Skip to main content
Log in

Continuous-Time Public Good Contribution Under Uncertainty: A Stochastic Control Approach

  • Published:
Applied Mathematics & Optimization Submit manuscript

Abstract

In this paper we study continuous-time stochastic control problems with both monotone and classical controls motivated by the so-called public good contribution problem. That is the problem of n economic agents aiming to maximize their expected utility allocating initial wealth over a given time period between private consumption and irreversible contributions to increase the level of some public good. We investigate the corresponding social planner problem and the case of strategic interaction between the agents, i.e. the public good contribution game. We show existence and uniqueness of the social planner’s optimal policy, we characterize it by necessary and sufficient stochastic Kuhn–Tucker conditions and we provide its expression in terms of the unique optional solution of a stochastic backward equation. Similar stochastic first order conditions prove to be very useful for studying any Nash equilibria of the public good contribution game. In the symmetric case they allow us to prove (qualitative) uniqueness of the Nash equilibrium, which we again construct as the unique optional solution of a stochastic backward equation. We finally also provide a detailed analysis of the so-called free rider effect.

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. See, e.g., the classical work of [33] and the recent papers by [10], [34] and [35] and references therein.

  2. A quasilinear utility is an additively separable utility function which is linear in one of its arguments and concave in the others.

  3. A stochastic process X is:

    1. (a)

      optional if it is measurable with respect to the optional sigma-field \(\mathcal {O}\) on \(\Omega \times [0,T]\) generated, e.g., by the right-continuous adapted processes;

    2. (b)

      lower-semicontinuous in expectation if for any stopping time \(\tau \) one has \(\liminf _{n\uparrow \infty }E[X(\tau _n)] \ge E[X(\tau )]\), whenever \(\{\tau _n\}_{n\in \mathbb {N}}\) is a monotone sequence of stopping times converging to \(\tau \);

    3. (c)

      of class (D) if \(\{X(\tau ),\,\,\tau \,\,\text {a stopping time}\}\) defines a uniformly integrable family of random variables on \((\Omega ,{\mathscr {F}},P)\).

    We refer the reader to [18], among others, for further details.

  4. An optional random measure on [0, T] is simply a random variable \(\nu \) valued in the space of nonnegative Borel measures on [0, T] (endowed with the topology of weak*-convergence) such that the process \(\nu (\omega ,t):=\nu (\omega , [0,t])\) is adapted. Our admissible public good contribution processes are the cumulative distributions of optional random measures, being adapted, right-continuous, and nonnegative.

  5. That is, \(\tau \) is a time of increase for \(C_*(\omega ,\cdot )\).

  6. Mathematically, the utility function u is supermodular (see [32]). Economically, this means that the private and the public goods are complements. Two goods are complements if the increase of a good’s demand corresponds to a decrease in the price of the other good.

  7. \(\tau \in \mathcal {T}\) is a point of increase of the optional, non-decreasing process \(\hat{C}\) if \(\hat{C}(t)>\hat{C}(\tau -)\) for all \(t\in (\tau ,T]\), \(P\)-a.s.

  8. Notice that the martingale property of Assumption 2 is without loss of generality in this case, one just has to correct r by \(\frac{1}{2}\sigma ^2\), i.e. by the Laplace exponents of \(\sigma \frac{W(t)}{t}\).

  9. Notice that such a result is in line with the well known finding from the economic literature (see, e.g., [23]) that under perfect reversibility the optimal investment criterion is to equate the marginal operating profit with the user cost of capital.

References

  1. Austen-Smith, D.: Individual contributions to public goods. Econ. Lett. 5, 359–361 (1980)

    Article  Google Scholar 

  2. Back, K., Paulsen, D.: Open loop equilibria and perfect competition in option exercise games. Rev. Financ. Stud. 22(11), 4531–4552 (2009)

    Article  Google Scholar 

  3. Balder, E.J.: New sequential compactness results for spaces of scalarly integrable functions. J. Math. Anal. Appl. 151(1), 1–16 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bank, P.: Optimal control under a dynamic fuel constraint. SIAM J. Control Optim. 44(4), 1529–1541 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bank, P., El Karoui, N.: A stochastic representation theorem with applications to optimization and obstacle problems. Ann. Probab. 32, 1030–1067 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bank, P., Föllmer, H.: American options, multi-armed bandits, and optimal consumption plans: a unifying view. In: Carmona, R., Çinlar, E., Ekeland, I., Jouini, E., Scheinkman, J., Touzi, N. (eds.) Paris-Princeton Lectures on Mathematical Finance 2002. Volume 1814 of Lecture Notes in Mathematics, pp. 1–42. Springer, Berlin (2003)

