Skip to main content
Log in

On the dynamic representation of some time-inconsistent risk measures in a Brownian filtration

  • Published:
Mathematics and Financial Economics Aims and scope Submit manuscript

Abstract

It is well-known from the work of Kupper and Schachermayer that most law-invariant risk measures are not time-consistent, and thus do not admit dynamic representations as backward stochastic differential equations. In this work we show that in a Brownian filtration the “Optimized Certainty Equivalent” risk measures of Ben-Tal and Teboulle can be computed through PDE techniques, i.e. dynamically. This can be seen as a substitute of sorts whenever they lack time consistency, and covers the cases of conditional value-at-risk and monotone mean-variance. Our method consists of focusing on the convex dual representation, which suggests an expression of the risk measure as the value of a stochastic control problem on an extended the state space. With this we can obtain a dynamic programming principle and use stochastic control techniques, along with the theory of viscosity solutions, which we must adapt to cover the present singular situation.

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.

Similar content being viewed by others

Notes

  1. In fact, the conditional value-at-risk (also known as expected shortfall) has been praised, and its adoption recommended, by the Basel III Committee in the following terms [34, p. 18]:

    “... the current framework’s reliance on VaR (value-at-risk) as a quantitative risk metric raises a number of issues, most notably the inability of the measure to capture the “tail risk” of the loss distribution. The Committee has therefore decided to use an expected shortfall (ES) measure for the internal models -based approach and will determine the risk weights for the revised standardised approach using an ES methodology...”

  2. We denote by \(l^*\) the convex conjugate of l.

  3. As pointed out by an anonymous referee, in some cases the upper lateral boundary condition may be expensive to obtain, since it involves an expected value. However, in the situation when \(\text {dom}(l^*)=[0,\infty )\), this boundary condition disappears and only the lower lateral boundary condition remains. The latter takes the form \(V(t,y,0)=-l^*(z)\) and is hence explicit.

  4. \(\rho (X+ c) = \rho (X) +c\) for all \(X \in L^\infty ({{{\mathcal {F}}}})\) and \(c \in {\mathbb {R}} \). Translation invariance is a synonym for this.

  5. In fact, it is \({\tilde{\rho }}(X):= \rho (-X)\) that satisfies the risk measures axioms as developed in Artzner et al. [1] and Föllmer and Schied [24], but we will work with the increasing functional \(\rho \) for ease of notation.

  6. This condition is also known as the flow property. We refer to Cheridito et al. [11], Delbaen [15], Artzner et al. [2], Ruszczyński and Shapiro [40], Detlefsen and Scandolo [17] for discussions on the consequences of time-consistency.

  7. This corresponds to the standard OCE risk measure up to a minus sign.

  8. \(A'\) is the transpose of A.

  9. To exemplify: Take \(m=d=1\) and \(\Gamma ^n_{1,2}=1,\Gamma ^n_{2,2}=-1/n\). Then \(H(t,y,1,\gamma ,\Gamma ^n)<\infty \) but in the limit (i.e. \(\Gamma _{1,2}=1,\Gamma _{2,2}=0\)) the Hamiltonian is infinite. Thus \(H<\infty \) is not closed and so there cannot be such continuous G.

  10. See also Remark 3.5 for some comments on the growth properties of V.

  11. The essential range of \(\eta Z\) is \(range(\eta Z):=[\mathop {\mathrm{ess}\,\mathrm{inf}}\eta Z,\mathop {\mathrm{ess}\,\mathrm{sup}} \eta Z]\).

  12. This is definitely a technical point, and we see no reason why Theorem 1.1 would not hold without such assumption on the domain of \(l^*\). In fact, during revision of the present article, the authors and A. Max Reppen [3] derived existence and uniqueness for a related PDE in the case when the domain of \(l^*\) is bounded, but under a different notion of viscosity sub/super solution.

  13. I.e. for all \(x_0=(s_0,y_0,z_0)\in [0,T]\times {\mathbb {R}}^m\times (0,+\infty )\) and \(\varphi \in C^2([0,T]\times {\mathbb {R}}^m\times (0,+\infty ))\) such that \(x_0\) is a local minimizer of \(w-\varphi \) and \(\varphi (x_0) = w(x_0)\), we have \(w(x_0)\ge \psi (y_0,z_0)\) if \(s_0=T\), and otherwise \( \partial _t\varphi (x_0) + H^n(x_0, D\varphi (x_0), D^2\varphi (x_0)) \le 0\).

