Skip to main content
Log in

Euler–Lagrange equations of stochastic differential games: application to a game of a productive asset

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

This paper analyzes a noncooperative and symmetric dynamic game where players have free access to a productive asset whose evolution is a diffusion process with Brownian uncertainty. A Euler–Lagrange equation is found and used to provide necessary and sufficient conditions for the existence and uniqueness of a smooth Markov Perfect Nash Equilibrium. The Euler–Lagrange equation also provides a stochastic Keynes–Ramsey rule, which has the form of a forward–backward stochastic differential equation. It is used to study the properties of the equilibrium and to make some comparative statics exercises.

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. EL equations can also be formulated in the general case by introducing multipliers and inequalities instead of equalities, although obviously in this case, they lose much of their direct applicability.

  2. Of course, this program succeeds in continuous time under some technical conditions guaranteeing the existence of solutions of the corresponding evolution equations.

  3. EL equations were found following the stochastic maximum principle conditions, which is a more general way than that used here, based on the value function; the second approach has the advantage of simplicity. In Josa-Fombellida and Rincón-Zapatero (2010), an EL equation of a Mayer problem, where the diffusion coefficient depends on the control, has been obtained and analyzed. It turns out that the EL equation in this case is much more complex.

  4. Note that in particular the one-player game never fulfills these assumptions.

  5. Note that this claim has no direct implications for the precautionary savings behavior of players, as the context is different. Precautionary savings means that the agent saves more today when there is uncertainty in his/her tomorrow’s income than when the uncertainty is eliminated by adding the expected income to the wealth process. We do not carry out this exercise here, but compare the Markov equilibrium strategies of two games with different diffusion coefficients, without modifying the income process (in this case, the productive asset process). Usually, precautionary savings appear when the marginal utility is convex, which can be easily proved in two period, discrete time models by means of Jensen’s inequality, see Leland (1968).

  6. The case \(m_i>n\) could also be analyzed, by means of a reduction to the case \(m_i=n\) as in Josa-Fombellida and Rincón-Zapatero (2007).

  7. But in some models, the MPNE is also Pareto optimal; see Martín-Herrán and Rincón-Zapatero (2005).

  8. The game is a particular case of the general framework described in Sect. 2 above. The class of admissible strategies is given in Definition 1.

  9. This relationship between \(\rho \) and \(\pi \) is general and follows from the identity \(R'=-1+RP\) and the definitions of \(\rho \) and \(\pi \).

  10. FBSDEs were introduced by Bismut (1973) for stochastic control problems and studied with more generality in Pardoux and Peng (1990). FBSDEs play a central role in the statement of the stochastic maximum principle, see Peng (1990), Yong and Zhou (1999), or Ma and Yong (1999). FBSDEs in optimization models play the same role as the Hamiltonian system for the state-costate variable in deterministic dynamics, that is, they constitute a part of the necessary conditions for optimality. In our game, the forward part corresponds to the new state \(Y\) and the backward part to the optimal strategy \(C\). The definition of a backward SDE is not straightforward, as the filtration of the Brownian motion is an increasing family of \(\sigma \)-algebras, but the stochastic process \(C\) evolves in the opposite direction of time. The measurability problems are overcome by introducing the process \(Z\), which is an integral part of the definition of the solution.

  11. The deterministic KR rule or “modified Ramsey rule” is obtained setting \(N=1\) and \(\sigma \equiv 0\) in (33), to get

    $$\begin{aligned} \frac{\dot{C}}{C} = \rho (C)(F'(X)-r), \end{aligned}$$

    where the (deterministic) process \(X\) satisfies

    $$\begin{aligned} \dot{X} = F(X)-NC. \end{aligned}$$
  12. The selection of the boundary conditions at \(1/n\) and \(n\) can be done differently, with the only requisite of being compatible with \(u_0(x)\), that is, conserving continuity and smoothness; in the proof below, we will use a different set of boundary conditions still compatible.

  13. Note that what we require here is \(\mu _1(M)\le \frac{\partial a(x,p)}{\partial p}\le \mu _2(M)\) for \(x\) belonging to a compact set of the kind \([1/M,M]\) and not simply to \(x\le M\). This is a slight variation that is not problematic in this framework.

References

  • Amir, R.: Continuous stochastic games of capital accumulation with convex transitions. Games Econ. Behav. 15, 111–131 (1996a)

    Article  Google Scholar 

  • Amir, R.: Strategic intergenerational bequests with stochastic convex production. Econ. Theory 8, 367–376 (1996b)

    Article  Google Scholar 

  • Amir, R.: A new look at optimal growth under uncertainty. J. Econ. Dyn. Control 22, 67–86 (1997)

    Article  Google Scholar 

  • Balbus, Ł., Reffett, K., Wózny, Ł.: A constructive study of Markov equilibria in stochastic games with strategic complementarities. J. Econ. Theory 150, 815–840 (2014)

    Article  Google Scholar 

  • Başar, T., Olsder, G.J.: Dynamic Non-Cooperative Game Theory. Society for Industrial and Applied Mathematics, Philadelphia (1999)

    Google Scholar 

  • Benhabib, J., Radner, R.: The joint exploitation of a productive asset: a game-theoretic approach. Econ. Theory 2, 155–190 (1992)

    Article  Google Scholar 

  • Bismut, J.M.: Conjugate convex functions in optimal stochastic control. J. Math. Anal. Appl. 44, 384–404 (1973)

    Article  Google Scholar 

  • Bourguignon, F.: A particular class of continuous-time stochastic growth models. J. Econ. Theory 9, 141–158 (1974)

    Article  Google Scholar 

  • Carroll, C.D., Kimball, M.S.: On the concavity of the consumption function. Econometrica 64, 981–992 (1996)

    Article  Google Scholar 

  • Case, J.: A dynamic version of Cournot’s problem. In: Roxin, O., Liu, P.T., Sternbergm, R.L. (eds.) Differential Games and Control Theory, Lecture Notes In Pure And Applied Mathematics vol 10, pp. 161–177. Dekker, New York (1974)

  • Chang, F.R.: The inverse optimal problem: a dynamic programming approach. Econometrica 56, 147–172 (1988)

  • Conrad, J.M.: Differential games in a stochastic fishery. Unpublished manuscript (2012)

  • Clemhout, S., Wan Jr, H.Y.: Cartelization conserves endangered species? In: Feichtinger, G. (ed.) Optimal Control Theory and Economic Analysis 2, pp. 549–568. North-Holland, Amsterdam (1985)

    Google Scholar 

  • Danskin, J.M.: The theory of max–min, with applications. SIAM J. Appl. Math. 14, 641–664 (1966)

    Article  Google Scholar 

  • Dockner, E., Jorgensen, S., Van Long, N., Sorger, G.: Differential Games in Economics and Management Science. Cambridge University Press, Cambridge (2000)

