{\int_{t}^{T}\Lambda_{N}^{\prime}(X_{s}^{(N)})\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\to\sigma X_{t}e^{-2a\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\;\textnormal{ a.s.}

We claim that this convergence also holds in L2.L^{2}. By the dominated convergence theorem, it suffices to find a L2L^{2}-dominating function. Considering that ΛN(x)0\Lambda_{N}^{\prime}(x)\leq 0 for x>0,x>0, we have

σXt(N)etTΛN(Xs(N))𝑑s12σ2(Tt)+σ(B^TB^t)\displaystyle\sigma X_{t}^{(N)}e^{\int_{t}^{T}\Lambda_{N}^{\prime}(X_{s}^{(N)})\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})} σXt(N)e12σ2(Tt)+σ(B^TB^t)\displaystyle\leq\sigma X_{t}^{(N)}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})} (D.11)
σX~te12σ2(Tt)+σ(B^TB^t),\displaystyle\leq\sigma\tilde{X}_{t}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\,,

and similarly, σXte2atTXs𝑑s12σ2(Tt)+σ(B^TB^t)σX~te12σ2(Tt)+σ(B^TB^t).\sigma X_{t}e^{-2a\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\leq\sigma\tilde{X}_{t}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\,. Thus,

|σXT(N)etTΛN(Xs(N))𝑑s12σ2(Tt)+σ(B^TB^t)σXTe2atTXs𝑑s12σ2(Tt)+σ(B^TB^t)|\displaystyle\quad\big{|}\sigma X_{T}^{(N)}e^{\int_{t}^{T}\Lambda_{N}^{\prime}(X_{s}^{(N)})\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}-\sigma X_{T}e^{-2a\int_{t}^{T}X_{s}\,ds-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\big{|} (D.12)
2σX~te12σ2(Tt)+σ(B^TB^t).\displaystyle\leq 2\sigma\tilde{X}_{t}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})}\,.

The random variable 2σX~te12σ2(Tt)+σ(B^TB^t)2\sigma\tilde{X}_{t}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})} is an L2L^{2}-dominating function since

𝔼[(2σX~te12σ2(Tt)+σ(B^TB^t))2]=4σ2𝔼[X~t4]1/2𝔼[e2σ2(Tt)+4σ(B^TB^t)]1/2<\mathbb{E}[(2\sigma\tilde{X}_{t}e^{-\frac{1}{2}\sigma^{2}(T-t)+\sigma(\hat{B}_{T}-\hat{B}_{t})})^{2}]=4\sigma^{2}\mathbb{E}[\tilde{X}_{t}^{4}]^{1/2}\mathbb{E}[e^{-2\sigma^{2}(T-t)+4\sigma(\hat{B}_{T}-\hat{B}_{t})}]^{1/2}<\infty