References

  1. Artzner, P., Delbaen, F., Eber, J.M., Heath, D.: Coherent measures of risk. Math. Finance 9, 203–228 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  2. Artzner, P., Delbaen, F., Eber, J.-M., Heath, D., Ku, H.: Coherent multiperiod risk adjusted values and Bellman’s principle. Ann. Oper. Res. 152, 5–22 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  3. Backhoff-Veraguas, J., Reppen, M., Tangpi, L.: Stochastic control of optimized certainty equivalent. Preprint, (2019). arXiv:2001.10108

  4. Bäuerle, N., Ott, J.: Markov decision processes with average-value-at-risk criteria. Math. Methods Oper. Res. 74(3), 361–379 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  5. Ben-Tal, A., Teboulle, M.: Expected utility, penalty functions, and duality in stochastic nonlinear programming. Manag. Sci. 32(11), 1445–1466 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  6. Ben-Tal, A., Teboulle, M.: An old-new concept of convex risk measures: the optimized certainty equivalent. Math. Finance 17(3), 449–476 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  7. Benth, F.E., Karlsen, K.H., Reikvam, K.: A semilinear black and scholes partial differential equation for valuing american options. Finance Stoch. 7(3), 277–298 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bokanowski, O., Bruder, B., Maroso, S., Zidani, H.: Numerical approximation for a superreplication problem under gamma constraints. SIAM J. Numer. Anal. 47(3), 2289–2320 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  9. Bokanowski, O., Picarelli, A., Zidani, H.: State-constrained stochastic optimal control problems via reachability approach. SIAM J. Control Optim. 54(5), 2568–2593 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  10. Bouveret, G., Chassagneux, J.-F.: A comparison principle for PDEs arising in approximate hedging problems: application to Bermudan options. Appl. Math. Optim., pp 1–23 (2017)

  11. Cheridito, P., Delbaen, F., Kupper, M.: Dynamic monetary risk measures for bounded discrete-time processes. Electron. J. Probab. 11(3), 57–106 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  12. Coquet, F., Hu, Y., Mémin, J., Peng, S.: Filtration-consistent nonlinear expectations and related \(g\)-expectations. Probab. Theory Related Fields 123, 1–27 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Da Lio, F., Ley, O.: Uniqueness results for second-order Bellman–Issacs equations under quadratic growth assumptions and applications. SIAM J. Control Optim. 45, 74–106 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  14. Da Lio, F., Ley, O.: Convex Hamilton–Jacobi equations under superlinear growth conditions on data. Appl. Math. Optim. 63(3), 309–339 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  15. Delbaen, F.: The structure of \(m\)-stable sets and in particular of the set of risk neutral measures. In: Memoriam Paul-André Meyer–Seminaire de Probabilités XXXIX, pp. 215–258. Springer, Berlin (2006)

  16. Delbaen, F., Peng, S., Rosazza Gianin, E.: Representation of the penalty term of dynamic concave utilities. Finance Stoch. 14, 449–472 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  17. Detlefsen, K., Scandolo, G.: Conditional and dynamic convex risk measures. Finance Stoch. 9(4), 539–561 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  18. Drapeau, S., Mainberger, C.: Stability and Markov properties of foward backward minimal supersolutions. Electron. J. Probab. 21(41), 1–15 (2016)

    MATH  Google Scholar 

  19. Drapeau, S., Kupper, M., Papapantoleon, A.: A Fourier approach to the computation of CVaR and optimized certainty equivalents. J. Risk 16(6), 3–29 (2014)

    Article  Google Scholar 

  20. Drapeau, S., Kupper, M., Gianin, E.R., Tangpi, L.: Dual representation of minimal supersolutions of convex BSDEs. Ann. Inst. H. Poincaré Probab. Stat. 52(2), 868–887 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  21. Ekeland, I., Lazrak, A.: The golden rule when preferences are time inconsistent. Math. Financ. Econ. 4(1), 29–55 (2010)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  23. Fleming, W.H., Soner, H.M.: Controlled Markov processes and viscosity solutions, vol. 25 of Stochastic Modelling and Applied Probability, second edition. Springer, New York (2006)