    Chapter  Google Scholar 

  7. Bank, P., Küchler, C.: On Gittin’s index theorem in continuous time. Stoch. Process. Appl. 117, 1357–1371 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bank, P., Riedel, F.: Optimal consumption choice with intertemporal substitution. Ann. Appl. Probab. 11, 750–788 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  9. Bank, P., Riedel, F.: (2003). Optimal dynamic choice of durable and perishable goods. Discussion Paper 28/2003, Bonn Graduate School of Economics

  10. Battaglini, M., Nunnari, S., Palfrey, T.: Dynamic free riding with irreversible investments. Am. Econ. Rev. 104(9), 2858–2871 (2014)

    Article  Google Scholar 

  11. Bertoin, J.: Lévy Processes. Cambridge University Press, Cambridge (1996)

    MATH  Google Scholar 

  12. Billingsley, P.: Probability and Measures. Wiley, New York (1986)

    MATH  Google Scholar 

  13. Borodin, A.N., Salminen, P.: Handbook of Brownian Motion-Facts and Formulae. Birkhäuser Verlag, Basel (2001)

    MATH  Google Scholar 

  14. Chiarolla, M.B., Ferrari, G.: Identifying the free boundary of a stochastic irreversible investment problem via the Bank-El Karoui representation theorem. SIAM J. Control Optim. 52(2), 1048–1070 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Chiarolla, M.B., Ferrari, G., Riedel, F.: Generalized Kuhn–Tucker conditions for \(N\)-firm stochastic irreversible investment under limited resources. SIAM J. Control Optim. 51(5), 3863–3885 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  16. Chiarolla, M.B., Haussmann, U.G.: The free boundary of the monotone follower. SIAM J. Control Optim. 32(3), 690–727 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  17. Cornes, R., Sandler, T.: The Theory of Externalities, Public Goods and Club Goods, 2nd edn. Cambridge University Press, Cambridge (1996)

    Book  MATH  Google Scholar 

  18. Dellacherie, C., Meyer, P.-A.: Probabilities and Potential, Volume 29 of Mathematics Studies. Oxford, Amsterdam (1978)

    Google Scholar 

  19. Eichberger, J., Kelsey, D.: (1999). Free Riders Do Not Like Uncertainty. Economic Series 9912, University of Saarland

  20. El Karoui, N., Karatzas, I.: A new approach to the Skorohod problem, and its applications. Stoch. Stoch. Rep. 34, 57–82 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  21. Ferrari, G.: On an integral equation for the free boundary of stochastic irreversible investment problems. Ann. Appl. Probab. 25(1), 150–176 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  22. Jacod, J.: Calcul Stochastique et Problèmes de Martingales. Lecture Notes in Mathematics, vol. 714. Springer, Berlin (1979)

    Book  MATH  Google Scholar 

  23. Jorgenson, D.W.: Capital theory and investment behavior. Am. Econ. Rev. 53, 247–257 (1963)

    Google Scholar 

  24. Kabanov, Y.: Hedging and liquidation under transaction costs in currency markets. Financ. Stoch. 3, 237–248 (1999)

    Article  MATH  Google Scholar 

  25. Karatzas, I.: The monotone follower problem in stochastic decision theory. Appl. Math. Optim. 7, 175–189 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  26. Karatzas, I., Shreve, S.E.: Connections between optimal stopping and singular stochastic control I. Monotone follower problems. SIAM J. Control Optim. 22, 856–877 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  27. Komlós, J.: A generalization of the problem of Steinhaus. Acta Math. Acad. Sci. Hung. 18, 217–229 (1967)

    Article  MathSciNet  MATH  Google Scholar 

  28. Laffont, J.-J.: Fundamentals of Public Economics. MIT Press, Cambridge (1988)

    Google Scholar 

  29. Revuz, D., Yor, M.: Continuous Martingales and Brownian Motion, 3rd edn. Springer, Berlin (1999)

    Book  MATH  Google Scholar 

  30. Riedel, F., Su, X.: On irreversible investment. Financ. Stoch. 15(4), 607–633 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  31. Steg, J.-H.: Irreversible investment in oligopoly. Financ. Stoch. 16(2), 207–224 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  32. Topkis, D.M.: Minimizing a submodular function on a lattice. Oper. Res. 26, 305–321 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  33. Varian, H.R.: Sequential contributions to public goods. J. Public Econ. 53, 165–186 (1994)

