Skip to main content
Log in

Optimal consumption and investment with Epstein–Zin recursive utility

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

We study continuous-time optimal consumption and investment with Epstein–Zin recursive preferences in incomplete markets. We develop a novel approach that rigorously constructs the solution of the associated Hamilton–Jacobi–Bellman equation by a fixed point argument and makes it possible to compute both the indirect utility and, more importantly, optimal strategies. Based on these results, we also establish a fast and accurate method for numerical computations. Our setting is not restricted to affine asset price dynamics; we only require boundedness of the underlying model coefficients.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Our analysis imposes no structural conditions on the underlying model coefficients, but requires them to be bounded; see (A1) and (A2) in Sect. 4 and (A1′) in Sect. 7.

  2. In particular, [42] covers specifications with (untruncated) affine dynamics as in Kim and Omberg [24] and Heston [23].

  3. Condition (3.1) holds if and only if one of the conditions (a), (b), (c), and (d) in [26, Proposition 3.2] is satisfied; see also [40, (2)]. We are not aware of rigorous results that ensure (E1) and (E2) for parametrizations not subsumed by (3.1).

  4. Typically, the pair \((X,Z)\) would be referred to as a solution of the BSDE (6.4). For simplicity of notation, and since \(Z\) is not required for our further analysis, here and in the following, we also refer to \(X\) alone as a solution of (6.4).

  5. Machine: Intel® Core™ i3-540 Processor (4M Cache, 3.06 GHz), 4 GB RAM.

  6. Here we slightly abuse notation since \(\langle u\rangle^{q}_{x}\) has only been defined for functions on \([0,T]\times \mathbb{R}^{d}\). Of course, for \(u:\ \mathbb{R}^{d}\to \mathbb{R}\) and \(q\in(0,1)\), we understand that \(\langle u\rangle^{q}_{x} :=\sup_{x,x' \in \mathbb{R}^{d},\ |x-x'| \leq 1} \frac {|u(x) - u(x')|}{|x-x'|^{q}}\).

References

  1. Antonelli, F.: Stability of backward stochastic differential equations. Stoch. Process. Appl. 62, 103–114 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  2. Barberis, N.C.: Investing for the long run when returns are predictable. J. Finance 55, 225–264 (2000)

    Article  Google Scholar 

  3. Berdjane, B., Pergamenshchikov, S.: Optimal consumption and investment for markets with random coefficients. Finance Stoch. 17, 419–446 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bertoldi, M., Lorenzi, L.: Estimates of the derivatives for parabolic operators with unbounded coefficients. Trans. Am. Math. Soc. 357, 2627–2664 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  5. Briand, P., Carmona, R.: BSDEs with polynomial growth generators. J. Appl. Math. Stoch. Anal. 13, 207–238 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  6. Briand, P., Hu, Y.: Quadratic BSDEs with convex generators and unbounded terminal conditions. Probab. Theory Relat. Fields 141, 543–567 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  7. Campbell, J.Y., Chacko, G., Rodriguez, J., Viceira, L.M.: Strategic asset allocation in a continuous-time VAR model. J. Econ. Dyn. Control 128, 2195–2214 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  8. Campbell, J.Y., Viceira, L.M.: Consumption and portfolio decisions when expected returns are time varying. Q. J. Econ. 114, 433–495 (1999)

    Article  MATH  Google Scholar 

  9. Campbell, J.Y., Viceira, L.M.: Strategic Asset Allocation. Oxford University Press, London (2002)

    Book  MATH  Google Scholar 

  10. Cerrai, S.: Elliptic and parabolic equations in \(\mathbb{R}^{n}\) with coefficients having polynomial growth. Commun. Partial Differ. Equ. 21, 281–317 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  11. Cerrai, S.: Second order PDEs. In: Finite and Infinite Dimension. Springer, Berlin (2001)

    Google Scholar 

  12. Chacko, G., Viceira, L.M.: Dynamic consumption and portfolio choice with stochastic volatility in incomplete markets. Rev. Financ. Stud. 18, 1369–1402 (2005)

    Article  Google Scholar 

  13. Christoffersen, P., Jacobs, K., Mimouni, K.: Volatility dynamics for the S&P500: evidence from realized volatility, daily returns, and option prices. Rev. Financ. Stud. 23, 3141–3189 (2010)

    Article  Google Scholar 

  14. Cochrane, J.Y.: A mean-variance benchmark for intertemporal portfolio theory. J. Finance 69, 1–49 (2014)

    Article  Google Scholar 

  15. Delarue, F.: On the existence and uniqueness of solutions to FBSDEs in a non-degenerate case. Stoch. Process. Appl. 99, 209–286 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  16. Delbaen, F., Hu, Y., Richou, A.: On the uniqueness of solutions to quadratic BSDEs with convex generators and unbounded terminal conditions. Ann. Inst. Henri Poincaré B, Probab. Stat. 47, 559–574 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  17. Duffie, D., Epstein, L.G.: Stochastic differential utility. Econometrica 60, 353–394 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  18. Duffie, D., Lions, P.L.: PDE solutions of stochastic differential utility. J. Math. Econ. 21, 577–606 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  19. Duffie, D., Skiadas, C.: Continuous-time security pricing. J. Math. Econ. 23, 107–131 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  20. El Karoui, N., Peng, S., Quenez, M.C.: Backward stochastic differential equations in finance. Math. Finance 7, 1–71 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  21. Elworthy, K., Li, X.M.: Formulae for the derivatives of heat semigroups. J. Funct. Anal. 125, 252–286 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  22. Epstein, L.G., Zin, S.E.: Substitution, risk aversion, and the temporal behavior of consumption and asset returns: a theoretical framework. Econometrica 57, 937–969 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  23. Heston, S.L.: A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev. Financ. Stud. 6, 327–343 (1993)

    Article  Google Scholar 

  24. Kim, T.S., Omberg, E.: Dynamic nonmyopic portfolio behavior. Rev. Financ. Stud. 9, 141–161 (1996)

    Article  Google Scholar 

  25. Kobylanski, M.: Backward stochastic differential equations and partial differential equations with quadratic growth. Ann. Appl. Probab. 28, 558–602 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kraft, H., Seifried, F.T., Steffensen, M.: Consumption-portfolio optimization with recursive utility in incomplete markets. Finance Stoch. 17, 161–196 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  27. Kreps, D.M., Porteus, E.L.: Temporal resolution of uncertainty and dynamic choice theory. Econometrica 46, 185–200 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  28. Kreps, D.M., Porteus, E.L.: Temporal von Neumann–Morgenstern and induced preferences. J. Econ. Theory 20, 81–109 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  29. Ladyzenskaja, O.A., Solonnikov, V., Ural’ceva, N.: Linear and Quasi-Linear Equations of Parabolic Type. Am. Math. Soc., Providence (1968)

    Google Scholar 

  30. Liu, J.: Portfolio selection in stochastic environments. Rev. Financ. Stud. 20, 1–39 (2007)

    Article  Google Scholar 

  31. Liu, J., Pan, J.: Dynamic derivative strategies. J. Financ. Econ. 69, 401–430 (2003)

    Article  Google Scholar 

  32. Ma, J., Protter, P., Yong, J.: Solving forward-backward stochastic differential equations explicitly—a four step scheme. Probab. Theory Relat. Fields 98, 339–359 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  33. Ma, J., Yin, H., Zhang, J.: On non-Markovian forward-backward SDEs and backward stochastic PDEs. Stoch. Process. Appl. 122, 3980–4004 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  34. Marinacci, M., Montrucchio, L.: Unique solutions for stochastic recursive utilities. J. Econ. Theory 145, 1776–1804 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  35. Pardoux, E., Peng, S.: Backward stochastic differential equations and quasilinear parabolic differential equations. In: Rozovskii, B.L., Sowers, R.B. (eds.) Stochastic Partial Differential Equations and Their Applications. Lecture Notes in Control and Information Sciences, vol. 176, pp. 200–217 (1992)