  24. Föllmer, H., Schied, A.: Stochastic finance. Walter de Gruyter & Co., Berlin, extended edition. An introduction in discrete time (2011)

  25. Friedman, A.: Partial Differential Equations of Parabolic Type. Prentice-Hall Inc, Englewood Cliffs (1964)

    MATH  Google Scholar 

  26. Ikeda, N., Watanabe, S.: Stochastic differential equations and diffusion processes, vol. 24 of North-Holland Mathematical Library, second edition . North-Holland Publishing Co., Amsterdam; Kodansha, Ltd., Tokyo(1989)

  27. Karnam, C., Ma, J., Zhang, J.: Dynamic approach for some time inconsistent problems. Ann. Appl. Probab. 27(6), 3435–3477 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  28. Kunita, H.: Stochastic differential equations based on lévy processes and stochastic flows of diffeomorphisms. In: Real and stochastic analysis, pp. 305–373. Springer, Berlin (2004)

  29. Kupper, M., Schachermayer, W.: Representation results for law invariant time consistent functions. Math. Financ. Econ. 2(3), 189–210 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  30. Maccheroni, F., Marinacci, M., Rustichini, A.: Ambiguity aversion, robustness, and the variational representation of preferences. Econometrica 74(6), 1447–1498 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  31. Mataramvura, S., Øksendal, B.: Risk minimizing portfolios and HJBI equations for stochastic differential games. Stochastics 80(4), 317–337 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  32. Miller, C.W., Yang, I.: Optimal control of conditional value-at-risk in continuous time. SIAM J. Control Optim. 55(2), 856–884 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  33. Neufeld, A., Nutz, M.: Nonlinear Lévy processes and their characteristics. Trans. AMS 369(1), 69–95 (2017)

    Article  MATH  Google Scholar 

  34. On Banking Supervision, B.C.: Fundamental review of the trading book: a revised market risk framework. Bank for international settlments (2014)

  35. Pflug, G.C., Pichler, A.: Time-inconsistent multistage stochastic programs: martingale bounds. Eur. J. Oper. Res. 249(1), 155–163 (2016a)

    Article  MathSciNet  MATH  Google Scholar 

  36. Pflug, G.C., Pichler, A.: Time-consistent decisions and temporal decomposition of coherent risk functionals. Math. Oper. Res. 41(2), 682–699 (2016b)

    Article  MathSciNet  MATH  Google Scholar 

  37. Pham, H.: Continuous-time stochastic control and optimization with financial applications. Stochastic Modelling and Applied Probability, vol. 61. Springer, Berlin (2009)

  38. Rockafellar, R., Uryasev, S.: Conditional value-at-risk for general loss distributions. J. Bank. Finance, pp. 1443–1471 (2002)

  39. Rouge, R., El Karoui, N.: Pricing via utility maximization and entropy. Math. Finance 10(2), 259–276 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  40. Ruszczyński, A., Shapiro, A.: Conditional risk mappings. Math. Oper. Res. 31(3), 544–561 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  41. Shapiro, A.: On a time consistency concept in risk averse multistage stochastic programming. Oper. Res. Lett. 37(3), 143–147 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  42. Touzi, N.: Optimal stochastic control, stochastic target problems, and backward SDE, vol. 29 of Fields Institute Monographs. Springer, New York; Fields Institute for Research in Mathematical Sciences, Toronto, ON, 2013. With Chapter 13 by Angès Tourin

  43. Vargiolu, T.: Existence, uniqueness and smoothness for the Black–Scholes–Barenblatt equation. Universita di Padova, Department of Pure and Applied Mathematics, Rapporto Interno, (5) (2001)

  44. Yong, J., Zhou, X.: Stochastic Controls, Hamiltonian Systems and HJB Equations. Springer, New York (2000)

    MATH  Google Scholar 

  45. Zhou, X.Y., Li, D.: Continuous-time mean-variance portfolio selection: a stochastic LQ framework. Appl. Math. Optim. 42(1), 19–33 (2000)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Beatrice Acciaio, Joaquín Fontbona, Asgar Jamneshan and Michael Kupper for their feedback on this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julio Backhoff-Veraguas.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