    Article  Google Scholar 

  34. Wang, W.-K., Ewald, C.O.: Dynamic voluntary provision of public goods with uncertainty: a stochastic differential game model. Decis. Econ. Financ. 33, 97–116 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  35. Yeung, D.W.K., Petrosyan, L.A.: Subgame consistent cooperative provision of public goods. Dyn. Games Appl. 3(3), 419–442 (2013)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The authors wish to thank two anonymous referees for their pertinent and useful comments. Financial support by the German Research Foundation (DFG) via Grant Ri 1142-4-2 is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giorgio Ferrari.

Appendix 1: Some Proofs and Technical Results

Appendix 1: Some Proofs and Technical Results

1.1 On the Proof of Proposition 3.3

In this section we prove Proposition 3.3. The proof is a generalization of Theorem 3.2 in [8] to the case of a multivariate optimal consumption problem with both monotone and classical absolutely continuous controls. Sufficiency easily follows from concavity of the utility functions \(u^i\), \(i=1,\ldots ,n\). On the other hand, the next Lemma accomplishes the proof of the necessity part. Necessity is proved by linearizing the original problem (3.2) around its optimal solution \((\underline{x}_{*}, \underline{C}_{*})\), by showing that \((\underline{x}_{*}, \underline{C}_{*})\) solves the linearized problem as well and that it satisfies some flat-off conditions as those of (3.8).

Recall the notation \(x:=\sum _{i=1}^n x^i\) and \(C:=\sum _{i=1}^n C^i\).

Lemma 6.1

Let Assumptions 1 and 2 hold and \((\underline{x}_{*}, \underline{C}_{*}) \in \mathcal {B}_w\) be optimal for problem (3.2) and set

$$\begin{aligned} \Psi _{*}(t):= E\biggl [\int _{t}^T e^{-\int _0^s r(u)\,du} \sum _{i=1}^n \gamma ^i\,u^{i}_c \big (x^i_{*}(s), C_{*}(s) \big )\,ds\Big |{\mathscr {F}}_{t}\biggr ]. \end{aligned}$$
(6.1)

Then \((\underline{x}_{*}, \underline{C}_{*})\)

  1. i.

    solves the linear optimization problem

    $$\begin{aligned} \sup _{(\underline{x}, \underline{C}) \in \mathcal {B}_w} E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_x \big (x^i_{*}(t),C_{*}(t) \big ) x^i(t)dt + \int _0^T\Psi _{*}(t)dC(t)\biggr ];\nonumber \\ \end{aligned}$$
    (6.2)
  2. ii.

    satisfies

    $$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \Big (e^{-\int _0^t r(s)ds} \gamma ^i u^i_x(x^i_{*}(t),C_{*}(t)) - M\psi _x(t)\Big )x^i_*(t) = 0 \quad \,\,P\otimes dt\text {-a.e.},\, i=1,\ldots ,n, \\ \displaystyle E\biggl [\int _0^T \Big (\Psi _{*}(t) - M \psi _c(t)\Big )dC_{*}(t)\biggr ] = 0, \end{array} \right. \end{aligned}$$
    (6.3)

    with

    $$\begin{aligned} M:={}&(P\otimes dt)\text {-}{{\mathrm{ess\,sup}}}\Biggl [\max \Biggl \{\frac{e^{-\int _0^t r(s)ds} \gamma ^i u^i_x(x^i_{*}(t),C_{*}(t))}{\psi _x(t)};\; i=1,\ldots ,n \Biggr \}\Biggr ] \nonumber \\ \vee \;&P\text {-}{{\mathrm{ess\,sup}}}\Biggl [\sup _{t\in [0,T]} \frac{\Psi _{*}(t)}{\psi _c(t)}\Biggr ]. \end{aligned}$$
    (6.4)

Proof

The proof splits into two steps.