    Chapter  Google Scholar 

  36. Schroder, M., Skiadas, C.: Optimal consumption and portfolio selection with stochastic differential utility. J. Econ. Theory 89, 68–126 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  37. Schroder, M., Skiadas, C.: Optimal lifetime consumption-portfolio strategies under trading constraints and generalized recursive preferences. Stoch. Process. Appl. 108, 155–202 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  38. Schroder, M., Skiadas, C.: Lifetime consumption-portfolio choice under trading constraints, recursive preferences, and nontradeable income. Stoch. Process. Appl. 115, 1–30 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  39. Schroder, M., Skiadas, C.: Optimality and state pricing in constrained financial markets with recursive utility under continuous and discontinuous information. Math. Finance 2, 199–238 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  40. Seiferling, T., Seifried, F.T.: Epstein–Zin stochastic differential utility: Existence, uniqueness, concavity, and utility gradients. Working paper (2016). Available online at http://ssrn.com/abstract=2625800

  41. Wachter, J.A.: Portfolio and consumption decisions under mean-reverting returns: an exact solution for complete markets. J. Financ. Quant. Anal. 37, 63–91 (2002)

    Article  Google Scholar 

  42. Xing, H.: Consumption–investment optimization with Epstein–Zin utility in incomplete markets. Finance Stoch. 21 (this issue, 2017). doi:10.1007/s00780-016-0297-z

  43. Zariphopoulou, T.: A solution approach to valuation with unhedgeable risks. Finance Stoch. 5, 61–82 (2001)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

All authors wish to thank Jakša Cvitanić (editor), the Associate Editor and two referees (anonymous) for very helpful comments. We thank Darrell Duffie, Bernard Dumas, Francis Longstaff, Claus Munk, Lukas Schmid, and Carsten Sørensen for very helpful discussions, comments and suggestions. We also thank the participants of the Bachelier Finance Society 8th World Congress, the 11th German Probability and Stochastics Days, the 9th Bachelier Colloquium and seminar participants at ETH Zürich, Copenhagen Business School, the University of Copenhagen, and the University of Southern Denmark for many helpful comments and suggestions. Holger Kraft gratefully acknowledges financial support from Deutsche Forschungsgemeinschaft (DFG) and the Center of Excellence SAFE, funded by the State of Hessen initiative for research LOEWE. Thomas Seiferling gratefully acknowledges financial support from Studienstiftung des Deutschen Volkes.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Frank Thomas Seifried.

Appendices

Appendix A: Proofs omitted from the main text

Proof of Lemma 4.6

Since \(h\) solves the reduced HJB equation (4.7), we have

$$\begin{aligned} H(z, \hat{\pi}, \hat{c}) :=& w_{t} + x(r + \hat{\pi}\lambda) w_{x} - \hat{c}w_{x} + \frac{1}{2} x^{2} \hat{\pi}^{2} \sigma ^{2} w_{xx} + \alpha w_{y} \\ &{} + \frac{1}{2} \beta^{2} w_{yy} + x \hat{\pi}\sigma \beta \rho w_{xy} + f(\hat{c},w) = 0, \end{aligned}$$

where \(z:=(t,x,y, w_{x}, w_{y}, w_{x_{y}} w_{xx}, w_{yy})\). Separating the terms in the function \(H\) as \(H(z, \pi ,c) :=u(z, \pi) + s(z, c) + q(z) \), it is easy to see that the candidate solutions \(\hat{\pi}\) and \(\hat{c}\) defined in (4.6) are the unique solutions of the associated first-order conditions

$$\begin{aligned} \begin{aligned} 0&= s_{c}(z, c) = -w_{x} + f_{c}(c, w),\\ 0&= u_{\pi}(z, \pi) =x \lambda w_{x} + \pi x^{2} \sigma ^{2} w_{xx} + x \sigma \beta \rho w_{xy}. \end{aligned} \end{aligned}$$
(A.1)

Concavity of \(u\) and \(s\) implies that \(H(z, \hat{\pi}, \hat{c}) = { \sup_{\pi\in \mathbb{R},\, c\in(0,\infty)}} H(z, \pi ,c)\). □

Proof of Lemma 4.7

By (A1) and (A2), \(\tilde{\alpha}\) and \(\tilde{r}\) are bounded. Moreover,

$$\begin{aligned} |\tilde{\alpha}(y) - \tilde{\alpha}(\bar{y})| \leq& \bigg|\frac {1-\gamma}{\gamma}\bigg| \rho \bigg(\bigg|\frac{\lambda(y)}{\sigma(y)} \bigg| |\beta(y) -\beta(\bar{y})| + \bigg|\frac {\beta(\bar{y})}{\sigma(y)}\bigg| | {\lambda (y)} - {\lambda(\bar{y})} |\bigg)\\ &{}+ |\beta(\bar{y}) \lambda(\bar{y})| \bigg| \frac {\sigma (\bar{y}) -\sigma(y)}{\sigma(y)\sigma(\bar{y})}\bigg| + |\alpha (y) - \alpha(\bar{y})| \end{aligned}$$

so that \(\tilde{\alpha}\) is Lipschitz-continuous. Finally,

$$\begin{aligned} k|\tilde{r}(y) -\tilde{r}(\bar{y})| \leq& |1-\gamma| |r(y)-r(\bar{y})| \\ &{}+ \bigg|\frac {1-\gamma}{\gamma}\bigg| \|\lambda\|_{\infty}\Big(\inf_{x \in \mathbb{R}} \sigma(x)\Big)^{-2} |\lambda(y) - \lambda(\bar{y})|\\ &{}+ \bigg|\frac {1-\gamma}{\gamma}\bigg| \|\lambda\|_{\infty}^{2} \|\sigma\|_{\infty}\Big(\inf_{x \in \mathbb{R}} \sigma(x)\Big)^{-4} |\sigma(\bar{y}) -\sigma (y)|. \end{aligned}$$

 □

Proof of Lemma 5.3

The candidate optimal wealth process \(\hat{X}\) has dynamics

$$\mathrm{d} \hat{X}_{t} = \hat{X}_{t}\bigg( \Big(r_{t} + \frac{1}{\gamma}\frac{\lambda_{t}^{2}}{\sigma_{t}^{2}} + \frac{k}{\gamma}\frac {\lambda_{t} \beta_{t} \rho}{\sigma_{t}} \frac{h_{y}}{h} - \delta^{\psi}h^{q-1}\Big)\,\mathrm{d} t + \Big(\frac{1}{\gamma}\frac{\lambda_{t}}{\sigma_{t}} + \frac{k}{\gamma}\beta_{t} \rho \frac{h_{y}}{h}\Big) \,\mathrm{d} W_{t}\bigg). $$