Proof of Proposition 2.4

In our language, we easily obtain \( \rho ^{l_z}(zX)=z\log \left( \frac{ E[e^{X}]}{z}\right) +z - 1\), and so for the value function, see Proposition 3.3, we have

$$\begin{aligned} V(s,y,z)&=-z\log z +z - 1+ z\log E \left[ \exp \left\{ f(Y_T^{s,y}) + \int _s^T g(t,Y_t^{s,y})dt \right\} \right] \\&=: -z\log z+z - 1 + z {\tilde{V}}(s,y). \end{aligned}$$

That \(\exp {({\tilde{V}})}\) is a viscosity solution of the backward Kolmogorov PDE associated to the diffusion Y follows e.g. from Fleming and Soner [23, Chap. V.9]. That \(\exp {({{\tilde{V}}})}\) is the uniqueness classical solution follows by e.g. Friedman [25, Theorems 1.7.12 and 2.4.10]. It is then clear that V is the unique classical solution of our HJB equation. \(\square \)

Proof of Lemma 3.1

A derivation of the dual representation

$$\begin{aligned} \rho (X)\,=\, \sup \left\{ E[XZ] -E[l^*( Z)]:Z\in L^{1}_{+} ,Z\in \text {dom}(l^*),\,E[ Z]=1 \right\} ,\quad X \in L^\infty \end{aligned}$$
(4.4)

can be obtained for instance from Ben-Tal and Teboulle [6], and elementary considerations. Deriving (3.1) from (4.4) is done by classical arguments which we give for completeness. We clearly have “\(\ge \)” in (3.1). Conversely, given \(\varepsilon >0\) and a feasible Z for (4.4) such that \(\rho (X)\le E[XZ] - E[l^*(Z)]-\varepsilon \), we must have \(l^*(Z)< \infty \). For every \(c\in (0,1)\), define \(Z^{c}:= c +(1-c)Z\), which is likewise feasible for (4.4), satisfies \(l^*(Z^c)<\infty \) and is such that \((Z^c)_c\) is uniformly integrable. Assume for the moment that \(l^*(Z^{c})\rightarrow l^*(Z)\)P-a.s. Then, by convexity, \(0\le l^*(Z^c)\le (1-c)l^*(Z^c)\le l^*(Z)\). Thus, we conclude by dominated convergence that \(\rho (X) \le \lim _{c\rightarrow 0}E[XZ^c]-E[l^*(Z^c)]-\varepsilon \), which yields the reverse inequality. The proof is finished after noticing that \(l^*\) must be continuous throughout its domain. In fact, the domain of \(l^*\) is an interval with end points denoted by \(a\in {\mathbb {R}}_+\) and \(b\in {\mathbb {R}}_+\cup \{+\infty \}\), and \(l^*\) is continuous on (ab). If \(+\infty >b \in \text {dom}(l^*)\), let \(x^n\rightarrow b\) and \(\lambda ^n\in (0,1)\) such that \(\lambda ^n\rightarrow 1\) and \(x^n= \lambda ^nb + (1-\lambda ^n)a\). By convexity, we have \(l^*(x^n) \le \lambda ^nl^*(b)+(1-\lambda ^n)l^*(a)\). This shows \(\limsup _{n\rightarrow \infty } l^*(x^n)\le l^*(b)\). We conclude by lower semicontinuity that \(l^*\) is continuous at b and the proof is similar if \(a\in \text {dom}(l^*)\).

Proof of Proposition 3.2

By the dual representation of \(\rho (X)\) in the completed Brownian filtration, we directly see that the r.h.s. of (3.2) is a lower bound for \(\rho (X)\). We shall establish the opposite inequality by repeated approximation arguments. By Lemma 3.1 we need only consider \(Z\in {{{\mathcal {Z}}}}\).

STEP 1: We may assume w.l.o.g. that \(E[l^*(Z)]<\infty \), as otherwise this Z is irrelevant for the problem. Letting \(Z_t:=E[Z|{\mathcal {F}}_t]\) and \(\tau ^n:=\inf \{0<t\le T:Z_t=n\}\wedge T\), we have by optional sampling and Jensen’s inequality that \(-E[l^*(Z)]\le -E[l^*(Z^{n})] \) and \(E[l^*(Z^{n})]<\infty \), with \(Z^n:=Z_{\tau ^n}\le n\). On the other hand \(E[Z^nX]\rightarrow E[ZX] \) by the martingale convergence theorem. Thus

$$\begin{aligned} E[ZX]-E[l^*(Z)]\le \liminf _{n\rightarrow \infty }(E[Z^nX]-E[l^*(Z^n)]). \end{aligned}$$

So we assume w.l.o.g. that Z is essentially bounded from above and essentially bounded away from 0.