  • Step 1. Let \((\underline{x}_{*}, \underline{C}_{*}) \in \mathcal {B}_w\) be optimal for problem (3.2). For \((\underline{x}, \underline{C}) \in \mathcal {B}_w\) and \(\epsilon \in [0,1]\), define the admissible strategy \((\underline{x}_{\epsilon }, \underline{C}_{\epsilon })\) with \(\underline{x}_{\epsilon }(t):= \epsilon \underline{x}(t) + (1-\epsilon )\underline{x}_{*}(t)\) and such that \(C_{\epsilon }(t)=\epsilon C(t) + (1-\epsilon )C_{*}(t)\). Notice that \(\underline{x}_{\epsilon }(t)\) and \(C_{\epsilon }(t)\) respectively converge to \(\underline{x}_{*}(t)\) and \(C_{*}(t)\) for all \(t \in [0,T]\) a.s. when \(\epsilon \downarrow 0\). Now, optimality of \((\underline{x}_{*}, \underline{C}_{*})\), Assumption 1.ii, concavity of \(u^i\) and an application of Fubini’s Theorem allow us to write

    $$\begin{aligned} 0\ge & {} \frac{1}{\epsilon }[U_{SP}(\underline{x}_{\epsilon }, \underline{C}_{\epsilon }) - U_{SP}(\underline{x}_{*}, \underline{C}_{*})] \nonumber \\\ge & {} E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x\big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )\big (x^i(t) - x^i_{*}(t)\big )\,dt\biggr ] \nonumber \\&+\, E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_c \big (x^i_{\epsilon }(t), C_{\epsilon }(t)\big ) \big (C(t) - C_{*}(t)\big )\,dt\biggr ]\nonumber \\= & {} E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big ) \big (x^i(t) - x^i_{*}(t) \big )\,dt\biggr ] \nonumber \\&+ E\biggl [\int _0^T \Phi _{\epsilon }(t) \big (dC(t) - dC_{*}(t) \big )\biggr ], \end{aligned}$$
    (6.5)

    where \(\Phi _{\epsilon }(t):= \int _{t}^T e^{-\int _0^s r(u)\,du} \sum _{i=1}^n \gamma ^i\,u^{i}_c(x^i_{\epsilon }(s), C_{\epsilon }(s))\,ds\). One has

    $$\begin{aligned}&\liminf _{\epsilon \downarrow 0} E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )x^i(t) dt \biggr ] \\&\quad \ge E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{*}(t), C_{*}(t) \big )x^i(t) dt \biggr ] \end{aligned}$$

    and

    $$\begin{aligned} \liminf _{\epsilon \downarrow 0} E\biggl [\int _0^T \Phi _{\epsilon }(t) dC(t) \biggr ] \ge E\biggl [\int _0^T \Phi _{*}(t) dC(t) \biggr ], \end{aligned}$$

    with \(\Phi _{*}:=\Phi _{0}\), by Fatou’s Lemma. We now claim (and we prove it later) that

    $$\begin{aligned}&\lim _{\epsilon \downarrow 0} E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )x^i_{*}(t) dt \biggr ] \nonumber \\&= E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{*}(t), C_{*}(t)\big )x^i_{*}(t) dt \biggr ] \end{aligned}$$
    (6.6)

    and

    $$\begin{aligned} \lim _{\epsilon \downarrow 0} E\biggl [\int _0^T \Phi _{\epsilon }(t) dC_{*}(t) \biggr ] = E\biggl [\int _0^T \Phi _{*}(t) dC_{*}(t) \biggr ]. \end{aligned}$$
    (6.7)

    Hence from (6.5)

    $$\begin{aligned}&E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{*}(t), C_{*}(t) \big )x^i(t) \,dt\biggr ] + E\biggl [\int _0^T \Phi _{*}(t)dC(t)\biggr ] \\&\quad \! \le E\biggl [\int _{0}^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{*}(t), C_{*}(t) \big )x^i_{*}(t) \,dt\biggr ] + E\biggl [\int _0^T \Phi _{*}(t)dC_{*}(t)\biggr ] \end{aligned}$$

    and by replacing \(\Phi _{*}\) with its optional projection \(\Psi _{*}\) as defined in (6.1) (cf. [22], Theorem 1.33) it follows that \((\underline{x}_{*}, \underline{C}_{*})\) is optimal for problem (6.2) as well. To conclude the proof we must prove (6.6) and (6.7). To prove (6.6) it suffices to show that the family \((\Gamma ^1_{\epsilon })_{\epsilon \in [0,\frac{1}{2}]}\) given by

    $$\begin{aligned} \Gamma ^1_{\epsilon }(t):= e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )x^i_{*}(t) \end{aligned}$$

    is \(P\otimes dt\)-uniformly integrable. Concavity of \(u^i\) and the fact that \(x^{i}_{\epsilon }(t) \ge \frac{1}{2}x^i_{*}(t)\) a.s. for \(\epsilon \in [0,\frac{1}{2}]\) and every \(t \in [0,T]\) lead to