Put \(a_{t}:=r_{t} + \frac{1}{\gamma}\frac{\lambda_{t}^{2}}{\sigma_{t}^{2}}+ \frac{k}{\gamma}\frac{\lambda_{t}\beta_{t} \rho}{\sigma_{t}} \frac{h_{y}}{h} - \delta^{\psi}h^{q-1}\) and \(b_{t}:=\frac{1}{\gamma}\frac{\lambda_{t}}{\sigma_{t}} + \frac{k}{\gamma}\beta_{t} \rho \frac {h_{y}}{h}\). Our assumptions on the coefficients and on \(h_{y}\) and \(h\) imply that both \(a\) and \(b\) are bounded. By Itô’s formula,

$$\hat{X}_{t}^{p} = x^{p} \, \exp\bigg(p{ \int_{0}^{t}} \Big(a_{s} + \frac{1}{2} (p-1) b_{s}^{2} \Big) \,\mathrm{d} s \bigg) {\mathcal {E}}_{t} \bigg(p{ \int_{0}^{{\,\cdot \,}}} b_{s} \,\mathrm{d} W_{s} \bigg), $$

where \(\mathcal {E}_{t}({\,\cdot \,})\) denotes the stochastic exponential. Choose the constant \(M>0\) such that \(|p a_{t}| + |p(p-1) b_{t}^{2}|, |p b_{t}| < M\) for all \(t\in[0,T]\). By Novikov’s condition, \({\mathcal {E}}_{t} (p\int_{0}^{{\,\cdot \,}}b_{s} \,\mathrm{d} W_{s} )\) is then an \(L^{2}\)-martingale; so using Doob’s \(L^{2}\)-inequality, we get

$$\operatorname{E}\bigg[{ \sup _{t \in [0,T]}} \hat{X}_{t}^{p}\bigg] \leq 2 x^{p} e^{M T} \operatorname{E}\bigg[{\mathcal {E}}_{T} \left(p{ \int_{0}^{{\,\cdot \,}}} b_{s} \,\mathrm{d} W_{s} \right)^{2} \bigg]^{\frac{1}{2}} < \infty. $$

 □

Proof of Lemma 5.4

Lemma 5.3 and the boundedness of \(\delta^{\psi}h(t,Y_{t})^{q-1}\) imply that \(\operatorname{E}[ {\sup_{t\in[0,T]}} |\hat{c}_{t}|^{p} ] < \infty\) for all \(p \in \mathbb{R}\). In particular, \(\hat{c} \) is in \(\mathcal{C}\). By Itô’s formula,

$$\begin{aligned} \mathrm{d} V_{t} = \bigg(&w_{t}+ \hat{X}_{t}(r_{t} + \hat{\pi}_{t} \lambda_{t}) w_{x} - \hat{c}_{t}w_{x} + \frac{1}{2} \hat{X}_{t}^{2} \hat{\pi}_{t}^{2} \sigma_{t} ^{2} w_{xx} + \alpha_{t} w_{y} + \frac{1}{2} \beta_{t}^{2} w_{yy} \\ &{}+ \hat{X}_{t} \hat{\pi}_{t} \sigma_{t} \beta_{t} \rho w_{xy} \bigg) \,\mathrm{d} t + \,\mathrm{d} M_{t}, \end{aligned}$$

where \(M\) is a local martingale. Hence \(\mathrm{d} V_{t} = - f(\hat{c}_{t}, V_{t}) \,\mathrm{d} t + \,\mathrm{d} M_{t}\) by Lemma 4.6. Moreover, exploiting the special form of \(w\), we get

$$\mathrm{d} M_{t}= V_{t}\bigg( \frac{1-\gamma}{\gamma}\frac{\lambda_{t}}{\sigma_{t}} + \frac{\rho k}{\gamma}\beta_{t} \frac{h_{y}}{h} \bigg) \,\mathrm{d} W_{t} + V_{t} k \sqrt{1-\rho^{2}} \beta_{t}\frac{h_{y}}{h} \,\mathrm{d} \bar{W}_{t}. $$

Here \(V_{t}\) can be rewritten as \(V_{t} = w(t, \hat{X}_{t}, Y_{t}) = \frac{1}{1-\gamma} \hat{X}_{t}^{1-\gamma} h(t,Y_{t})^{k}\). By (4.8), the function \(h\) is bounded and bounded away from zero. Thus we have for all \(p \in \mathbb{R}\) by Lemma 5.3 that \(\operatorname{E}[{\sup_{t\in[0,T]}} |V_{t}|^{p} ] <\infty\). Hence \(V\) is a utility process associated with \(\hat{c}\); by (E1), it follows that \(V=V^{\hat{c}}\). Finally, the first-order condition (A.1) for the optimal consumption implies that \(w_{x}(t, \hat{X}_{t}, Y_{t}) = f_{c}(t, w(t, \hat{X}_{t}, Y_{t})) = f_{c}(\hat{c}_{t}, \hat{V}_{t})\). □

Proof of Lemma 5.5

For simplicity of notation, we set \(r_{t}:=r(Y_{t})\), \(\lambda_{t}:=\lambda(Y_{t})\) and \(\sigma_{t}:=\sigma(Y_{t})\). We have \(\mathrm{d} Z^{\pi, c}_{t} = \hat{m}_{t} c_{t} \,\mathrm{d} t + \hat{m}_{t} \,\mathrm{d} X_{t}^{\pi,c} + X_{t}^{\pi ,c} \,\mathrm{d} \hat{m}_{t} + \,\mathrm{d} [\hat{m},X^{\pi,c}]_{t}\) by the product rule. Inserting the dynamics of \(X^{\pi ,c}\) from (4.1), we get

$$ \mathrm{d} Z^{\pi, c}_{t} = \hat{m}_{t} X_{t}^{\pi ,c}\big((r_{t} + \pi_{t} \lambda_{t}) \,\mathrm{d} t + \pi_{t} \sigma _{t} \,\mathrm{d} W_{t}\big) + X_{t}^{\pi ,c} \,\mathrm{d} \hat{m}_{t} + \,\mathrm{d} [ \hat{m}, X^{\pi ,c }]_{t}. $$

By Lemma 5.4, \(\hat{V}_{t}=w(t,\hat{X}_{t},Y_{t})\) and \(\hat{m}_{t} = e^{\int_{0}^{t} f_{v}(\hat{c}_{s}, \hat{V}_{s}) \,\mathrm{d} s} w_{x}(t,\hat{X}_{t},Y_{t})\). From here on, we abbreviate \(f_{v} = f_{v}(\hat{c}_{t}, \hat{V}_{t})\), \(w_{x} = w_{x}(t, \hat{X}_{t}, Y_{t})\) etc. Clearly, we have \(\mathrm{d} \hat{m}_{t} = \hat{m}_{t} ( f_{v} \,\mathrm{d} t + \frac{\mathrm{d} w_{x}}{w_{x}} )\).