STEP 2: Define \(Z^n:=E[Z|{\mathcal {G}}_n]\), where \({\mathcal {G}}_n=\sigma \{W_{kT2^{-n}}:k=0,\dots ,2^n\}\). It holds

$$\begin{aligned} Z^n = u_n(W_{T/n},\dots ,W_{kT/n} ,\dots ,W_{T} ), \end{aligned}$$

for some bounded positive Borel function \(u_n\) which is bounded away from 0 and with range in \(dom(l^*)\). Moreover, \(Z^n\rightarrow Z\)P-a.s. and by continuity of \(l^*\) in its domain (see the proof of Lemma 3.1), \(l^*(Z^n)\) is essentially bounded uniformly in n. Using martingale convergence again and dominated convergence we have \( E[ZX]-E[l^*(Z)] = \lim _{n\rightarrow \infty }(E[Z^nX]-E[l^*(Z^n)])\). Thus, we may further assume w.l.o.g. that Z is of the form

$$\begin{aligned} Z = u(W_{T/n},\dots ,W_{kT/n} ,\dots ,W_{T} )\quad \text {for some } n\in {\mathbb {N}}, \end{aligned}$$
(4.5)

with u as above.

STEP 3: Let \(\Phi \in C^\infty _c({\mathbb {R}}^n) \) be the mollifier \(\textstyle \Phi (x)={\mathbb {I}}_{\{\Vert x\Vert <1\}}\lambda \exp \left\{ \frac{1}{\Vert x\Vert ^2-1} \right\} \), with \(\lambda \ge 0\) such that \(\int _{{\mathbb {R}}^n}\Phi (x)\,dx=1\). Define by convolution \(u^{\delta }(x):= u*\delta ^{-n}\Phi (x/\delta ) \), \(\delta >0\). Then \(u^\delta \in C^\infty ({\mathbb {R}}^n)\) is bounded from above, bounded away from 0 and the derivative \( \nabla u^\delta = u*\nabla (\delta ^{-n}\Phi (\cdot /\delta )) \) is bounded. We can also choose a sequence \(\delta _m\rightarrow 0\) so that \(u^{\delta _m} \rightarrow u\), Lebesgue a.e. in \({\mathbb {R}}^n\); this follows from the convergence over compacts of \(u^\delta \) to u in \(L^1(dx)\) and a diagonalization argument. This shows that \(u^{\delta _m}(W_{T/n},\dots ,W_{kT/n} ,\dots ,W_{T} )\) converges to Z almost surely as \(m\rightarrow \infty \), since the law of the Gaussian vector \((W_{T/n},\dots ,W_{kT/n} ,\dots ,W_{T} )\) is equivalent to Lebesgue in \({\mathbb {R}}^n\). Arguing as in the previous step, we conclude that we may further assume w.l.o.g. that Z is given by (4.5) where u is smooth and with bounded derivatives.

STEP 4: From the previous steps, it remains to show that for every n, the random variable \(Z = u(W_{T/n},\dots ,W_{kT/n} ,\dots ,W_{T} )\) can be written as \({{{\mathcal {E}}}}(\int \beta \,dW)_T\), with \(\beta \in {{{\mathcal {L}}}}_b\). For \(t\in \left[ \frac{n-1}{n}T,T\right] \) we have

$$\begin{aligned} E[Z|{\mathcal {F}}_t]&= \int u(W_{T/n},\dots ,W_{T(n-1)/n},W_{t}+x)] N^{T-t}(dx) \\&=: U^n\left( t,W_t\,;\,W_{T/n},\dots ,W_{\frac{n-1}{n}T}\right) , \end{aligned}$$

with \(N^{T-t}(dx)\) the law of a centred Gaussian with variance \((T-t)\times I_d\), with \(I_d\) the identity of \({\mathbb {R}}^{d\times d}\). By the mean value theorem and dominated convergence, the function \(U^n\) is differentiable in the spacial arguments and the derivatives are bounded, uniformly in the time argument. Smoothness in the time argument is apparent from the density of \(N^{T-t}(dx)\). In addition, \(U^n\) is bounded away from 0.