    Book  Google Scholar 

  • Dockner, E., Sorger, G.: Existence and properties of equilibria for a dynamic game on productive assets. J. Econ. Theory 71, 209–227 (1996)

    Article  Google Scholar 

  • Dockner, D., Wagener, F.: Markov perfect Nash equilibria in models with a single capital stock. Econ. Theory 56, 585–625 (2014)

    Article  Google Scholar 

  • Dutta, P.K., Sundaram, R.K.: How different can strategic models be? J. Econ. Theory 60, 42–61 (1993a)

    Article  Google Scholar 

  • Dutta, P.K., Sundaram, R.K.: The tragedy of the commons? Econ. Theory 3, 413–426 (1993b)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Fleming, W.H., Soner, H.M.: Controlled Markov Processes and Viscosity Solutions. Springer, New York (2006)

    Google Scholar 

  • Foldes, L.: The optimal consumption function in a Brownian model of accumulation. Part A: the consumption function as solution of a boundary value problem. J. Econ. Dyn. Control 25, 1951–1971 (2001)

    Article  Google Scholar 

  • Ikeda, N., Watanabe, S.: A comparison theorem for solutions of stochastic differential equations and its applications. Osaka J. Math. 14, 619–633 (1977)

    Google Scholar 

  • Josa-Fombellida, R., Rincón-Zapatero, J.P.: New approach to stochastic optimal control. J. Optim. Theory Appl. 135, 163–177 (2007)

    Article  Google Scholar 

  • Josa-Fombellida, R., Rincón-Zapatero, J.P.: On a PDE arising in one-dimensional stochastic control problems. J. Optim. Theory Appl. 146, 1–26 (2010)

    Article  Google Scholar 

  • Kamihigashi, T.: Almost sure convergence to zero in stochastic growth models. Econ. Theory 29, 231–237 (2006)

    Article  Google Scholar 

  • Kimball, M.S.: Precautionary saving in the small and in the large. Econometrica 58, 53–73 (1990)

    Article  Google Scholar 

  • Ladyzhenskaya, O.A., Solonnikov, V.A., Ural’ceva, N.N.: Linear and Quasi-Linear Equations of Parabolic Type. Translations of Mathematical Monographs, vol. 23. American Mathematical Society, Providence (1968)

    Google Scholar 

  • Lancaster, K.: The dynamic inefficiency of capitalism. J. Polit. Econ. 87, 1092–1109 (1973)

    Article  Google Scholar 

  • Leland, H.E.: Saving and uncertainty: the precautionary demand for saving. Q. J. Econ. 82, 465–473 (1968)

    Article  Google Scholar 

  • Levhari, D., Mirman, L.J.: The great fish war: an example using a dynamic Cournot–Nash solution. Bell J. Econ. 1, 322–334 (1980)

    Article  Google Scholar 

  • Ma, J., Yong, J.: Forward–Backward Stochastic Differential Equations and Their Applications. Lecture Notes in Mathematics 1702. Springer, Berlin (1999)

    Google Scholar 

  • Mao, X.: Stochastic Differential Equations and Their Applications. Horwood Publishing, Chichester (1997)

    Google Scholar 

  • Martín-Herrán, G., Rincón-Zapatero, J.P.: Efficient Markov perfect Nash equilibria: theory and application to dynamic fishery games. J. Econ. Dyn. Control 29, 1073–1096 (2005)

    Article  Google Scholar 

  • Melhmann, A.: Applied Differential Games. Plenum Press, New York (1988)

    Book  Google Scholar 

  • Merton, R.C.: An assymptotic theory of growth under uncertainty. Rev. Econ. Stud. 42, 375–393 (1975)

    Article  Google Scholar 

  • Mitra, T., Roy, S.: Optimal exploitation of renewable resources under uncertainty and the extinction of species. Econ. Theory 28, 1–23 (2006)

    Article  Google Scholar 

  • Mitra, T., Sorger, G.: Extinction in common property resource models: an analytically tractable example. Econ. Theory 57, 41–57 (2014)

    Article  Google Scholar 

  • Nowak, A.: On perfect equilibria in stochastic models of growth with intergenerational altruism. Econ. Theory 28, 73–83 (2006)

    Article  Google Scholar 

  • Pardoux, E., Peng, S.: Adapted solution of a backward stochastic differential equation. Syst. Control Lett. 14, 55–61 (1990)

    Article  Google Scholar 

  • Peng, S.: A general stochastic maximum principle for optimal control problems. SIAM J. Control Optim. 28, 966–979 (1990)

    Article  Google Scholar 

  • Polyanin, A.D., Zaitsev, V.F.: Handbook of Exact Solutions for Ordinary Differential Equations. CRC Press, Boca Raton (1995)

    Google Scholar 

  • Ramsey, F.P.: Mathematical theory of saving. Econ. J. 38, 543–559 (1928)

    Article  Google Scholar 

  • Reinganum, J.F., Stokey, N.L.: Oligopoly extraction of a common property natural resource: the importance of the period of commitment in dynamic games. Int. Econ. Rev. 26, 161–173 (1985)

    Article  Google Scholar 

  • Rincón-Zapatero, J.P.: Characterization of Markovian equilibria in a class of differential games. J. Econ. Dyn. Control 28, 1243–1266 (2004)

    Article  Google Scholar 

  • Rincón-Zapatero, J.P., Martínez, J., Martín-Herrán, G.: New method to characterize subgame perfect Nash equilibria in differential games. J. Optim. Theory Appl. 96, 377–395 (1998)

    Article  Google Scholar 

  • Shimomura, K.: The feedback equilibria of a differential game of capitalism. J. Econ. Dyn. Control 15, 317–338 (1991)

    Article  Google Scholar 

  • Sorger, G.: Markov-perfect Nash equilibria in a class of resource games. Econ. Theory 11, 79–100 (1998)

    Article  Google Scholar 

  • Sundaram, R.K.: Perfect equilibrium in non-randomized strategies in a class of symmetric dynamic games. J. Econ. Theory 47, 153–177 (1989)

    Article  Google Scholar 

  • Tsutsui, S., Mino, K.: Nonlinear strategies in dynamic duopolistic competition with sticky prices. J. Econ. Theory 52, 136–161 (1990)

    Article  Google Scholar 

  • Van Long, N.: Dynamic games in the economics of natural resources: a survey. Dyn. Games Appl. 1, 115–148 (2011)

    Article  Google Scholar 

  • Yong, J., Zhou, X.Y.: Stochastic Controls. Hamiltonian Systems and HJB Equations. Springer, New York (1999)

    Google Scholar 

  • Zhiyuan, H.: A comparison theorem for solutions of stochastic differential equations and its applications. Proc. Am. Math. Soc. 91, 611–617 (1984)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Juan Pablo Rincón-Zapatero.