    $$\begin{aligned}&\Gamma ^1_{\epsilon }(t) \le 2 e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_x( x^i_{\epsilon }(t),C_{\epsilon }(t))x^i_{\epsilon }(t) \nonumber \\&\quad \le 2 e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i \bigl [u^{i} \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )-u^{i}(0, C_{\epsilon }(t))\bigr ], \end{aligned}$$

    and the last term in the right-hand side above is \(P\otimes dt\)-uniformly integrable by Assumption 1.ii. Then (6.6) holds by Vitali’s Convergence Theorem. As for (6.7) note that by Fubini’s Theorem

    $$\begin{aligned} \int _0^T \Phi _{\epsilon }(t)dC_{*}(t) = \int _0^T e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_c \big (x^i_{\epsilon }(t), C_{\epsilon }(t) \big )C_{*}(t)dt. \end{aligned}$$

    Hence, to have (6.7) it suffices to show that the family

    $$\begin{aligned} \Gamma ^2_{\epsilon }(t):= e^{-\int _0^t r(s)\,ds} \sum _{i=1}^n \gamma ^i u^{i}_c(x^i_{\epsilon }(t), C_{\epsilon }(t))C_{*}(t) \end{aligned}$$

    is \(P\otimes dt\)-uniformly integrable, but this follows by employing arguments similar to those used for \((\Gamma ^1_{\epsilon })_{\epsilon \in [0,\frac{1}{2}]}\).

  • Step 2. We now show that the flat-off conditions (6.3) hold for any solution \((\hat{\underline{x}}, \hat{\underline{C}})\) of the linear problem (6.2). Then, by Step 1, they also hold for \((\underline{x}_{*},\underline{C}_{*})\). Notice that for every \((\underline{x}, \underline{C}) \in \mathcal {B}_w\) one has

    $$\begin{aligned}&E\biggl [\int _0^T \sum _{i=1}^n e^{-\int _0^t r(s)ds} \gamma ^i u^i_x(x^i_{*}(t),C_{*}(t)) x^i(t)dt + \int _0^T \Psi _{*}(t) dC(t)\biggr ] \nonumber \\&\quad \le M E\biggl [\int _0^T \sum _{i=1}^n \psi _x(t) x^i(t) dt + \int _0^T \psi _c(t)dC(t)\biggr ] = Mw \end{aligned}$$
    (6.8)

    by definition of M (cf. (6.4)). Obviously, if \((\underline{x}, \underline{C})\) satisfies (6.3) we then have equality in (6.8). On the other hand, if

    $$\begin{aligned}&\sup _{(\underline{x}, \underline{C}) \in \mathcal {B}_w} E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_x(x^i_{*}(t),C_{*}(t)) x^i(t)dt + \int _0^T \Psi _{*}(t) dC(t)\biggr ]\nonumber \\&\quad = Mw, \end{aligned}$$
    (6.9)

    then equality holds through (6.8) and we obtain (6.3). It therefore remains to prove (6.9). To this end take \(K<M\) and define the investment strategies

    $$\begin{aligned} x^{i}_K(t):= {\left\{ \begin{array}{ll} \alpha &{} \text {if } e^{-\int _0^t r(s)ds} \gamma ^i u^i_x(x^i_{*}(t),C_{*}(t)) \ge K \psi _x(t), \\ 0 &{} \text {else} \end{array}\right. } \quad \text {and}\quad C_K(t):=\alpha \mathbbm {1}_{[\sigma _K,T]}(t), \end{aligned}$$

    with the stopping time

    $$\begin{aligned} \sigma _K:= \inf \{t \in [0,T): \Psi _{*}(t) \ge K \psi _c(t)\} \wedge T \end{aligned}$$

    and some \(\alpha \) such that \(E[\int _0^T \sum _{i=1}^n \psi _x(t)x^i_K(t) dt + \int _0^T \psi _c(t)dC_K(t)] =w\). Note that one can find such \(\alpha \). Indeed, suppose to the contrary that for \(\alpha =1\), \(\max \{x^{i}_K(t);i=1,\ldots ,n\}=C_K(t)=0\) on \([0,T) P \otimes dt-\)a.e. for some \(K>0\). This would mean \(M \le K\) by the definition of M (cf. (6.4)). We now have