by (Zhao, 2009, Proposition 2.1). This completes the proof. ∎

References

  • Alòs and Ewald (2008) Alòs, E. and Ewald, C.-O. (2008). Malliavin differentiability of the Heston volatility and applications to option pricing. Advances in Applied Probability, 40(1):144–162.
  • Alos et al. (2007) Alos, E., León, J. A., and Vives, J. (2007). On the short-time behavior of the implied volatility for jump-diffusion models with stochastic volatility. Finance and Stochastics, 11(4):571–589.
  • Alòs and Shiraya (2019) Alòs, E. and Shiraya, K. (2019). Estimating the Hurst parameter from short term volatility swaps: a Malliavin calculus approach. Finance and Stochastics, 23(2):423–447.
  • Benhamou (2003) Benhamou, E. (2003). Optimal Malliavin weighting function for the computation of the Greeks. Mathematical Finance, 13(1):37–53.
  • Bi and Cai (2019) Bi, J. and Cai, J. (2019). Optimal investment-reinsurance strategies with state dependent risk aversion and VaR constraints in correlated markets. Insurance: Mathematics and Economics, 85:1–14.
  • Borovička et al. (2014) Borovička, J., Hansen, L. P., and Scheinkman, J. A. (2014). Shock elasticities and impulse responses. Mathematics and Financial Economics, 8(4):333–354.
  • Carassus and Rásonyi (2007) Carassus, L. and Rásonyi, M. (2007). Optimal strategies and utility-based prices converge when agents’ preferences do. Mathematics of Operations Research, 32(1):102–117.
  • Carr and Willems (2019) Carr, P. and Willems, S. (2019). A lognormal type stochastic volatility model with quadratic drift. Available at SSRN 3421304.
  • Ćurgus and Read (2002) Ćurgus, B. and Read, T. T. (2002). Discreteness of the spectrum of second-order differential operators and associated embedding theorems. Journal of Differential Equations, 184(2):526–548.
  • Delong (2019) Delong, Ł. (2019). Optimal investment for insurance company with exponential utility and wealth-dependent risk aversion coefficient. Mathematical Methods of Operations Research, 89(1):73–113.
  • Everitt (2005) Everitt, W. N. (2005). A catalogue of Sturm–Liouville differential equations. In Sturm-Liouville Theory, pages 271–331. Springer.
  • Fleming and Sheu (2000) Fleming, W. H. and Sheu, S. (2000). Risk-sensitive control and an optimal investment model. Mathematical Finance, 10(2):197–213.
  • Fournié et al. (1999) Fournié, E., Lasry, J.-M., Lebuchoux, J., Lions, P.-L., and Touzi, N. (1999). Applications of Malliavin calculus to Monte Carlo methods in finance. Finance and Stochastics, 3(4):391–412.
  • Fulton et al. (2005) Fulton, C., Pruess, S., and Xie, Y. (2005). The automatic classification of Sturm-Liouville problems. J. Appl. Math. Comput, 124:149–186.
  • Guasoni and Robertson (2015) Guasoni, P. and Robertson, S. (2015). Static fund separation of long-term investments. Mathematical Finance, 25(4):789–826.
  • Hansen (2012) Hansen, L. P. (2012). Dynamic valuation decomposition within stochastic economies. Econometrica, 80(3):911–967.
  • Hansen and Scheinkman (2009) Hansen, L. P. and Scheinkman, J. (2009). Long-term risk: An operator approach. Econometrica, 77(1):177–234.
  • Hansen and Scheinkman (2012) Hansen, L. P. and Scheinkman, J. (2012). Pricing growth-rate risk. Finance and Stochastics, 16(1):1–15.
  • Jeanblanc et al. (2009) Jeanblanc, M., Yor, M., and Chesney, M. (2009). Mathematical methods for financial markets. Springer Science & Business Media.
  • Jouini and Napp (2004) Jouini, E. and Napp, C. (2004). Convergence of utility functions and convergence of optimal strategies. Finance and Stochastics, 8(1):133–144.
  • Kallenberg (2006) Kallenberg, O. (2006). Foundations of modern probability. Springer Science & Business Media.
  • Karatzas and Shreve (1991) Karatzas, I. and Shreve, S. (1991). Brownian Motion and Stochastic Calculus, volume 113. Springer Science & Business Media.
  • León et al. (2003) León, J. A., Navarro, R., and Nualart, D. (2003). An anticipating calculus approach to the utility maximization of an insider. Mathematical Finance: An International Journal of Mathematics, Statistics and Financial Economics, 13(1):171–185.
  • Löcherbach (2015) Löcherbach, E. (2015). Ergodicity and speed of convergence to equilibrium for diffusion processes.
  • Malliavin and Paul (2006) Malliavin, T. and Paul, A. (2006). Stochastic calculus of variations in mathematical finance. Springer.
  • Mocha and Westray (2013) Mocha, M. and Westray, N. (2013). The stability of the constrained utility maximization problem: a BSDE approach. SIAM Journal on Financial Mathematics, 4(1):117–150.
  • Nualart and Nualart (2018) Nualart, D. and Nualart, E. (2018). Introduction to Malliavin calculus, volume 9. Cambridge University Press.
  • Nutz (2012) Nutz, M. (2012). Risk aversion asymptotics for power utility maximization. Probability Theory and Related Fields, 152(3-4):703–749.
  • Paravisini et al. (2017) Paravisini, D., Rappoport, V., and Ravina, E. (2017). Risk aversion and wealth: Evidence from person-to-person lending portfolios. Management Science, 63(2):279–297.
  • Park (2018) Park, H. (2018). Sensitivity analysis of long-term cash flows. Finance and Stochastics, 22(4):773–825.
  • Park (2019) Park, H. (2019). Convergence rates of large-time sensitivities with the Hansen–Scheinkman decomposition. arXiv preprint arXiv:1912.03404.
  • Park and Sturm (2019) Park, H. and Sturm, S. (2019). A sensitivity analysis of the long-term expected utility of optimal portfolios. Available at SSRN 3401532.
  • Protter (2004) Protter, P. E. (2004). Stochastic integration and stochastic differential equations. Springer, Berlin.
  • Qin and Linetsky (2016) Qin, L. and Linetsky, V. (2016). Positive eigenfunctions of Markovian pricing operators: Hansen-Scheinkman factorization, Ross recovery, and long-term pricing. Operations Research, 64(1):99–117.
  • Robertson and Xing (2015) Robertson, S. and Xing, H. (2015). Large time behavior of solutions to semilinear equations with quadratic growth in the gradient. SIAM Journal on Control and Optimization, 53(1):185–212.
  • Zariphopoulou (2001) Zariphopoulou, T. (2001). A solution approach to valuation with unhedgeable risks. Finance and stochastics, 5(1):61–82.
  • Zariphopoulou and Zhou (2009) Zariphopoulou, T. and Zhou, T. (2009). Investment performance measurement under asymptotically linear local risk tolerance. In Handbook of numerical analysis, volume 15, pages 227–253. Elsevier.
  • Zhao (2009) Zhao, B. (2009). Inhomogeneous geometric Brownian motion. Available at SSRN 1429449.