Additional information

We gratefully acknowledge the constructive comments of the anonymous referees and the editor. Part of this work was done while Juan Pablo Rincón-Zapatero was visiting the Department of Economics at Indiana University with the support of a Salvador de Madariaga Scholarship. The authors were also supported by the Spanish Ministerio de Ciencia e Innovación under Projects ECO2008-02358 and ECO2011-24200, and first author from Junta de Castilla y León under Project VA056A09.

Appendix: Proofs

Appendix: Proofs

In this appendix, we collect the proofs of the results that are not shown in the body text. It also include some auxiliary results that are used in the proofs.

1.1 Proof of Theorem 2

Let \(R_T=[0,T]\times (0,\infty )\). We first state an auxiliary result.

Existence of solutions to quasilinear parabolic PDEs.

Theorem 13

There exists at least one bounded solution in \(R_T\) of the Cauchy problem

$$\begin{aligned} u_{\tau }-\frac{\partial }{\partial x} a(\tau ,x,u,u_x) = b(\tau ,x,u,u_x), \end{aligned}$$

with initial condition

$$\begin{aligned} u(0,x)=u_0(x),\quad x>0. \end{aligned}$$

if all of the following conditions are satisfied.

C1.:

\(u_0\) is of class \(C^4\) and bounded.

C2.:

Functions \(a\) and \(b\) are of class \(C^3\) and \(C^2\), respectively.

C3.:

There are non-negative constants \(b_1\) and \(b_2\) such that for all \(x\) and \(u\)

$$\begin{aligned} \left( b(\tau ,x,u,0)+\frac{\partial }{\partial x} a(\tau ,x,u,0)\right) u\le b_1 u^2+b_2. \end{aligned}$$
C4.:

For all \(M>0\), there are constants \(\mu _2(M)\ge \mu _1(M)>0\) such that, if \(\tau , x\) and \(u\) are bounded in modulus by \(M\), then for arbitrary \(p\)

$$\begin{aligned} \mu _1(M)\le \frac{\partial }{\partial p} a(\tau ,x,u,p) \le \mu _2(M) \end{aligned}$$

and

$$\begin{aligned} \left( |a|+\left| \frac{\partial a}{\partial u}\right| \right) (1+|p|)+\left| \frac{\partial a}{\partial x}\right| +|b|\le \mu _2(M)(1+|p|)^2. \end{aligned}$$

The problem admits no more than a classical solution in \(R_T\) that is bounded together with its derivatives of first and second orders if the following additional conditions hold.

C5.:

For all \(M>0\) there are non-negative constants \(\nu _1(M)\) and \(\nu _2(M)\) such that

$$\begin{aligned}&\max _{\begin{array}{c} (t,x)\in R_T\\ |u|, |p|\le M \end{array}}\left| \frac{\partial ^2 a}{\partial p\partial u}, \frac{\partial ^2 a}{\partial p^2}, \frac{\partial A}{\partial p}\right| \le \nu _1(M),\\&\min _{\begin{array}{c} (t,x)\in R_T\\ |u|, |p|\le M \end{array}} \frac{\partial A}{\partial u}\ge -\nu _2(M), \end{aligned}$$

where

$$\begin{aligned} A=a-\frac{\partial a}{\partial u} p - \frac{\partial a}{\partial x}. \end{aligned}$$

Proof

The result is a consequence of Theorem 8.1 of Ladyzhenskaya et al (1969), the only difference being that we have set the problem in \([0,T]\times (0,\infty )\) instead of \([0,T]\times {\mathbb {R}}\) and that we require more smoothness. The method of proof of Theorem 8.1 in Ladyzhenskaya et al consists in considering truncated problems on the strip \([0,T]\times [1/n,n]\) with boundary conditions \(u_n(0,x)=u_0(x)\) for all \(x\in [1/n,n]\) andFootnote 12 \(u_n(t,1/n)=u_0(1/n), u_n(t,n)=u_0(n)\) for all \(t\ge 0\). These solutions converge smoothly to a solution \(u\) of the original equation as \(n\rightarrow \infty \). \(\square \)

Motivated by the necessity to drop the boundedness of the data defining the game, we will consider the EL Eq. (10) for the function \(u=\phi /f\) written in the divergence form

$$\begin{aligned} u_{\tau } - \frac{{\partial }}{{\partial }x} a(x,u_x) = b(x, u, u_x), \end{aligned}$$
(39)

where \(\tau =T-t\) and

$$\begin{aligned} a(x,p)&=\frac{\sigma (x)^2}{2} p,\\ b(x, u, p)&= \Big ( F -Nu f +(N-1) {R}(u f )+\frac{f' }{f } \sigma ^2 \Big )p\\&\quad + \Big ( F -Nu f +(N-1) {R}(u f )+ \sigma ' \sigma \Big )u \frac{f' }{f }\\&\quad +\frac{1}{2} \sigma ^2 u \frac{f'' }{f } -\frac{\sigma ^2}{2} \frac{ {P}(u f )}{f} (pf+u f')^2+\frac{r-F'}{f} {R(u f)}. \end{aligned}$$

The initial condition is

$$\begin{aligned} u(0,x)=\varphi _0 (x)=\frac{\varphi (x)}{f(x)}. \end{aligned}$$
(40)

Note that even if the initial condition \(\varphi _0\) is now bounded by assumption, still we cannot apply the above theorem directly. The difficulties are two: (i) the function \(\sigma \) vanishes at \(x=0\), thus it is not uniformly bounded away from zero; and (ii) the function \(P\) is in general not defined at \(0\) and in fact \(\lim _{c\rightarrow 0^+} P(c) = \infty \) for problems with CRRA utility, where \(P(c)=(1+\delta )/c\). To deal with (i) we consider the PDE (39) on bounded subintervals \(I_n=[1/n,n], n=1,2, \ldots \), and then we take a limit as \(n\rightarrow \infty \), while for (ii) we will prove that the solutions \(u_n\) found in the subintervals above remain uniformly bounded away from 0, in the sense that there exists a lower bound \(l_m>0\) such that \(u_n\ge l_m\) for all \(n\ge m\) for all \(x\in I_m\). As a byproduct of the proof, we obtain the estimates claimed in the theorem.

  • C.1. is fulfilled, since \(\varphi _0(x)=\varphi (x)/f(x)\) is bounded and smooth on \((0,\infty )\), by assumption.

  • C.2. holds, as the function \(a\) has the required smoothness. As explained above, we will prove below that \(u\) never vanishes on \((0,\infty )\), thus the term \(P(uf)\) does not pose any problem at all for the smoothness of function \(b\).