    $$\begin{aligned} Mw\ge & {} \sup _{(\underline{x}, \underline{C}) \in \mathcal {B}_w} E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_x \big (x^i_{*}(t),C_{*}(t) \big ) x^i(t)dt + \int _0^T \Psi _{*}(t) dC(t)\biggr ] \\\ge & {} E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_x \big (x^i_{*}(t),C_{*}(t) \big ) x^i_K(t)dt + \int _0^T \Psi _{*}(t) dC_K(t)\biggr ]\\\ge & {} K\,E\biggl [\int _0^T \sum _{i=1}^n \psi _x(t)x^i_K(t) dt + \alpha \psi _c(\sigma _K)\mathbbm {1}_{\{\sigma _K < T\}}\biggr ]\\\ge & {} K\,E\biggl [\int _0^T \sum _{i=1}^n \psi _x(t)x^i_K(t) dt + \int _0^T \psi _c(t)dC_K(t)\biggr ] = Kw, \end{aligned}$$

    which yields (6.9) by letting \(K \uparrow M\).\(\square \)

We are now able to prove Proposition 3.3.

Proof of Proposition 3.3

Proof

Sufficiency follows from concavity of utility function \(u^i\), \(i=1,\ldots ,n\), (cf. Assumption 1). Indeed, for \((\underline{x}_{*}, \underline{C}_{*}) \in \mathcal {B}_w\) satisfying (3.8) and for \((\underline{x}, \underline{C})\) any other admissible policy we may write

$$\begin{aligned}&U_{SP}(\underline{x}_{*}, \underline{C}_{*}) - U_{SP}(\underline{x}, \underline{C}) \\&\quad \ge E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_{x} \big (x^i_{*}(t),C_{*}(t) \big )(x^i_{*}(t) - x^i(t))dt\biggr ] \\&\quad \quad + E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_{c} \big (x^i_{*}(t),C_{*}(t) \big )(C_{*}(t) - C(t)) dt\biggr ] \\&\quad = E\biggl [\int _0^T e^{-\int _0^t r(s)ds} \sum _{i=1}^n \gamma ^i u^i_{x} \big (x^i_{*}(t),C_{*}(t) \big )(x^i_{*}(t) - x^i(t))dt\biggr ] \\&\quad \quad + E\biggl [\int _0^T \biggl (\int _{t}^T e^{-\int _0^s r(u)\,du} \sum _{i=1}^n \gamma ^i u^{i}_c \big (x^i_{*}(s), C_{*}(s)\big )\,ds\biggr ) \big (dC_{*}(t) - dC(t)\big )\biggr ] \\&\quad \ge \lambda (w - w) = 0, \end{aligned}$$

where (3.8) and Fubini’s Theorem lead to the second inequality, whereas the last one is implied by the first and the fourth of (3.8) and by the budget constraint. Finally, Lemma 6.1 yields the proof of the necessity part. \(\square \)

1.2 A Useful Simple Lemma

Lemma 6.2

Let Assumption 1 hold and set \(h^i(\psi ,c):=u_c^i(g(\psi ,c),c)\) for every \(i=1,\ldots ,n\), \(\psi ,c>0\), where \(g^i(\cdot ,c)\) is the inverse of \(u_x^i(\cdot ,c)\). Then \(c \mapsto h^i(\psi ,c)\) is strictly decreasing for any \(\psi >0\). Moreover, if also \(u_{xc}\ge 0\), then \(\psi \mapsto h(\psi ,c)\) is nonincreasing for any \(c>0\).

Proof

\(h^i\) is strictly decreasing in c by strict concavity of \(u^i\). Indeed, since \(u^i_x(g^i(\psi ,c),c)=\psi \) is constant in c for all \(\psi >0\), by implicit differentiation \(g^i_c(\psi ,c)=-u^i_{xc}(g^i(\psi ,c),c)/u^i_{xx}(g^i(\psi ,c),c)\). Then \(h^i_c(\psi ,c)=u^i_{xc}(g^i(\psi ,c),c)g^i_c(\psi ,c)+u^i_{cc}(g^i(\psi ,c),c)=-\bigl [u^i_{xc}(g^i(\psi ,c),c)\bigr ]^2/u^i_{xx}(g^i(\psi ,c),c)+u^i_{cc}(g^i(\psi ,c),c)<0\), as the Hessian of \(u^i\) is negative definite by Assumption 1.i.