From the explicit expression \(f_{v}(c, v) = \delta \frac {\phi -\gamma}{1-\phi} c^{1 -\phi} ((1-\gamma)v)^{\frac{\phi-1}{1-\gamma}} - \delta \theta\), we obtain \(f_{v}(\hat{c}_{t}, w(t, \hat{X}_{t}, Y_{t}))= \frac{\phi-\gamma}{1-\phi} \delta^{\psi} h^{q -1} - \delta\theta\). By Itô’s formula,

$$\mathrm{d} w_{x} = w_{x} \bigg(\frac {w_{xt} }{w_{x}} \,\mathrm{d} t + \frac {w_{xx} }{w_{x}} \,\mathrm{d} \hat{X}_{t} + \frac{1}{2} \frac {w_{xxx} }{w_{x}} \,\mathrm{d} [\hat{X}]_{t} + \frac{1}{2} \frac {w_{xyy} }{w_{x}} \,\mathrm{d} [Y]_{t} + \frac {w_{xxy} }{w_{x}} \,\mathrm{d} [\hat{X}, Y]_{t} \bigg). $$

Substituting for \(w\), we find

$$\begin{aligned} \frac {\mathrm{d} w_{x}}{k w_{x}} =& \frac {h_{t}}{h} \,\mathrm{d} t -\frac{\gamma}{k} \frac {\,\mathrm{d} \hat{X}_{t} }{\hat{X}_{t}} + \frac {h_{y}}{h} \,\mathrm{d} Y_{t} + \frac{1}{2} \frac {\gamma (1+ \gamma)}{k} \frac {\,\mathrm{d} [\hat{X}]_{t}}{\hat{X}_{t}^{2}}\\ &{} +\frac{1}{2} \bigg((k-1) \frac {h_{y}^{2}}{h^{2}} + \frac {h_{yy}}{h} \bigg)\,\mathrm{d} [Y]_{t} - \frac{\gamma}{\hat{X}_{t}} \frac {h_{y}}{h} \,\mathrm{d} [\hat{X}, Y]_{t}. \end{aligned}$$

Plugging in the candidate \(\hat{\pi}\) from (5.1) and the dynamics of \(\hat{X}\) and \(Y\) yields

$$\frac{\mathrm{d} w_{x}}{k w_{x}} = A^{1}_{t} \,\mathrm{d} t + A^{2}_{t} \,\mathrm{d} t - \frac{1}{k} \frac {\lambda_{t}}{\sigma _{t}} \,\mathrm{d} W_{t} + \sqrt{1- \rho^{2}}\beta_{t} \frac {h_{y}}{h} \,\mathrm{d} \bar{W}_{t}, $$

where

$$\begin{aligned} A_{t}^{1} &:=\frac {h_{t}}{h} -\frac{\gamma}{k} r_{t} + \frac{1}{2} \frac{1}{k} \frac {1- \gamma}{\gamma}\frac {\lambda_{t}^{2}}{\sigma_{t}^{2}} + \frac{1}{\gamma}\frac {\lambda_{t} \beta_{t} \rho}{\sigma}\frac {h_{y}}{h} + \frac{\gamma}{k} \delta^{\psi}h^{q-1} + \frac{k}{2} \frac {1+ \gamma}{\gamma}\beta_{t}^{2} \rho^{2} \frac{h_{y}^{2}}{h^{2}},\\ A_{t}^{2} &:=\frac {h_{y}}{h} \bigg( \alpha_{t} - \frac{\rho \beta_{t} \lambda_{t}}{\sigma_{t}}\bigg) + \frac {h_{y}^{2} }{h^{2}} \left( \frac {k-1}{2} \beta_{t}^{2} - k \beta_{t}^{2} \rho^{2} \right) + \frac {\beta_{t}^{2}}{2} \frac {h_{yy}}{h}. \end{aligned}$$

For the sum of the \(\frac {h_{y}^{2}}{h^{2}}\)-terms, we have

$$\frac{k}{2} \frac {1+ \gamma}{\gamma}\beta_{t}^{2} \rho^{2} \frac{h_{y}^{2}}{h^{2}} + \frac {h_{y}^{2} }{h^{2}} \bigg( \frac {k-1}{2} \beta_{t}^{2} - k \beta_{t}^{2} \rho^{2}\! \bigg) = \beta_{t}^{2} \frac{h_{y}^{2}}{h^{2}} \bigg( \frac{k}{2} \rho ^{2} \frac {1+ \gamma}{\gamma}+ \frac {k-1}{2} - \rho^{2} k \bigg) = 0 $$

by our choice of \(k\). Combining the above, we obtain

$$\begin{aligned} \mathrm{d} \hat{m}_{t} =& k \hat{m}_{t} \bigg(\frac {h_{t}}{h} + \frac{1}{k} \Big( -\gamma r_{t} + \frac{1}{2} \frac {1- \gamma}{\gamma}\frac {\lambda_{t} ^{2}}{\sigma_{t}^{2}} - \delta \theta \Big) \\ &\quad\ \quad{}+ \tilde{\alpha}_{t} \frac {h_{y}}{h} + \frac {\beta_{t}^{2}}{2} \frac {h_{yy}}{h} + \frac {\phi \theta}{k} \delta ^{\psi}h^{q-1} \bigg)\\ &{}+ k \hat{m}_{t} \bigg(- \frac{1}{k} \frac {\lambda_{t}}{\sigma _{t}} \,\mathrm{d} W_{t} + \sqrt{1- \rho^{2}}\beta_{t} \frac {h_{y}}{h} \,\mathrm{d} \bar{W}_{t} \bigg), \end{aligned}$$

and it follows that \(\mathrm{d} [\hat{m}, X^{\pi , c}]_{t} = - \lambda_{t} \pi_{t} \hat{m}_{t} X_{t}^{\pi,c} \,\mathrm{d} t\). Since \(h\) solves (4.7), we get

$$\begin{aligned} \mathrm{d} Z^{\pi, c}_{t} &= \hat{m}_{t} X_{t}^{\pi ,c}\big((r_{t} + \pi_{t} \lambda_{t}) \,\mathrm{d} t + \pi_{t} \sigma _{t} \,\mathrm{d} W_{t}\big) + X_{t}^{\pi ,c} \,\mathrm{d} \hat{m}_{t} + \,\mathrm{d} [ \hat{m}, X^{\pi ,c }]_{t}\\ &= \hat{m}_{t} X_{t}^{\pi ,c} \frac{1}{h} \bigg({h_{t}} - \tilde{r}_{t} h + \tilde{\alpha}_{t} {h_{y}} + \frac{1}{2}\beta_{t}^{2} h_{yy} + \frac{\delta^{\psi}}{1-q} h^{q} \bigg) \,\mathrm{d} t + \,\mathrm{d} M_{t} = \,\mathrm{d} M_{t}, \end{aligned}$$

where \(\mathrm{d} M_{t}:=\hat{m}_{t} X_{t}^{\pi,c} ((\pi_{t} \sigma_{t} - \frac {\lambda_{t}}{\sigma _{t}} ) \,\mathrm{d} W_{t} + k\sqrt{1- \rho^{2}}\beta_{t} \frac {h_{y}}{h} \,\mathrm{d} \bar{W}_{t} )\) defines a local martingale \(M\). A direct calculation using the definition of \(\hat{\pi}\) yields the statement for \(Z^{\hat{\pi}, \hat{c}}\). □

Proof of Lemma 5.6

Recall that \(\underline{h}\leq h\leq \overline{h}\) so that

$$f_{v}(\hat{c}_{s},\hat{V}_{s}) = \frac{\phi-\gamma}{1-\phi} \delta^{\psi} h(s,Y_{s})^{q-1} - \delta\theta \leq \bigg|\frac{\phi-\gamma}{1-\phi}\bigg| \delta^{\psi} \left( \underline{h}^{q-1} + \overline{h}^{q-1}\right) + |\delta\theta| =: m_{1} $$

and we get \(0 \leq \exp(p{\int_{0}^{T}} f_{v}(\hat{c}_{s}, \hat{V}_{s}) \,\mathrm{d} s ) \leq e^{Tp m_{1}}\). On the other hand, Lemma 5.4 implies that \(\operatorname{E}[{\sup_{t \in [0,T]}} f_{c}(\hat{c}_{t}, \hat{V}_{t})^{p}] < \infty\) for all \(p \in \mathbb{R}\). It follows that

$$\operatorname{E}\bigg[ { \sup_{t\in[0,T]}} \hat{m}_{t}^{p} \bigg] < \infty\quad\text{for all }p>1. $$