  • C.3. There are constants \(b_1\) and \(b_2\) such that

    $$\begin{aligned} b(x,u,0) u&= \Big (F - Nu f +(N-1) {R}(u f )+ \sigma ' \sigma \Big )u^2 \frac{f' }{f } \\&\quad +\frac{1}{2} \sigma ^2 u^2 \frac{f'' }{f } -\frac{\sigma ^2}{2} \frac{ {P}(u f )}{f} (u f')^2 u+\frac{r-F'}{f} {R(u f)}u \le b_1u^2+b_2 \end{aligned}$$

    (since \(a(x,0)=0\)). To see this, note that, thanks to our assumptions,

    $$\begin{aligned} b(x,u,0)u&\le \alpha ^+u^3f' + \left( (F +\sigma ' \sigma )\frac{f'}{f} +\frac{1}{2} \sigma ^2\frac{f''}{f} - \frac{\pi ^{-}}{2}\sigma ^2\left( \frac{f'}{f}\right) ^2 + \gamma ^+\right) u^2\\&< \alpha ^+(\min _{x\in (0,\infty )}f') u^3 + \beta ^+u^2\le \beta ^+u^2, \end{aligned}$$

    since \(\alpha ^+< 0\) and the solution \(u>0\) (this will be proved below).

  • C.4. first part, is also fulfilled, as

    $$\begin{aligned} \frac{\partial a(x,p)}{\partial p} = \frac{\sigma ^2(x)}{2} \end{aligned}$$

    is positive for \(x>0\) and continuous, thus it is bounded away from 0, as well as bounded above in any compact subset \([1/M,M]\) of \((0,\infty )\).Footnote 13 The second part of C.4 is a local assumption, that also holds because function \(|b|\) is quadratic in \(p\), with continuous coefficients, and thus bounded on compact subsets of \((0,\infty )\); the same is true for

    $$\begin{aligned} \left( |a|+\left| \frac{\partial a}{\partial u}\right| \right) (1+|p|)+\left| \frac{\partial a}{\partial x}\right| = \frac{\sigma ^2(x)}{2} |p| (1+|p|) + |\sigma (x)\sigma '(x)||p|, \end{aligned}$$

    which is also quadratic in \(p\), and since both \(\sigma \) and \(\sigma '\) are continuous and thus they are bounded in compact subsets of \([0,\infty )\).

  • C.5. Both \(\sigma \) and \(\sigma '\) are continuous, and therefore bounded over any compact interval of the state space, achieving uniqueness of the bounded solution

We now show that we can construct a solution \(u>0\) of the PDE (39) as limit of positive truncated solutions \(u_n\) in \([0,T]\times [1/n,n]\) as \(n\rightarrow \infty \). Then, C.2 holds, as \(P\) is smooth in \((0,\infty )\). In fact we prove more than that, as we obtain upper and lower estimates for the solution. The latter will imply in particular that the solution is positive for \(x>0\).

Let the PDE (39) with initial condition (40) and boundary conditions at the extreme points of the interval \(I_n\) given by

$$\begin{aligned} \begin{aligned} u_n(\tau ,1/n)&= \frac{\beta ^+e^{\tau \beta ^+}\varphi _0(1/n)}{\alpha ^+\varphi _0(1/n) f'(1/n)(1-e^{\tau \beta ^+})+\beta ^+},\\ u_n(\tau ,n)&= \frac{\beta ^+e^{\tau \beta ^+}\varphi _0(n)}{\alpha ^+\varphi _0(n) f'(n)(1-e^{\tau \beta ^+})+\beta ^+}. \end{aligned} \end{aligned}$$
(41)

The reason for this particular selection of the boundary conditions is shown next. Formulating these approximating problems, we have eliminated the degeneration in the truncated equation. It is easy to check that all conditions in Theorem 8.1 of Ladyzhenskaya et al. (1968) are fulfilled in a small neighborhood of \(\tau = 0\), since \(\varphi _0(x)\) is bounded away from zero in \(I_n\) and hence a solution \(u_n\) of the truncated problem exists which does not vanish. From this, our aim is to extend this truncated solution to all \([0,T]\) and after this, to get the solution \(\phi (t,x)\) with initial condition \(\phi (0,x)=\varphi (x)\) as the smooth limit of \(u_n(\tau ,x) f(x) \) as \(n\rightarrow \infty \). As explained above, this is the procedure used in Ladyzhenskaya et al. (1968). We show next that the local solution can be extended in time and space. To this end, we define

$$\begin{aligned} M_n(\tau )=\max _{y\in [1/n,n]}u_n(\tau ,y). \end{aligned}$$

By Danskin’s Theorem, function \(M_n\) is almost everywhere differentiable, and at points of differentiability, the derivative \(\dot{M}_n(\tau )=u_{n,\tau }(\tau ,x_n(\tau ))\), where \(u_n(\tau ,x_n(\tau ))=M_n(\tau )\).

We prove that for any \(\tau \)

$$\begin{aligned} M_n(\tau ) \le \frac{\beta ^+e^{\tau \beta ^+}M_n(0)}{\alpha ^+(\min _{I_n}f') M_n(0)(1-e^{\tau \beta ^+})+\beta ^+}. \end{aligned}$$
(42)

where \(M_n(0)=\sup _{x\in I_n}\varphi _0(x)\). Suppose, by way of contradiction, that \(M_n(\tau _0)\) is greater that the right-hand side of (42) for some \(\tau _0>0\). Let \(\overline{ \tau }_0\) be the inferior of all the \(\tau _0\)s satisfying this property, and hence, by continuity of \(M_n, M_n(\overline{\tau }_0)\) equals the right-hand side of (42). Then \(x_n(\tau )\) is interior to \(I_n\) for every \(\tau \in [\overline{\tau }_0, \tau _0]\), due to the boundary conditions (41). In consequence, \(u_{n,x}(\tau ,x_n(\tau ))=0\) and \(u_{n,xx}(\tau ,x_n(\tau )) \le 0\) for any \(\tau \in [\overline{\tau }_0, \tau _0]\). Hence, \((\partial /\partial x) (\sigma ^2 u_{n,x})|_{(\tau , x_n(\tau ))} \le 0\). This information, used in the Eq. (39) for \(\nu _n\), provides the following chain of inequalities

$$\begin{aligned} \begin{aligned} \dot{M}_n(\tau )&\le b( x_n(\tau ), M_n (\tau ),0)\\&= \Big ( F-NfM_n (\tau ) +(N-1) {R}(fM_n (\tau )) +\sigma ' \sigma \Big )M_n (\tau )\frac{f'}{f}\\&\quad +\frac{1}{2} \sigma ^2 M_n(\tau )\frac{f''}{f} -\frac{1}{2} \sigma ^2 \frac{ {P}(fM_n(\tau ) )}{f} (M_n (\tau )f')^2 + \frac{r-F'}{f} {R}(fM_n (\tau )) \\&\le \alpha ^+M_n^2(\tau )f' + \Big ( (F +\sigma ' \sigma )\frac{f'}{f} \!+\!\frac{1}{2} \sigma ^2\frac{f''}{f} - \frac{\pi ^{-}}{2}\sigma ^2\left( \frac{f'}{f}\right) ^2\!+\!\gamma ^+\Big )M_n (\tau )\\&< \alpha ^+(\min _{I_n}f') M_n^2(\tau ) + \beta ^+M_n(\tau ), \quad \text {a.e. }\tau . \end{aligned} \end{aligned}$$
(43)