On the other hand, \(h^i(\psi ,c)=u^i_c(g^i(\psi ,c),c)\) is nonincreasing in \(\psi \) if also \(u^i_{xc}\ge 0\), since \(g^i(\cdot ,c)\) is strictly decreasing like \(u^i_x(\cdot ,c)\), which it is the inverse of. \(\square \)

1.3 Proof of Proposition 3.4

Proof

For any given \(\lambda > 0\), set \(X(t):=\lambda \psi _c(t)\mathbbm {1}_{\{t < T\}}\). Such a process vanishes at T, it is of class (D) and lower semicontinuous in expectation by Assumption 2. Moreover, we define the atomless, optional random Borel measure \(\mu (\omega ,dt):= e^{-\int _0^t r(\omega ,s) ds}dt\) and the random field

$$\begin{aligned} f^{i}(\omega ,t,\ell ):= {\left\{ \begin{array}{ll} h^i(\frac{\lambda }{\gamma ^i}e^{\int _0^t r(\omega ,s) ds}\psi _x(\omega ,t), -\frac{1}{\ell }) &{} \text {if } \ell <0, \\ -\ell &{} \text {if } \ell \ge 0, \end{array}\right. } \end{aligned}$$

for some \(\gamma ^i>0\). Notice that \((\omega ,t)\mapsto f^{i}(\omega ,t,\ell )\) is progressively measurable and \(P\otimes \mu (dt)\) integrable for any given and fixed \(\ell \in \mathbb {R}\). Moreover, since \(c \mapsto h^i(\psi , c)\) is strictly decreasing (cf. Lemma 6.2) and, by assumption, satisfies the Inada conditions

$$\begin{aligned} \lim _{c \downarrow 0}h^i(\psi ,c)=+\infty , \qquad \lim _{c \uparrow \infty }h^i(\psi ,c)=0, \end{aligned}$$
(6.10)

then \(f^{i}(\omega ,t,\cdot )\) is strictly decreasing from \(+\infty \) to \(-\infty \). All these properties are clearly inherited by the function \(\sum _{i=1}^n \gamma ^i h^i(\psi , \cdot )\), being \(\gamma ^i > 0\), \(i=1,\ldots ,n\).

Following the arguments, e.g., in the proof of Proposition 3.4 of [21] (see also the proof of Theorem 2.4 in [9]), we can apply Theorem 3 of [5] to have existence of an optional signal process \(l^*\) solving (3.11). Such a process is also upper right-continuous and therefore it is unique up to indistinguishability by [5], Theorem 1, and Meyer’s optional section theorem (see, e.g., [18], Theorem IV.86) (cf. again [21], proof of Proposition 3.4, or [7], proof of Theorem 1). \(\square \)

1.4 Proof of Proposition 5.4

Due to Theorem 4.4, to find the Nash equilibrium strategy of the public good contribution game (4.1) in our homogeneous and symmetric setting it suffices to solve backward equation (4.9).

Recall that \(h^i(\psi ,c):=u^i_c(g^i(\psi ,c),c)\) with \(g^i(\cdot , c)\) the inverse of \(u^i_x(\cdot , c)\). For any \(\lambda ^i > 0\), straightforward computations lead to \(h^i(\lambda ^i e^{rt}\psi _x(t), C(t)) = \delta (\lambda ^i \mathcal {E}_x(t))^{\frac{\alpha }{\alpha -1}}C^{\frac{\alpha + \beta -1}{1 - \alpha }}(t)\), with \(\delta :=\frac{\beta }{\alpha }\left( \frac{\alpha + \beta }{\alpha }\right) ^{\frac{1}{\alpha -1}}\). Set \(\hat{C}^i(t)= \sup _{0\le s \le t}\hat{l}(s) \vee 0\) for some progressively measurable process \(\hat{l}\) solving

$$\begin{aligned} E\biggl [\int _{\tau }^{\infty } \delta e^{-rs} (\lambda ^i \mathcal {E}_x(s))^{\frac{\alpha }{\alpha -1}} \Big ( n\sup _{\tau \le u \le s} \hat{l}(u)\Big )^{\frac{\alpha + \beta -1}{1 - \alpha }} ds \biggm |{\mathscr {F}}_{\tau }\biggr ] = \lambda ^i e^{-r\tau } \mathcal {E}_c(\tau ), \end{aligned}$$

i.e.,

$$\begin{aligned} E\biggl [\int _{0}^{\infty } \delta e^{-ru} (\lambda ^i)^{\frac{\alpha }{\alpha -1}} \frac{{\mathcal {E}_x}^{\frac{\alpha }{\alpha -1}}(u+\tau )}{\mathcal {E}_c(\tau )} \inf _{0 \le s \le u}\Big ( (n{\hat{l}})^{\frac{\alpha + \beta -1}{1 - \alpha }}(s+\tau )\Big ) du \biggm |{\mathscr {F}}_{\tau }\biggr ] = \lambda ^i.\nonumber \\ \end{aligned}$$
(6.11)