We now proceed by reverse induction. Assume that we have constructed a function \(U^{k+1}\) such that \(E[Z|\mathcal{F}_t]=U^{k+1}(t,W_t\,;\,W_{T/n},\dots ,W_{\frac{k}{n}T})\) on \([kT/n,(k+1)T/n]\) with \(U^{k+1}\) smooth, bounded from above and away from zero, as well as having bounded derivatives in \((x,w_1,\dots ,w_{k})\) uniformly in time. By the tower property, for \(t\in [(k-1)T/n,kT/n]\) we have:

$$\begin{aligned} E[Z|{\mathcal {F}}_t]&= E\left[ U^{k+1}(kT/n,W_{kT/n}\,;\,W_{T/n},\dots ,W_{T(k-1)/n},W_{kT/n}) |{\mathcal {F}}_t\right] \\&= \int U^{k+1}(kT/n,W_{t}+x\,;\,W_{T/n},\dots ,W_{T(k-1)/n},W_{t}+x)] N^{(kT/n)-t}(dx) \\&=: U^k\left( t,W_t\,;\,W_{T/n},\dots ,W_{\frac{k-1}{n}T}\right) . \end{aligned}$$

By essentially the same argument as above, \(U^k\) is smooth in time and space arguments, bounded from above and away from zero, and has bounded derivatives in \((x,w_1,\dots ,w_{k-1})\) uniformly in time.

Using Itô’s formula and by uniqueness in the martingale representation, we obtain that

$$\begin{aligned} \beta _t=\left[ { U^k\left( t,W_t\,;\,W_{T/n},\dots ,W_{\frac{k-1}{n}T}\right) }\right] ^{-1}{\nabla _x U^k\left( t,W_t\,;\,W_{T/n},\dots ,W_{\frac{k-1}{n}T}\right) }, \end{aligned}$$

on \(t\in [(k-1)T/n,kT/n]\). Since there are only finitely many such intervals for fixed n, it follows that \(\beta \) is essentially bounded. This concludes the proof. \(\square \)

We close the appendix including a technical lemma which was crucial in the proof of our main theorem. We assume \({{{\mathcal {O}}}}=(0,+\infty )\) from here on. The argument is inspired by [43].

Lemma 4.4

Let \(w:[0,T]\times {\mathbb {R}}^m\times [0,+\infty ) \) be a continuous viscosity supersolution ofFootnote 13

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _tu + H^n(t, y,z, Du, D^2u) = 0,\quad (t,y,z)\in [0,T)\times {\mathbb {R}}^m\times (0,+\infty )\\ u(T,y,z) = \psi (y,z),\quad (y,z)\in {\mathbb {R}}^m\times (0,+\infty ). \end{array}\right. } \end{aligned}$$
(4.6)

Then \({{\tilde{w}}}(s,y,z):=w(s,y,|z|)\) is a continuous viscosity supersolution of

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _tu + H^n(t, y,z, Du, D^2u) = 0,\quad (t,y,z)\in [0,T)\times {\mathbb {R}}^m\times {\mathbb {R}}\\ u(T,y,z) = \psi (y,|z|),\quad (y,z)\in {\mathbb {R}}^m\times {\mathbb {R}}. \end{array}\right. } \end{aligned}$$
(4.7)

The analogue statement holds for continuous viscosity (sub)solutions.

Proof

Let \(x_0=(s_0,y_0,z_0)\in [0,T]\times {\mathbb {R}}^m\times {\mathbb {R}}\) and \(\varphi \in C^2([0,T]\times {\mathbb {R}}^m\times {\mathbb {R}})\) such that \(x_0\) is a local minimizer of \({{\tilde{w}}}-\varphi \) and \(\varphi (x_0) = {{\tilde{w}}}(x_0)\). If \(s_0=T\), and otherwise if \(z_0 > 0\), there is nothing to prove, owing to the supersolution property of w. Otherwise, if \(z_0<0\), we observe that the terms \(\partial _{zz}\varphi \) and \(z\partial _{yz}\varphi \) are of second-order in the z-variable, meaning that we can leverage again the supersolution property of w to easily obtain \( \partial _t\varphi (x_0) + H^n(x_0, D\varphi (x_0), D^2\varphi (x_0)) \le 0\).