We have used the definitions of \(\alpha ^+\) and \(\beta ^+\) done in Assumption (A3) and (A5), respectively, and the fact that \(\alpha ^+< 0\). From this differential inequality of Ricatti we get the estimate

$$\begin{aligned} M_n(\tau ) \le \frac{\beta ^+e^{(\tau -\overline{\tau }_0) \beta ^+}M_n(\overline{\tau }_0)}{\alpha ^+(\min _{I_n}f') M_n(\overline{\tau }_0)(1-e^{(\tau -\overline{\tau }_0) \beta ^+})+\beta ^+}. \end{aligned}$$

Once the expression for \(M_n(\overline{\tau }_0)=\frac{\beta ^+e^{\overline{\tau }_0 \beta ^+}M_n(0)}{\alpha ^+(\min _{I_n}f') M_n(0)(1-e^{\overline{\tau }_0\beta ^+})+\beta ^+}\), which holds by our contradiction argument is substituted into (A.1) inequality (42) easily follows.

Following the same technique, we get the lower estimate. Consider the solution \(v_n\) of the PDE with the boundary conditions

$$\begin{aligned} v_n(\tau ,1/n)&= \frac{\beta ^{-}e^{\tau \beta ^{-}}\varphi _0(1/n)}{\alpha ^-\varphi _0(1/n)f'(1/n)(1-e^{\tau \beta ^{-}})+\beta ^{-}},\\ v_n(\tau ,n)&= \frac{\beta ^{-}e^{\tau \beta ^{-}}\varphi _0(n)}{\alpha ^-\varphi _0(n) f'(n)(1-e^{\tau \beta ^{-}})+\beta ^{-}}. \end{aligned}$$

and

$$\begin{aligned} m_n(\tau )=\min _{y\in [1/n,n]}v_n(\tau ,y), \end{aligned}$$

with the minimum attained at some \(y_n(\tau )\). Similar arguments as done above for the maximum lead to the estimate

$$\begin{aligned} m_n(\tau ) \ge \frac{\beta ^{-}e^{\tau \beta ^{-}}m_n(0)}{\alpha ^-m_n(0) (\max _{I_n} f') (1-e^{\tau \beta ^{-}})+\beta ^{-}}. \end{aligned}$$
(44)

as follows. By way of contradiction one finds

$$\begin{aligned} \dot{m}_n(\tau )\ge g(y_n(\tau ), m_n (\tau ),0)\ge \alpha ^-(\max _{I_n} f') m_n^2(\tau ) +\beta ^{-}m_n(\tau ), \quad \text {a.e. }\tau . \end{aligned}$$

Reasoning much as for the case of \(M_n\), one gets the estimate (44) easily. Thus, we have shown that the local solution is strictly uniformly bounded away from zero in the intervals \(I_n\) and that it is also bounded above. Since this fact is independent of \(\tau \), as well as the upper bound obtained above, the solution can be extended up to the whole \([0,T]\), for any \(T\).

Hence, since \(v_n\rightarrow v\), we have that \(v\) is bounded away from zero, and hence \(v\) is a solution of the Cauchy problem, since all the conditions of the Theorem of Ladyzhenskaya et al. (1968) are fulfilled. Taking limits as \(n\rightarrow \infty \) one has

$$\begin{aligned} \frac{\phi (t,x)}{f(x)} = u(T-t,x) \ge m(T-t) \ge \frac{\beta ^{-}e^{(T-t) \beta ^{-}}m(0)}{\alpha ^-m(0)(\sup _{(0,\infty )} f') (1-e^{(T-t)\beta ^{-}})+\beta ^{-}}. \end{aligned}$$

By the above estimates the limit \(\phi (t,x)= u(T-t,x)f(x)\) is a solution of the Cauchy problem (10) satisfying

$$\begin{aligned} \frac{\phi (t,x)}{f(x)}= v(T-t,x) \le M(T-t) \le \frac{\beta ^+e^{(T-t) \beta ^+}M}{\alpha ^+(\inf _{(0,\infty )}f') M(1-e^{(T-t)\beta ^+})+\beta ^+}, \end{aligned}$$

since \(M_n(0)\rightarrow M\) as \(n\rightarrow \infty \), and \(\inf _{(0,\infty )} f'\le \min _{I_n} f'\).

1.2 Proof of Theorem 7

With CRRA preferences, \(\rho ^{+}=\rho ^{-}=\rho =1/\delta , \alpha ^+=\alpha ^-=\alpha \) and \(\pi ^{+}=\pi ^{-}=b=1+\delta \). See the definition of these constants in Sect. 3, Assumption (A3) and in (12). See also the computations done after Definition 4. We will follow the proof of Theorem 2, using the same notation. Retaking inequality (43) in the aforementioned proof, in the case of CRRA preferences we have

$$\begin{aligned} \begin{aligned} \dot{M}_n(\tau )&\le b( x_n(\tau ), M_n (\tau ),0)\\&= \Big ( F-NfM_n (\tau ) +(N-1) afM_n (\tau ) +\sigma ' \sigma \Big )M_n (\tau )\frac{f'}{f}\\&\quad +\frac{1}{2} \sigma ^2 M_n(\tau )\frac{f''}{f} -\frac{1}{2} \sigma ^2 \frac{b}{f^2M_n(\tau )} (M_n (\tau )f')^2 + \frac{r-F'}{f}M_n (\tau )a\\&\le \alpha f' M_n^2(\tau )\! +\! \Big ( (F +\sigma ' \sigma ) f' \!+\!\frac{1}{2} \sigma ^2 {f''} \!-\!\frac{1}{2} \sigma ^2 \frac{b}{f} f'^2 \!+\! a (r\!-\!F')f\Big ) \frac{M_n (\tau )}{f} \end{aligned} \end{aligned}$$

Since \(f\) is solution of the stationary EL Eq. (10) with CRRA preferences, then

$$\begin{aligned} ( F +\sigma ' \sigma ) {f'} +\frac{1}{2} \sigma ^2 {f''} -\frac{1}{2} \sigma ^2 \frac{b}{f} f'^2 + (r-F')af= -\alpha f f'. \end{aligned}$$