Now take \(\hat{l}(t):=\frac{\kappa }{n}{\mathcal {E}_{c}}^{\frac{1 -\alpha }{\alpha + \beta -1}}(t){\mathcal {E}_x}^{\frac{\alpha }{\alpha + \beta -1}}(t)\) for some constant \(\kappa \) and use independence and stationarity of Lévy increments to rewrite (6.11) as

$$\begin{aligned} \kappa ^{\frac{\alpha + \beta - 1}{1 - \alpha }} E\biggl [\int _{0}^{\infty } \delta e^{-ru} \inf _{0 \le s \le u}\left( \mathcal {E}_c(s)\mathcal {E}_x^{\frac{\alpha }{\alpha -1}}(u-s)\right) du \biggr ] = {(\lambda ^i)}^{\frac{1}{1-\alpha }}. \end{aligned}$$
(6.12)

We claim (and we discuss later) that (cf. (5.17)) \(A:= E[\int _{0}^{\infty }\delta e^{- r u} \inf _{0 \le s \le u} \left( \mathcal {E}_c(s) \mathcal {E}_x^{-\frac{\alpha }{1 - \alpha }}(u-s)\right) du]\) is finite. Then by solving (6.12) for \(\lambda ^i\) one obtains

$$\begin{aligned} \lambda ^i= A^{1-\alpha } \kappa ^{\alpha + \beta - 1}=:\hat{\lambda }^i. \end{aligned}$$

But now \(\hat{x}^i(t) = [\hat{\lambda }^i \left( \frac{\alpha + \beta }{\alpha }\right) \mathcal {E}_x(t) \hat{C}^{-\beta }(t)]^{\frac{1}{\alpha - 1}}\), and therefore

$$\begin{aligned} \hat{x}^i(t) = (\hat{\lambda }^i)^{-\frac{1}{1-\alpha }}\biggl [\left( \frac{\alpha + \beta }{\alpha }\right) \mathcal {E}_x(t) {\kappa }^{-\beta }\inf _{0 \le s \le t}\left( \mathcal {E}_c^{\frac{\beta (1-\alpha )}{1-\alpha - \beta }}(s)\mathcal {E}_x^{\frac{\alpha \beta }{1-\alpha - \beta }}(s)\right) \biggr ]^{-\frac{1}{1-\alpha }}; \end{aligned}$$

that is,

$$\begin{aligned} \hat{x}^i(t) = \kappa \gamma (t) \end{aligned}$$
(6.13)

with \(\gamma (t)\) as in (5.14).

To determine \(\kappa \) we use the budget constraint \(\mathbb {E}[\int _{0}^{\infty }\psi _x(t) \hat{x}^i(t) dt + \int _{0}^{\infty }\psi _c(t) d\hat{C}^i(t)] = w\). Indeed, by (6.13) we have

$$\begin{aligned} \kappa E\biggl [\int _0^{\infty } \psi _x(t) \gamma (t) dt + \frac{1}{n}\int _{0}^{\infty }\psi _c(t)d\theta (t)\biggr ] = w, \end{aligned}$$
(6.14)

since \(\hat{C}^i(t) = \sup _{0 \le s \le t}\hat{l}(s) = \frac{\kappa }{n} \sup _{0 \le s \le t}(\mathcal {E}_c^{-\frac{1 - \alpha }{1 - \alpha - \beta }}(s) \mathcal {E}_x^{-\frac{\alpha }{1 - \alpha - \beta }}(s)) = \frac{\kappa }{n} \theta (t)\) with \(\theta (t)\) as in (5.15), and if \(E[\int _0^{\infty } \psi _x(t) \gamma (t) dt + \frac{1}{n}\int _{0}^{\infty }\psi _c(t)d\theta (t)]<\infty \). Now the result follows by solving (6.14) for \(\kappa \). Finally, arguments as those in the proof of Proposition 5.3 allow to show that under Assumption 3.4. all the quantities above are finite, thus completing the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ferrari, G., Riedel, F. & Steg, JH. Continuous-Time Public Good Contribution Under Uncertainty: A Stochastic Control Approach. Appl Math Optim 75, 429–470 (2017). https://doi.org/10.1007/s00245-016-9337-5

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00245-016-9337-5

Keywords

JEL Classification

Mathematics Subject Classification

Navigation