It remains to check the case \(s_0<T\) and \(z_0=0\). As conventional in this theory, we may further assume that \(x_0\) was a strict local minimizer of \({{\tilde{w}}}-\varphi \). By definition, \(x_0\) is also a strict local minimizer of \( w-\varphi \) restricted to \([0,T]\times {\mathbb {R}}^m\times [0,+\infty )\). By Lemma 4.5 below, we have \(\partial _t\varphi (x_0) + H^n(x_0, D\varphi (x_0), D^2\varphi (x_0)) \le 0\), finishing the proof for supersolution. Of course the same arguments apply to subsolution, and then to solutions too. \(\square \)

The above proof rests on the following lemmas, which are reminiscent of [7, Lemma 4.1]. We only provide the proof of Lemma 4.5, as the one for Lemma 4.6 is straightforward and lengthy.

Lemma 4.5

Take w as in Lemma 4.4 and let \(\varphi \) be of class \(C^{2}\) on \([0,T]\times {\mathbb {R}}^m\times [0,+\infty )\). Suppose that \(w-\varphi \) has a strict local minimum at \(x_0:=(s_0,y_0,0)\). Then \(\partial _t\varphi (x_0) + H^n(x_0, D\varphi (x_0), D^2\varphi (x_0)) \le 0\).

Proof

Define \(v:=w-\varphi \). We apply Lemma 4.6, obtaining the existence of \(x_\epsilon :=(s_\epsilon ,y_\epsilon ,z_\epsilon )\rightarrow x_0:=(s_0,y_,0)\) with the properties listed therein. In particular, \(x_\epsilon \) is a local minimum of \(w-\varphi _\epsilon \) with \(z_\epsilon >0\), where \(\varphi _\epsilon (s,y,z):=\varphi (s,y,z)-\epsilon /z\). By definition of w, we have

$$\begin{aligned} \begin{aligned}&\partial _t\varphi _\epsilon (x_\epsilon )+b(s_\epsilon ,y_\epsilon )\partial _y\varphi _\epsilon (x_\epsilon )+\frac{1}{2}tr\left( \sigma (s_\epsilon ,y_\epsilon )\sigma (s_\epsilon ,y_\epsilon )'\partial ^2_{yy}\varphi _\epsilon (x_\epsilon )\right) \\&\quad +\textstyle \sup _{|\beta |\le n}\left[ \frac{1}{2} z_\epsilon ^2|\beta |^2\partial ^2_{zz}\varphi _\epsilon (x_\epsilon )+ z_\epsilon \,\partial ^2_{yz}\varphi _\epsilon (x_\epsilon ) \sigma (s_\epsilon ,y_\epsilon )\beta -\frac{\epsilon |\beta |^2}{z_\epsilon }\right] + z_\epsilon g(s_\epsilon ,y_\epsilon ) \le 0. \end{aligned} \end{aligned}$$

We conclude by sending \(\epsilon \rightarrow 0\), by continuity and since the supremum defining the Hamiltonian \(H^n\) is taken over a compact set. \(\square \)

Lemma 4.6

Let \(v:[0,T]\times {\mathbb {R}}^m\times [0,\infty )\rightarrow {\mathbb {R}}\) be continuous and such that \((s_0,y_,0)\) is a strict local minimizer of v. Then there exists \((s_\epsilon ,y_\epsilon ,z_\epsilon )\rightarrow (s_0,y_,0)\) such that

$$\begin{aligned}&(1)\, z_\epsilon>0\, (\forall \epsilon >0),\quad (2)\, \frac{\epsilon }{z_\epsilon }\rightarrow 0 \text { as } \epsilon \searrow 0, \quad (3)\, (s_\epsilon ,y_\epsilon ,z_\epsilon ) \text { locally minimizes }v_\epsilon (s,y,z)\\&\quad := v(s,y,z)+\frac{\epsilon }{z}. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Backhoff-Veraguas, J., Tangpi, L. On the dynamic representation of some time-inconsistent risk measures in a Brownian filtration. Math Finan Econ 14, 433–460 (2020). https://doi.org/10.1007/s11579-020-00261-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11579-020-00261-2

Keywords

Mathematics Subject Classification

Navigation