To show that \(Z^{\hat{\pi}, \hat{c}}\) is a martingale, note that \(\frac {1- \gamma}{\gamma}\frac {\lambda_{t}}{\sigma_{t}} + \frac{k}{\gamma}\beta_{t} \rho \frac {h_{y}}{h}\) is uniformly bounded by some \(c>0\). Hence by Lemma 5.3, we have

$${\int_{0}^{T}} \operatorname{E}\bigg[\hat{m}_{t}^{2} \hat{X}_{t}^{2} \bigg(\frac{1- \gamma}{\gamma}\frac {\lambda_{t}}{\sigma_{t}} + \frac{k}{\gamma}\beta_{t} \rho \frac {h_{y}}{h}\bigg)^{2} \bigg] \,\mathrm{d} t \leq c^{2} { \int_{0}^{T}} \sqrt {\operatorname{E}[\hat{m}_{t}^{4}] \operatorname{E}[\hat{X}_{t}^{4}] } \,\mathrm{d} t< \infty. $$

Analogously, we obtain that \(\int_{0}^{T} \operatorname{E}[ \hat{m}_{t}^{2} \hat{X}_{t}^{2} (k\sqrt{1- \rho^{2}}\beta_{t} \frac {h_{y}}{h})^{2}] \,\mathrm{d} t <\infty\). From this and Lemma 5.5, we conclude that \(Z^{\hat{\pi},\hat{c}}\) is an \(L^{2}\)-martingale. □

Proof of Proposition 6.4

For any fixed \(\kappa > c + \varrho\), define a metric \(d\) equivalent to \(\|{\,\cdot \,}\|_{\infty}\) by \(d(X,Y):=\mathop {\mathrm {ess}\,\mathrm {sup}}_{\,\mathrm{d} t \otimes \,\mathrm{d} \mathrm{P}} e^{-\kappa (T-t)} |X_{t} - Y_{t}|\). Then \((A,d)\) is a complete metric space. By definition, \(|X_{s}-Y_{s}| \leq e^{\kappa (T-s)} d(X,Y) \,\mathrm{d} t \otimes \,\mathrm{d}\mathrm{P}\)-a.e., so

$$\begin{aligned} e^{-\kappa (T-t)} |(S X)_{t} - (S Y)_{t} | &\leq e^{-\kappa (T-t)} c { \int_{t}^{T}} e^{(s-t)\varrho} e^{\kappa (T-s)} d(X,Y) \,\mathrm{d} s\\ &\leq \frac {c}{\kappa- \varrho } d(X,Y), \end{aligned}$$

and we conclude that \(d(SX,SY) \leq \frac {c}{\kappa- \varrho} d(X,Y)\), where \(\frac {c}{\kappa- \varrho}<1\). Hence \(S\) is a contraction on \((A,d)\). Thus by Banach’s fixed point theorem, there is a unique \(X\in A\) with \(S X = X\), and we have \(d(X_{(n)},X) \leq ( \frac {c}{\kappa-\varrho})^{n} d(X_{(0)},X)\) for all \(n\in \mathbb{N}\). Hence it follows that

$$\begin{aligned} |(X_{(n)})_{t} - X_{t}|&\leq e^{\kappa T} d(X_{(n)},X) \leq \bigg(\frac{c}{\kappa-\varrho}\bigg)^{n} e^{\kappa T} d(X_{(0)},X)\\ &\leq e^{\kappa T} \big(\|X_{(0)}\|_{\infty}+ \|X\|_{\infty}\big) \bigg(\frac{c}{\kappa-\varrho}\bigg)^{n} \end{aligned}$$

and thus \(\|X_{(n)}-X\|_{\infty}\le e^{\kappa T} (\|X_{(0)}\|_{\infty}+ \|X\|_{\infty}) (\frac {c}{\kappa - \varrho} )^{n}\), for every \(n\in \mathbb{N}\) and every choice of \(\kappa>c+\varrho\). Setting \(\kappa=\frac{n+T\varrho}{T}\) for \(n>cT\), we obtain the asserted error bound. □

Appendix B: Stochastic Gronwall inequality

This appendix provides a ramification of the stochastic Gronwall–Bellman inequality which is required for the proofs in this article. Related results can be found in [17, 1, 36]. We work on a general probability space \((\varOmega, \mathcal {F},\mathrm{P})\) that is endowed with a filtration \((\mathcal {F}_{t})_{t \geq 0}\) that is right-continuous and complete.

Proposition B.1

Suppose \(A=(A_{t})_{t\in[0,T]}\) is bounded and progressively measurable, \(Z \in L^{p}(\mathrm{P})\) and \(B=(B_{t})_{t\in[0,T]}\) is a progressively measurable process in \(L^{p}(\,\mathrm{d} t \otimes \,\mathrm{d}\mathrm{P})\) for some \(p>1\). Moreover, let \(X = (X_{t})_{t \in [0,T]}\) be right-continuous and adapted with \(\operatorname{E}[\sup_{t\in [0,T]} |X_{t}|]<\infty\). If

$$ 1_{\{\tau > t\}} X_{t} \geq \operatorname{E}_{t}\left[ 1_{\{\tau > t\}} { \int _{t}^{\tau}} \left(A_{s} X_{s}+ B_{s} \right) \,\mathrm{d} s + 1_{\{\tau > t\}}X_{\tau}\right] \quad \textit{a.s. for }t \in [0,T] $$
(B.1)

for every stopping time \(\tau\) and \(X_{T} \geq Z\), then

$$X_{t} \geq \operatorname{E}_{t} \left[{ \int_{t}^{T}} e^{\int_{t}^{s} A_{u} \,\mathrm{d} u} B_{s} \,\mathrm{d} s + e^{\int_{t}^{T} A_{s} \,\mathrm{d} s} Z \right] \quad \textit{for all } t \in [0,T]\ \textit{a.s.} $$

Proof

We set

$$M_{t}:=\operatorname{E}_{t} \left[{ \int_{0}^{T}} e^{\int_{0}^{s} A_{u} \,\mathrm{d} u} B_{s} \,\mathrm{d} s + e^{ \int_{0}^{T} A_{s} \,\mathrm{d} s}Z\right]. $$

Since \(A\) is bounded above, \(Z\in L^{p}(\mathrm{P})\) and \(B \in L^{p}(\,\mathrm{d} t \otimes \,\mathrm{d}\mathrm{P})\), it follows from Doob’s \(L^{p}\)-inequality that \(\operatorname{E}[\sup_{t\in [0,T]} |M_{t}|^{p}]<\infty\). In particular, \(M\) is well defined as a uniformly integrable martingale. Now set

$$Y_{t}:=e^{- \int_{0}^{t} A_{s} \,\mathrm{d} s} \left(M_{t} - { \int_{0}^{t}} e^{\int_{0}^{s} A_{u} \,\mathrm{d} u} B_{s} \,\mathrm{d} s \right). $$