Plugging this into the inequality above it simplifies to

$$\begin{aligned} \dot{M}_n(\tau ) \le \alpha f' M_n^2(\tau )-\alpha f' M_n(\tau )\le \alpha (\inf _{I_n} f') M_n^2(\tau ) - \alpha (\sup _{I_n} f') M_n(\tau ). \end{aligned}$$

Hence

$$\begin{aligned} M_n(\tau ) \le \frac{M_n(0) e^{-\tau \alpha (\sup _{I_n}f')}}{- M_n(0) (1-e^{-\tau \alpha (\sup _{I_n}f')})+1}. \end{aligned}$$

A similar computation for the minimum shows

$$\begin{aligned} m_n(\tau ) \ge \frac{m_n(0) e^{-\tau \alpha (\inf _{I_n}f')}}{- m_n(0) (1-e^{-\tau \alpha (\inf _{I_n}f')})+1}. \end{aligned}$$

Notice that in the former case \(f'\) is evaluated at the point where \(u_n\) attains a maximum in \(I_n=[1/n,n]\), say \(x_n\), whereas in the latter case it is at the point where \(u_n\) attains a minimum, say \(y_n\). Thus,

$$\begin{aligned} \frac{m_n(0) e^{-\tau \alpha (\inf _{I_n}f')}}{- m_n(0) (1-e^{-\tau \alpha (\inf _{I_n}f')})+1}\le u_n(t,x)\le \frac{M_n(0) e^{-\tau \alpha (\sup _{I_n}f') }}{- M_n(0) (1-e^{-\tau \alpha (\sup _{I_n}f')})+1}. \end{aligned}$$

Taking the limit as \(n\) tends to \(\infty \) and since \(u_n(t,x)\rightarrow \phi (t,x)/f(x)\) as \(n\rightarrow \infty \), we find

$$\begin{aligned} \frac{m e^{-\tau \alpha (\inf _{I_n} f')}}{- m (1-e^{-\tau \alpha (\inf _{I_n} f')})+1}\le \frac{\phi (\tau ,x)}{f(x)}\le \frac{M e^{-\tau \alpha (\sup _{I_n} f')}}{- M (1-e^{-\tau \alpha (\sup _{I_n} f')})+1}. \end{aligned}$$

As \(T \rightarrow \infty , \tau \rightarrow \infty \) and one finally find that \(\phi (t,x)\) converges to \(f(x)\).

1.3 Proof of Theorem 8

We follow the same scheme of proof as in Theorems 2 and 7. Notice that the hypotheses of the theorem imply \(\varphi '(x) = S''(x)/L''(S'(x)) \ge 0\). Now, derive the EL Eq. (10) with respect to \(x\) and let \(v=\phi _x\). Then \(w\) solves the Cauchy problem

$$\begin{aligned} v_{\tau } -\frac{1}{2} \frac{{\partial }}{{\partial }x} (\sigma (x)^2 v_x)= & {} g_x(x,\phi ,v) + v g_c(x,\phi ,v) + v_x g_v(x,\phi ,v)\\&\quad + v \sigma '(x)^2 + v\sigma (x) \sigma ''(x) + v_x \sigma (x)\sigma '(x). \\ v(0,x)= & {} \varphi '(x)\ge 0. \end{aligned}$$

Function \(g\) is

$$\begin{aligned} g(x,c,v) = (F(x) - N c + (N-1) {R}(c))v - \frac{1}{2} \sigma (x)^2 {P} (c) v^2 + (r - F'(x)) {R}(c). \end{aligned}$$

We will follow a similar strategy of proof as in Theorem 2, considering truncated intervals for the variable \(x, I_n=[1/n,n]\) and the solution \(v_n(\tau ,x)\) in \([0,T]\times I_n\) satisfying the boundary conditions

$$\begin{aligned} v_n(\tau ,1/n) = \varphi '(1/n)>0 ,\quad v_n(\tau ,n) = \varphi '(n)>0. \end{aligned}$$

Let \(\nu _n(\tau ) = \min _{x\in I_n} v_n(\tau ,x)\). A reasoning by contradiction, assuming the existence of \(\tau _0\) satisfying \(\nu _n(\tau _0) < \nu _n(0)\), will lead to a contradiction as follows. Let \(\overline{ \tau }_0\) be the inferior of all the \(\tau \)s satisfying this property; by continuity of \(\nu _n, \nu _n(\overline{\tau }_0)=\nu _n(0)\). Then, we obtain the inequality

$$\begin{aligned} \dot{\nu }_n \ge g_x(x,\phi ,\nu _n)+\nu _n \big (g_c(x,\phi _n,\nu _n)+\sigma '(x)^2+\sigma (x)\sigma ''(x)\big ). \end{aligned}$$

We have used the same arguments as those used in the proof of Theorem 2, hence we do not repeat it here. The notation \(\phi _n\) is used for the solution of the EL Eq. (10) in \([0,T]\times I_n\). The term \(\phi _n(\tau ,\zeta (\tau ))\), where \(\zeta (\tau )\) minimizes \(v_n\) over \(I_n\), does not pose any problem at all. Given that \( g_x(x,c,v) = v\big (F'(x)-v\sigma (x)\sigma '(x) {P}(c)\big )-F''(x) {R}(c), \) we have

$$\begin{aligned} \dot{\nu }_n&\ge - F''(x) {R}(\phi _n)\\&\quad + \nu _n \big (F'(x)-\nu _n\sigma (x)\sigma '(x) {P}(\phi _n) + g_c(x,\phi _n,\nu _n)+\sigma '(x)^2+\sigma (x)\sigma ''(x)\big ), \end{aligned}$$

with \(\nu _n(\overline{\tau })=0\). Since \(F''\) is concave and \( {R}\) is non-negative, \(-F''(x) {R}(c)\ge 0\), thus

$$\begin{aligned} \nu _n(\tau ) \ge \nu _n (\overline{\tau }_0) e^{\int _{\overline{\tau }_0}^{\tau } \{\cdots \} \, d\tau } = 0\quad \forall \tau \in [\overline{\tau }_0,\tau _0], \end{aligned}$$

in contradiction with \(\nu _n(\tau _0) < 0\). Hence \(0\le \nu _n(\tau )\le v_n(\tau ,x)\) for all \( \tau , x\in I_n\), and then the limit function \(\phi _x(\tau ,x)\ge 0\).