Since \(A\) is bounded below, we have \(\operatorname{E}[\sup_{t\in[0,T]}|Y_{t}|^{p}]<\infty\), and integration by parts yields

$$ \mathrm{d} Y_{t} = e^{-\int_{0}^{t} A_{s} \,\mathrm{d} s}\big(\,\mathrm{d} M_{t} - e^{\int_{0}^{t} A_{u} \,\mathrm{d} u} B_{t} \,\mathrm{d} t \big) - Y_{t-} A_{t} \,\mathrm{d} t = -(A_{t} Y_{t} + B_{t} ) \,\mathrm{d} t + \,\mathrm{d} N_{t}, $$

where \(N_{t}:=\int_{0}^{t} e^{-\int_{0}^{s} A_{u}\,\mathrm{d} u}\,\mathrm{d} M_{s}\) is a uniformly integrable martingale. For an arbitrary stopping time \(\tau\), we obtain

$$Y_{t} - Y_{\tau}= -{ \int_{0}^{t}} (A_{s}Y_{s} + B_{s}) \,\mathrm{d} s + N_{t} + { \int_{0}^{\tau}} (A_{s} Y_{s} + B_{s})\,\mathrm{d} s - N_{\tau} \quad \text{on } \{\tau>t\}, $$

so that

$$ 1_{\{\tau >t \}} Y_{t} = 1_{\{\tau >t \}} { \int_{t}^{\tau}} (A_{s} Y_{s} + B_{s}) \,\mathrm{d} s + 1_{\{\tau >t \}}(N_{t} - N_{\tau}) + 1_{\{\tau >t \}} Y_{\tau}. $$

Since \((1_{\{\tau >t \}}(N_{t} - N_{\tau}))_{s\in[t,T]}\) is a martingale, it follows that

$$ 1_{\{\tau >t \}} Y_{t} = \operatorname{E}_{t} \left[1_{\{\tau >t \}} { \int_{t}^{\tau}} (A_{s} Y_{s} + B_{s}) \,\mathrm{d} s + 1_{\{\tau >t \}} Y_{\tau}\right]. $$
(B.2)

We set \(\Delta_{t}:=X_{t} -Y_{t}\) and obtain \(\Delta_{T} = X_{T}-Z \ge 0\) and \(\operatorname{E}[\sup_{t\in[0,T]}|\Delta_{t}|]<\infty\). Moreover, (B.1) and (B.2) imply that for any stopping time \(\tau\),

$$ 1_{\{\tau > t\}} \Delta_{t} \geq \operatorname{E}_{t}\left[ 1_{\{\tau > t\}} { \int _{t}^{\tau}} A_{s} \Delta_{s} \,\mathrm{d} s + 1_{\{\tau > t\}}\Delta_{\tau}\right] \quad \text{a.s. for all }t \in [0,T]. $$

Thus Proposition C.2 in [40] applies to yield \(\Delta_{t} \geq 0\) for all \(t\in[0,T]\) a.s., i.e.,

$$X_{t} \geq Y_{t} = e^{- \int_{0}^{t} A_{u} \,\mathrm{d} u} \operatorname{E}_{t} \left[ { \int_{t}^{T}}e^{\int_{0}^{s} A_{u} \,\mathrm{d} u} B_{s} \,\mathrm{d} s + e^{\int_{0}^{T} A_{s} \,\mathrm{d} s} Z \right]. $$

 □

Appendix C: Some facts on parabolic partial differential equations

This appendix collects the relevant results on linear and semilinear parabolic partial differential equations that are used in this article. Following [29], we first introduce the Hölder spaces \(H^{r/2,r}([0,T] \times \mathbb{R}^{d})\) for \(r\in \mathbb{R}_{+}\). For a continuous function \(u: [0,T]\times \mathbb{R}^{d} \to \mathbb{R}, (t,x) \mapsto u(t,x)\), and \(q\in(0,1)\), we define the Hölder coefficient \(\langle u\rangle^{q}_{x}\) in space via

$$\langle u\rangle^{q}_{x} :=\sup_{t \in [0,T],\ x,x' \in \mathbb{R}^{d},\ |x-x'| \leq 1} \frac {|u(t,x) - u(t,x')|}{|x-x'|^{q}} $$

and the Hölder coefficient \(\langle u\rangle^{q}_{t}\) in time via

$$\langle u\rangle^{q}_{t} :=\sup_{t,t' \in [0,T],\ x\in \mathbb{R}^{d},\ |t-t'| \leq 1} \frac {|u(t,x) - u(t,x')|}{|t-t'|^{q}}. $$

The space \(H^{r/2,r}([0,T]\times \mathbb{R}^{d})\) consists of all functions \(u: [0,T]\times \mathbb{R}^{d}\to \mathbb{R}\) that are continuous along with all derivatives \(D^{\alpha}_{t} D^{\beta}_{x} u\) with “order” \(2|\alpha| + |\beta| \le r\) and satisfy \(\|u\|_{H}^{r/2,r} < \infty\). Here the norm \(\|u\|_{H}^{r/2,r}\) of \(u\) is given by

$$\|u\|_{H}^{r/2,r} :=\langle u \rangle^{r/2,r}_{\bullet}+ \sum_{2|\alpha| + |\beta|\leq \lfloor r \rfloor } \| D^{\alpha}_{t} D_{x}^{\beta}u\|_{\infty}, $$

where the mixed space-time Hölder coefficient \(\langle u\rangle^{r/2,r}_{\bullet}\) of \(u\) is given by

$$\langle u \rangle^{r/2,r}_{\bullet} :=\sum_{2|\alpha| + |\beta| = \lfloor r \rfloor } \langle D^{\alpha}_{t} D_{x}^{\beta}u \rangle^{r - \lfloor r \rfloor}_{x} + \sum_{r-2< 2|\alpha|+|\beta|< r} \langle D^{\alpha}_{t} D_{x}^{\beta}u \rangle^{\frac {r-2|\alpha|-|\beta|}{2}}_{t}. $$

Thus \(\|u\|_{H}^{r/2,r}\) sums up the \(L^{\infty}\)-norms of all relevant derivatives plus the Hölder coefficients of the highest-order derivatives. Analogously, for \(r\in \mathbb{R}_{+}\), the space \(H^{r}(\mathbb{R}^{d})\) is defined as the collection of all \(\lfloor r \rfloor\) times continuously differentiable functions \(u: \mathbb{R}^{d} \to \mathbb{R}\) with \(\|u\|_{H}^{r} < \infty\), whereFootnote 6

$$\|u\|_{H}^{r} :=\langle u\rangle^{r}_{\bullet}+ \sum_{|\beta| \leq \lfloor r \rfloor} \| D^{\beta}u \|_{\infty}\quad \text{and}\quad \langle u\rangle^{r}_{\bullet} :=\sum_{|\beta| = \lfloor r \rfloor} \langle D^{\beta}u \rangle^{r- \lfloor r \rfloor}. $$

Linear Cauchy problem

Consider a linear second-order differential operator

$$L u:=\frac {\partial u}{\partial t} - { \sum_{i,j=1}^{d}} a_{ij}(t,x) \frac {\partial^{2} u}{\partial x_{i} \partial x_{j}} - { \sum_{i=1}^{d}} b_{i}(t,x) \frac {\partial u}{\partial x_{i}} - c(t,x)u, $$

where the coefficients \(a,b,c\) are defined on \([0,T]\times \mathbb{R}^{d}\) and \((a_{i,j}(t,x))_{i,j}\) is a symmetric matrix for all \((t,x) \in [0,T]\times \mathbb{R}^{d}\). The main existence and uniqueness result for linear Cauchy problems in \(\mathbb{R}^{d}\), Theorem C.1 below, relies on the following conditions:

\((\mathrm{P1})\) :

The operator \(L\) is uniformly parabolic, i.e., there exist \(0< c_{1}< c_{2}<\infty\) such that for every \((t,x)\in[0,T]\times \mathbb{R}^{d}\), we have

$$c_{1} |y|^{2} \leq { \sum_{i,j=1}^{d}} a_{ij}(t,x) y_{i} y_{j} \leq c_{2} |y|^{2} \quad \text{for all }y\in \mathbb{R}^{d}. $$
\((\mathrm{P2})^{r}\) :

For all \(i,j=1,\dots,d\), we have \(a_{i,j}, b_{i}, c \in H^{r/2,r}([0,T]\times \mathbb{R}^{d})\).

Theorem C.1

Suppose \((\mathrm{P1})\) and \((\mathrm{P2})^{r}\) are satisfied with \(r\in \mathbb{R}_{+}\), \(r\notin \mathbb{N}\), and let \(\varphi\in H^{r+2}(\mathbb{R}^{d})\) and \(f\in H^{r/2,r}([0,T]\times \mathbb{R}^{d})\). Then there exists a unique function \(u\in H^{(r+2)/2,r+2}([0,T]\times \mathbb{R}^{d})\) such that

$$L u = f, \quad u(0,{\,\cdot \,}) = \varphi. $$

Moreover, \(u\) satisfies

$$\|u \|_{H}^{r/2 +1, r+2} \leq c \left( \|\varphi\|_{H}^{r+2} + \|f\|^{r/2, r}_{H} \right), $$

where \(c>0\) is a global constant that is independent of \(\varphi\) and \(f\).

Proof

See [29, Theorem IV.5.1]. □

As a special case, we obtain the result we have used in the proof of Lemma 7.1.

Corollary C.2

Suppose that

(C1):

\(a, b, c: \mathbb{R}\to \mathbb{R}\) are bounded and Lipschitz-continuous;

(C2):

the function \(a\) has a bounded Lipschitz-continuous derivative and satisfies \(\inf_{y \in \mathbb{R}} a (y) >0\);

(\(\mathrm{C3}^{\prime}\)):

\(\hat{\varepsilon}\in H^{r+2}(\mathbb{R})\) for some \(r \in (0,1)\).

Then for each bounded and Lipschitz-continuous function \(f: [0,T] \times \mathbb{R}\to \mathbb{R}\), there exists a unique \(g\in C_{b}^{1,2}([0,T] \times \mathbb{R})\) that solves

$$ 0 = g_{t} + a g_{yy} + b g_{y} + c g + f, \quad g(T,{\,\cdot \,}) = \hat{\varepsilon}. $$

Proof

Consider the second-order differential operator

$$L u = \frac {\partial u}{\partial t} - a \frac {\partial^{2} u}{\partial y \partial y} - b \frac {\partial u}{\partial y} - c u. $$

By assumptions (C1) and (C2), the differential operator \(L\) satisfies (P1) and \((\mathrm{P2})^{r}\) for \(r \in (0,1)\). Moreover, \(f\) is in \(H^{r/2,r}([0,T] \times \mathbb{R})\) since it is Lipschitz-continuous. Hence Theorem C.1 yields \(u\in H^{(r+2)/2,r+2}([0,T] \times \mathbb{R})\) such that

$$Lu = f(T-t, \cdot),\quad u(0,{\,\cdot \,}) = \hat{\varepsilon}\quad\text{and}\quad \|u\|_{C^{1,2}} \leq \|u \|_{H}^{(r+2)/2,r+2} < \infty. $$

Thus defining \(g\in C_{b}^{1,2}([0,T]\times \mathbb{R})\) by \(g(t,y) :=u(T-t, y)\), we obtain

$$0 = g_{t} + a g_{yy} + b g_{y} + c g + f, \quad g(T,\cdot) = \hat{\varepsilon}. $$

 □

Quasilinear Cauchy problem

Next consider the nonlinear differential operator

$$Lu:=u_{t} - { \sum_{i=1}^{d}} \left(\frac {\,\mathrm{d}}{\,\mathrm{d} x_{i}} a_{i}(t,x, u, u_{x}) \right) + a(t,x,u,u_{x}) $$

with principal part in divergence form. We set

$$\begin{aligned} \begin{aligned} a_{ij}(t,x,u,p)&:=\frac {\partial a_{i}(x,t,u,p)}{\partial p_{j}}, \\ A (t,x,u,p)&:=a(t,x,u,p) - { \sum_{i=1}^{d}} \bigg( \frac {\partial a_{i}}{\partial u} p_{i} + \frac {\partial a_{i}}{\partial x_{i}} \bigg). \end{aligned} \end{aligned}$$
(C.1)

We now state the conditions required for Theorem C.3.

  1. (Q1)

    For all \(t\in(0,T]\), \(x,p\in \mathbb{R}^{d}\) and \(u\in \mathbb{R}\), we have

    $${ \sum_{i,j=1}^{d}} a_{ij}(t,x,u,p) y_{i} y_{j} \geq 0\quad \text{for all } y \in \mathbb{R}^{d}. $$
  2. (Q2)

    There exist \(b_{1}, b_{2} \geq 0\) such that for all \(t \in (0,T]\), \(x \in \mathbb{R}^{d}\) and \(u \in \mathbb{R}\), we have

    $$A(t,x,u,0) \geq - b_{1} u^{2} - b_{2}. $$
  3. (Q3)

    The functions \(a\) and \(a_{i}\) are continuous, and \(a_{i}\) is differentiable with respect to \(x\), \(u\) and \(p\). Moreover, there exist \(c_{1}, c_{2} >0\) such that for all tuples \(v =(t,x,u,p)\) in \([0,T]\times \mathbb{R}^{d} \times \mathbb{R}\times \mathbb{R}^{d}\), we have

    $$c_{1}|y|^{2} \leq { \sum_{i,j=1}^{n}} a_{ij}(v) y_{i} y_{j} \leq c_{2} |y|^{2}\quad \text{for all } y\in \mathbb{R}^{d} $$

    and, with \(a_{ij}\) given by (C.1),

$$|a(v)| + { \sum_{i=1}^{d}} \bigg(|a_{i}(v)| + \bigg| \frac {\partial a_{i}(v)}{\partial u}\bigg| \bigg) (1+ |p|) + { \sum_{i,j=1}^{d}} |a_{ij}(v)|\\ \leq c_{2}(1+ |u|) (1 + |p|)^{2}. $$
\((\mathrm{Q4})^{\beta}\) :

There exists \(\beta\in(0,1)\) such that for all compact sets \(K\subset \mathbb{R}\), \(\bar{K} \subset \mathbb{R}^{d}\), the functions

$$a_{i}, a, a_{ij}, \frac{\partial a_{i}}{\partial u}, \frac{\partial a_{i}}{\partial x_{i}}: [0,T] \times \mathbb{R}^{d} \times K \times \bar{K} \to \mathbb{R}$$

are Hölder-continuous in \(t,x,u\) and \(p\) with exponents \(\beta/2\), \(\beta\), \(\beta\) and \(\beta\), respectively.