1.4 Proof of Theorem 9

We first establish a lemma about the concavity of the consumption rate at the final time \(T\). It is established for general \(L\), not only in the class CRRA. In the lemma, \(\rho _{\{\}}\) stands for the elasticity of the marginal utility and \(\pi _{\{\}}\) for the relative prudence index of a given utility function. For the linear game, the lemma implies that \(\varphi \) is concave if and only if the bequest function \(S\) satisfies

$$\begin{aligned} \frac{S'(x) S'''(x)}{{S''}^2(x)} \ge 1+\frac{1}{\delta }. \end{aligned}$$

Lemma 3

\(\varphi ''\le 0 \) if and only if for all \(x> 0\)

$$\begin{aligned} \rho _S(x)\pi _S(x) \ge \rho _L(\varphi (x))\pi _L(\varphi (x)). \end{aligned}$$

Proof

Deriving twice in \(L'(\varphi (x)) = S'(x)\) we get

$$\begin{aligned}&L''(\varphi (x))\varphi '(x) = S''(x),\\&L'''(\varphi (x))\varphi '(x)^2 + L''(\varphi (x))\varphi ''(x) = S'''(x). \end{aligned}$$

Solving for \(\varphi ''(x)\) and imposing \(\varphi ''(x)\le 0\) we obtain the inequality (we eliminate arguments)

$$\begin{aligned} \frac{S'''}{S''^2} \ge \frac{L'''}{L''^2} \end{aligned}$$

or equivalently, multiplying both sides of the inequality by \(S'>0\)

$$\begin{aligned} \left( \frac{-xS'''}{S''}\right) \left( \frac{-S'}{xS''}\right) \ge \left( \frac{-\varphi L'''}{L''} \right) \left( \frac{-S'}{\varphi L''}\right) . \end{aligned}$$

Noting that \(S'=L'(\varphi )\), and plugging this equality into the right-hand side of the inequality above, we obtain the claim of the lemma. \(\square \)

We now proceed with the proof of Theorem 9 with the same techniques as those used in the proofs of Theorems 2 and 8, deriving twice in the EL Eq. (10) to find a PDE for \(\phi _{xx}\). We refer the reader to the proofs of those theorems for filling in the missing details. Deriving twice in (10) we get

$$\begin{aligned} (\phi _{xx})_{\tau }-\frac{\partial }{\partial x} \left( \frac{\sigma ^2 x^2}{2} \phi _{xxx}\right) = j(x,\phi ,\phi _x,\phi _{xx},\phi _{xxx}) \end{aligned}$$

where the function \(j\) is defined as

$$\begin{aligned} j(x,c,v,w,z)&=w\left( 3\sigma ^2+2F'(x)+3\alpha v +\frac{r-F'(x) }{\delta } - 4\sigma ^2 x \frac{1+\delta }{c} v \right. \\&\quad \left. +\frac{5}{2} \sigma ^2x^2\frac{1+\delta }{c^2} v^2-\sigma ^2x^2\frac{1+\delta }{c}w\right) \\&\quad - \sigma ^2 v^2 \frac{1+\delta }{c} \left( 1-\frac{xv}{c}\right) ^2\\&\quad -F'''(x)\frac{c}{\delta }+F''(x)\left( 1-\frac{2}{\delta }\right) v\\&\quad + \left( 2\sigma ^2 x + F(x) +g_N(c) - \sigma ^2x^2 \frac{1+\delta }{c} v\right) z. \end{aligned}$$

Recall that in the CRRA case, \(\alpha =-N+(N-1)/\delta \). Following the same method of proof as in the above referenced theorems, and defining \( w_n \) as the solution of the PDE above in the interval \(I_n\), we get

$$\begin{aligned} w_{n,\tau } - \frac{\partial }{\partial x} \left( \frac{\sigma ^2 x^2}{2} w_{n,x}\right) = j(x,\phi _n,\phi _{n,x},w_{n},w_{n,x}). \end{aligned}$$
(45)

Here, \(\phi _n\) denotes the restriction of \(\phi \) to \([0,T]\times I_n\). Let \(\omega _n(\tau ) = \max _{I_n} w_n (\tau ,x)\). Reasoning by contradiction supposing that \(\omega _n(\tau ) >0\) at some \(\overline{\tau }>0\), one has that \(w_{n,x}(\overline{\tau })= 0\) and \(\frac{\partial }{\partial x} \left( \frac{\sigma ^2 x^2}{2} w_{n,x}\right) \le 0\), thus plugging this into Eq. (45) we get

$$\begin{aligned} \dot{\omega }_n \le \omega _n\left\{ \cdots \right\} - \sigma ^2 \phi _{n,x}^2 \frac{1+\delta }{\phi _n} \left( 1-\frac{x\phi _{n,x}}{\phi _n}\right) ^2 -F'''(x)\frac{\phi _n}{\delta }+F''(x)\left( 1-\frac{2}{\delta }\right) \phi _{n,x}. \end{aligned}$$

In the linear game \(F''=F'''=0\), hence the second summand in the above expression is non-positive. Given that \(\omega _n(0)=\sup _{[0,\infty )} \varphi ''(x) \le 0\) is also non-positive, we arrive to a contradiction, because it is never possible to have \(\omega _n(\overline{\tau }) >0\) from the above estimate for \(\dot{\omega }_n\).

In the following proofs, recall that \(g_N(c)=-Nc+(N-1)R(c)\).

1.5 Proof of Theorem 10

Let us show that if \(\sigma _i\in \varSigma , i=1,2\) and \(\sigma _1\le \sigma _2\), then \(Y^{\sigma _1}\le Y^{\sigma _2}\) a.s. Clearly, for \(\sigma \in \varSigma , \mu (Y(s))\) satisfies the Novikov condition and then

$$\begin{aligned} M(s) = \exp {\left\{ \int _t^s \mu (Y(a)) d(a) - \frac{1}{2} \int _t^s \mu ^2(Y(a))da\right\} }, \quad s\in [t,T] \end{aligned}$$

is a \(\mathbf {P}\)-martingale, where \(\mathbf {P}\) is the objective probability measure. Define now the probability measure \(\mathbf {\widetilde{P}}\) by

$$\begin{aligned} \frac{d \mathbf {\widetilde{P}}}{d \mathbf {P}} = M(T). \end{aligned}$$

It is known that \(\mathbf {\widetilde{P}}\) is absolutely continuous with respect to \(\mathbf {P}\). By Girsanov’s Theorem, \(\widetilde{\omega }(s) = \omega (s) - \int _t^s \mu (Y(a)) d(a)\) is a \(\mathbf {\widetilde{P}}\)-Brownian motion and, in the new measure, \(Y\) satisfies

$$\begin{aligned} dY (s) = \frac{1}{2}\sigma (Y(s))\sigma '(Y(s)) ds + \sigma (Y(s)) dw (s). \end{aligned}$$

Now, given that \(0<\sigma _1\le \sigma _2\), it holds that \(\int _x ^y \frac{dz}{\sigma _1(z)} \ge \int _x ^y \frac{dz}{\sigma _2(z)}\) for all \(y\ge x\), which is the sufficient condition of Example 2 of Zhiyuan (1984) assuring \(Y^{\sigma _1}\le Y^{\sigma _2}\) for every \(s\in [t,T], \mathbf {\widetilde{P}}\)-a.s., whence \(\mathbf {P}\)-a.s. Then \(\varphi (Y^{\sigma _1}(T))\le \varphi (Y^{\sigma _2}(T))\,\mathbf {P}\)-a.s., since \(\varphi \) is non-decreasing by hypothesis.

Now, let us define \(f^i(\omega , s, C,Z)\) as the drift term in the backward SDE for \(C\) in (36) when the forward process is \(Y^{\sigma _i}(s), i=1,2\), that is

$$\begin{aligned} f^i(\omega , s, C,Z) = C \left( \rho (C) (F'(Y^{\sigma _i})-r) + \frac{1}{2} \pi (C) Z^2 - \frac{g_N(C)Z}{\sigma (Y^{\sigma _i})} \right) , \end{aligned}$$

as well as \(\varphi ^i(T) = \varphi (Y^{\sigma _i}(T))\). In the above, \(\omega \) denotes an element of the sample space \(\varOmega \) where the probability \(\mathbf {P}\) is defined, in order to stress the dependence of the drift term with respect to the stochastic process \(Y^{\sigma _i}\). Let also \((C^1,Z^ 1), (C^2,Z^ 2)\) be solutions of the BSDE (36) when \(\sigma = \sigma _i, i=1,2\). We can write (36) in integral form

$$\begin{aligned} C^i(s) = \varphi ^i - \int _s^T f^i(\omega ,v, C(v),Z(v))dv - \int _s^T Z^ i(v) dw(v). \end{aligned}$$

Let us check that \(-f^1(\omega , s, C^ 2,Z^ 2)\le -f^2 (\omega , s, C^ 1,Z^ 1)\). Note that there are two terms that depend on \(i\) in the definition of \(f^i\): one is \(- \frac{g_N(C)Z}{\sigma (Y^{\sigma _i})} \); we have already proved that \({\sigma (Y^{\sigma _1})}\le {\sigma (Y^{\sigma _2})}\). Since \(\sigma \) is increasing, that \(Z\ge 0\) for all \(s \, \mathbf {P}\)-a.s., and that \(g_N(C)\le 0\) (see Sect. 5), then \( - \frac{g_N(C)}{\sigma (Y^{\sigma _2})} \le - \frac{g_N(C)}{\sigma (Y^{\sigma _1})}\). The other term is \(F'(Y^ {\sigma _i})\); since \(F\) is concave, \(F'\) is non-increasing, thus \(-F'(Y^{\sigma _1}) \le -F'(Y^{\sigma _2})\). Hence, \(-f^1(\omega , s, C^ 2,Z^ 2)\le -f^2 (\omega , s, C^ 1,Z^ 1)\), as claimed. Now, the Comparison Theorem 2.2 in Karoui et al. (1997) ensures that \(C^1(s) \le C^2(s)\) for all \(s, \mathbf {P}\)-a.s. Since the process \(C^ i\) is deterministic at \((t,x)\) and \(C^ i(t)=\phi ^ {\sigma ^ i}(t,x)\), we have that \(\phi ^{\sigma _1}(t,x)\le \phi ^{\sigma _2}(t,x)\) and the proof is finished.

1.6 Proof of Theorem 11

To show the result, we will use a representation of the MPNE by means of an FBSDE, alternative to the KR rule introduced in Sect. 5, more amenable for our purposes. Let the process \(Y\) that satisfies

$$\begin{aligned} dY =(F(Y) + (\sigma {\sigma }')(Y))\, ds+\sigma (Y)\, d\omega ,\quad Y(t)=x, \end{aligned}$$

where we have omitted the argument \(s\) to shorten notation. Then, for \(C(s)=\phi (s,Y(s))\) and the square-integrable adapted process (\(Z=\sigma (Y(s))\phi _x(s,Y(s))\)), where \(\phi \) is a solution of the EL Eq. (10) we have, as in Sect. 6.4

$$\begin{aligned} dC (s)= & {} \left( R(C(s)) (F'(Y(s))-r) - \frac{1}{2} {P}({C}(s)) Z^2(s) + \frac{g_N(C(s))}{\sigma (Y(s))}Z(s) \right) ds \\&+ Z(s) dw (s), \end{aligned}$$

and \(C(T) = \varphi (Y(T))\). Note that the process \(Z\) does not depend on the number of players, \(N\). Let

$$\begin{aligned} f^ N (\omega ,s,C,Z) = R(C) (F'(Y(s))-r) - \frac{1}{2} {P}({C}) Z^2 + \frac{g_N(C)}{\sigma (Y(s))}Z \end{aligned}$$

be the drift term of the SDE of \(C\) when the number of players is \(N\) and let \((C^ {N},Z^ N)\) be the corresponding solution. We have included \(\omega \in \varOmega \) into the notation to stress the dependence with respect to the process \(Y\). We know that \(g_N\) is negative thanks to Assumption (A3) (a) (see also (12)). On the other hand, it is easy to see that \(g_N\), as a function of \(N\), is monotonous increasing if \(\rho (c)>1\) and decreasing if \(\rho (c)<1\). Let us suppose first that \(\rho (c)>1\). Given that \(Z\ge 0\) a.s. by Theorem 5, then \(\frac{g_{N_1}(C)}{\sigma }Z \le \frac{g_{N_2}(C)}{\sigma }Z \) a.s. if \(N_1\le N_2\). Hence \(-f^ {N_1} (\omega ,s,C^ {N_2},Z^ {N_2}) \le -f^{N_2}(\omega , s, C^ {N_2},\xi ^ {N_2})\), as well as \(C^{N_1}(T) = C^{N_2}(T)\). According to the Comparison Theorem 2.2 of Karoui et al. (1997), \(C^{N_1}(s)\le C^{N_2}(s)\) for all \(t\le s\le T\), a.s. The case \(\rho (c) <1\) is analyzed analogously. Since \(\phi ^ {N_i}(t,x)=C^{N_i}(t)\) is deterministic, the theorem is proved.

1.7 Proof of Theorem 12

We follow the same steps as in the proof of Theorem 11, using the same FBSDE representation. The process \(Y\) is independent of the preference rate and the drift term of the SDE of \(C\) in (A.6), \(f^ {r}\), is decreasing in the preference rate. Hence \(r_1\le r_2\) implies \(-f^{r_1}\le -f^{r_2}\) and, by the Comparison Theorem 2.2 of Karoui et al. (1997), we have the result.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Josa-Fombellida, R., Rincón-Zapatero, J.P. Euler–Lagrange equations of stochastic differential games: application to a game of a productive asset. Econ Theory 59, 61–108 (2015). https://doi.org/10.1007/s00199-015-0873-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-015-0873-z

Keywords

JEL Classification

Navigation