Here we say that \(f:[0,T] \times \mathbb{R}^{d} \times K \times \bar{K} \to \mathbb{R}, z= (z^{1},z^{2}, z^{3}, z^{4})\mapsto f(z)\) is \(\beta\)-Hölder-continuous in \(z^{i}\) if

$$\langle u\rangle^{\beta}_{i}:=\sup_{z, \bar{z} \in \operatorname{dom}(f),\ z^{j} = \bar{z}^{j}, \ j \neq i ,\ |z^{i}-\bar{z}^{i}| \leq 1} \frac {|f(z) - f(\bar{z})|}{|z^{i} - \bar{z}^{i}|^{\beta}} < \infty. $$

Theorem C.3

Suppose \(\psi_{0}\) is in \(H^{2+\beta}(\mathbb{R}^{d})\) and (Q1), (Q2), (Q3) and (Q4)β are satisfied for some \(\beta \in (0, 1)\). Then there exists a solution \(u \in H^{(2+ \beta)/2, 2+ \beta}([0,T] \times \mathbb{R}^{d})\) of the Cauchy problem

$$Lu = 0,\qquad u(0,\cdot) =\psi_{0}. $$

Proof

See [29, Theorem V.8.1]. □

In the proof of Theorem 6.10, we require the following ramification of this result.

Corollary C.4

Suppose that

(C1):

\(a, b, c: \mathbb{R}\to \mathbb{R}\) are bounded and Lipschitz-continuous;

(C2):

the function \(a\) has a bounded Lipschitz-continuous derivative and satisfies \(\inf_{y \in \mathbb{R}} a (y) >0\);

(\(\mathrm{C3}^{\prime}\)):

\(\hat{\varepsilon}\in H^{r+2}(\mathbb{R})\) for some \(r \in (0,1)\);

and let \(f\in C^{1}_{b}(\mathbb{R})\). Then the semilinear PDE

$$ 0 = g_{t} + ag_{yy} +b g_{y} + cg + f(g), \qquad g(T,\cdot) = \hat{\varepsilon}$$

has a solution \(g\in C_{b}^{1,2}([0,T]\times \mathbb{R})\).

Proof

After setting \(a_{1}(t,x,u,p):=p a(x)\) and

$$ \bar{a}(t,x,u,p):=-b(x) p - c(x) u - f(u) + p a'(x), $$

we can represent the relevant differential operator as

$$\begin{aligned} L u :=& u_{t} - \frac {\,\mathrm{d}}{\,\mathrm{d} x} a_{1}(t,x,u, u_{x}) + \bar{a}(t,x,u, u_{x})\\ \ =& u_{t} - \frac {\,\mathrm{d}}{\,\mathrm{d} x} \big(u_{x} a(x)\big) - b(x) u_{x} -c(x) u - f(u) + u_{x} a'(x)\\ \ =& u_{t} - a(x) u_{xx} - b(x) u_{x} -c(x) u - f(u). \end{aligned}$$

Hence if \(u \in C_{b}^{1,2}([0,T]\times \mathbb{R})\) solves \(Lu=0\), \(u(0,\cdot)=\hat{\varepsilon}\), then \(g(t,x):=u(T-t,x)\) defines a member of \(C_{b}^{1,2}([0,T]\times \mathbb{R})\) that satisfies

$$ 0= g_{t}+ a g_{yy} + bg_{y} +cg + f(g),\quad g(T,\cdot) = \hat{\varepsilon}. $$

We now verify the assumptions of Theorem C.3 for \(L\). Note that

$$a_{11}(t,x,u,p) = \frac {\partial a_{1}(x,t,u,p)}{\partial p_{1}} =a(x), $$

so (Q1) holds since

$$a_{11}(t,x,u,p) y^{2} = a(x) y^{2} \geq 0 \quad \text{by (C2).} $$

Next observe that

$$\begin{aligned} A(t,x,u,p) &= \bar{a}(t,x,u,p) - \frac {\partial a_{1}(t,x,u,p)}{\partial u} p - \frac {\partial a_{1}(t,x,u,p)}{\partial x} \\ &= - b(x) p -c (x) u - f(u). \end{aligned}$$

Thus (Q2) is satisfied since

$$A(t,x,u,0) = -c (x) u - f(u) \geq - \|c\|_{\infty}|u| - \|f\|_{\infty}\geq - b_{1} u^{2} - b_{2} $$

with \(b_{1}:=\|c\|_{\infty}\) and \(b_{2}:=\|c\|_{\infty}+ \|f\|_{\infty}\). To check (Q3), note that by (C1) and (C2), the functions \(a_{1}\) and \(\bar{a}\) are continuous, and \(a_{1}\) is differentiable; moreover,

$$\inf_{x \in \mathbb{R}} a(x) |y|^{2} \leq a_{11}(t,x,u,p) y^{2} \leq \|\beta\|_{\infty}|y|^{2} $$

for all \(t \in [0,T]\) and \(x,u,p,y \in \mathbb{R}\). For all \(v = (t,x,u,p) \in [0,T] \times \mathbb{R}\times \mathbb{R}\times \mathbb{R}\), we further have

$$\begin{aligned} & |\bar{a}(v)| + \bigg(|a_{1}(v)| + \bigg| \frac {\partial a_{1}(v)}{\partial u}\bigg|\bigg) (1+ |p|) + |a_{11}(v)|\\ &\quad{}\leq \|b\|_{\infty}|p| + \|c\|_{\infty}|u| + \|f\|_{\infty}+ \|a'\|_{\infty}|p| + |p| \|a\|_{\infty}(1+ |p|) + \|a\|_{\infty}\\ &\quad{}\leq \big(\|a\|_{\infty}+ \|b\|_{\infty}+ \|c\|_{\infty}+ \|f\|_{\infty}+ \|a'\|_{\infty}\big) (1+ |u|) (1 + |p|)^{2}, \end{aligned}$$

since \(a,b,c,f\) and \(a'\) are bounded. Thus (Q3) holds with

$$c_{2}:=\|a\|_{\infty}+ \|b\|_{\infty}+ \|c\|_{\infty}+ \|f\|_{\infty}+ \|a'\|_{\infty}$$

and \(c_{1}:=\inf_{x \in \mathbb{R}} a(x) >0\). Finally, for any compact set \(K \subset \mathbb{R}\), the functions

$$\begin{aligned} &a_{1}(v) =p a(x), \quad a(v)=-b(x) p - c(x) u - f(u) + p a'(x),\\ &a_{11}(v)= a(x),\quad \frac{\partial a_{1}}{\partial u}(v)=0, \quad \frac{\partial a_{1}}{\partial p}(v) = a'(x) \end{aligned}$$

restricted to \([0,T] \times \mathbb{R}\times K \times K\) are Lipschitz-continuous in \(x\), \(u\) and \(p\), because \(a\), \(a'\), \(b\), \(c\) and \(f\) are bounded and Lipschitz by (C1), (C2) and since \(f \in C_{b}^{1}(\mathbb{R})\). Hence (Q4)\(^{\frac{1}{2}}\) holds as well. Thus by Theorem C.3, the Cauchy problem

$$L u = 0, \quad u(0,\cdot) =\hat{\varepsilon}$$

has a solution \(u \in H^{5/4,5/2}([0,T] \times \mathbb{R}^{d}) \subset C_{b}^{1,2} ([0,T] \times \mathbb{R}^{d})\). Uniqueness follows from standard BSDE arguments; see e.g. [20, Proposition 4.3]. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kraft, H., Seiferling, T. & Seifried, F.T. Optimal consumption and investment with Epstein–Zin recursive utility. Finance Stoch 21, 187–226 (2017). https://doi.org/10.1007/s00780-016-0316-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-016-0316-0

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation