Skip to main content
Log in

Optimal Consumption Under Deterministic Income

  • Published:
Journal of Optimization Theory and Applications Aims and scope Submit manuscript

Abstract

We consider an individual or household endowed with an initial wealth, having an income and consuming goods and services. The wealth development rate is assumed to be a deterministic continuous function of time. The objective is to maximize the discounted consumption over a finite time horizon. Via the Hamilton–Jacobi–Bellman approach, we prove the existence and the uniqueness of the solution to the considered problem in the viscosity sense. Furthermore, we derive an algorithm for explicit calculation of the value function and optimal strategy. It turns out that the value function is in general not continuous. The method is illustrated by two examples.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Fleming, W.H., Rishel, R.W.: Deterministic and Stochastic Optimal Control. Springer, New York (1975)

    Book  MATH  Google Scholar 

  2. Bertsekas, D.P.: Dynamic Programming: Deterministic and Stochastic Models. Prentice Hall, New York (1987)

    MATH  Google Scholar 

  3. Bardi, M., Capuzzo-Dolcetta, I.: Optimal Control and Viscosity Solutions of Hamilton–Jacobi–Bellman Equations. Birkhäuser, Boston (1997)

    Book  MATH  Google Scholar 

  4. Schmidli, H.: Stochastic Control in Insurance. Springer, London (2008)

    MATH  Google Scholar 

  5. Giannessi, F., Maugeri, A. (eds.): Variational Analysis and Applications. Springer, New York (2005)

    MATH  Google Scholar 

  6. Aoki, M.: Optimal Control and System Theory in Dynamic Economic Analysis. North Holland, New York (1976)

    MATH  Google Scholar 

  7. Bryson, A.E. Jr., Ho, Y.: Applied Optimal Control. Blaisdell, Waltham (1969)

    Google Scholar 

  8. Chow, G.C.: Analysis and Control of Dynamic Economic Systems. Wiley, New York (1975)

    MATH  Google Scholar 

  9. Dixit, A.K.: Optimization in Economic Theory. Oxford University Press, Oxford (1990)

    Google Scholar 

  10. Dorfman, R.: An economic interpretation of optimal control theory. Am. Econ. Rev. 59(5), 817–831 (1969)

    MathSciNet  Google Scholar 

  11. Macki, J.W., Strauss, A.: Introduction to Optimal Control Theory. Springer, New York (1982)

    Book  MATH  Google Scholar 

  12. Intriligator, M.D.: Applications of optimal control theory in economics. Synthese 31, 271–288 (1975)

    Article  MATH  MathSciNet  Google Scholar 

  13. Kendrick, D.A.: Stochastic Control for Economic Models. McGraw-Hill, New York (1981)

    MATH  Google Scholar 

  14. Kendrick, D.A.: Stochastic control for economic models: past, present and the paths ahead. J. Econ. Dyn. Control 29, 3–30 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  15. Murray, R.M. (ed.): Control in an Information Rich World: Report of the Panel on Future Directions in Control, Dynamics, and Systems. SIAM, Philadelphia (2003)

    Google Scholar 

  16. Neck, R.: Control theory and economic policy: balance and perspectives. Annu. Rev. Control 33(1), 79–88 (2009)

    Article  Google Scholar 

  17. Moyer, H.G.: Deterministic Optimal Control: an Introduction for Scientists. Trafford, Victoria (2004)

    Google Scholar 

  18. Seierstad, A., Sydsæter, K.: Optimal Control Theory with Economic Applications. North-Holland, Amsterdam (1987)

    MATH  Google Scholar 

  19. Weber, T.A.: Optimal Control Theory with Applications in Economics. MIT Press, Cambridge (2011). Foreword by A.V. Kryazhimskiy

    Book  MATH  Google Scholar 

  20. Karatzas, I., Lehoczky, J.P., Sethi, S.P., Shreve, S.E.: Explicit solution of a general consumption/investment problem. Math. Oper. Res. 11, 261–294 (1986)

    Article  MathSciNet  Google Scholar 

  21. Karatzas, I., Lehoczky, J.P., Shreve, S.E.: Optimal portfolio and consumption decisions for a “small investor” on a finite horizon. SIAM J. Control Optim. 25, 1157–1586 (1987)

    Article  MathSciNet  Google Scholar 

  22. Cox, J.C., Huang, C.F.: Optimal consumption and portfolio policies when asset prices follow a diffusion process. J. Econ. Theory 49, 33–83 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  23. Avanzi, B.: Strategies for dividend distribution: a review. N. Am. Actuar. J. 13, 217–251 (2009)

    Article  MathSciNet  Google Scholar 

  24. Albrecher, H., Thonhauser, S.: Optimality results for dividend problems in insurance. RACSAM Rev. R. Acad. Cienc. Exactas Fis. Nat. Ser. A Mat. 103(2), 295–320 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  25. Grandits, P.: Optimal consumption in a Brownian model with absorption and finite time horizon. Appl. Math. Optim. 67(2), 197–241 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  26. Barron, E.N., Jensen, R.: Semicontinuous viscosity solutions for Hamilton–Jacobi equations with convex Hamiltonians. Commun. Partial Differ. Equ. 15(12), 293–309 (1990)

    Article  MathSciNet  Google Scholar 

  27. Evans, L.C.: Partial Differential Equations. Am. Math. Soc., Providence (1998)

    MATH  Google Scholar 

  28. Fleming, W.H., Soner, H.M.: Controlled Markov Processes and Viscosity Solutions. Springer, New York (1993)

    MATH  Google Scholar 

  29. Sontag, E.D.: Mathematical Control Theory. Springer, New York (1990)

    Book  MATH  Google Scholar 

  30. Zabczyk, J.: Mathematical Control Theory: an Introduction. Birkhäuser, Boston (1992)

    MATH  Google Scholar 

  31. Sulem, A.: Explicit solution of a two-dimensional deterministic inventory problem. Math. Oper. Res. 11(1), 134–146 (1986)

    Article  MATH  MathSciNet  Google Scholar 

  32. Sulem, A.: Résolution explicite d’Inéquations quasi-variationelles associées à des problèmes de gestion de stock. Thèse 3e cycle. Paris IX-Dauphine (1983)

  33. Sulem, A.: A solvable one-dimensional model of a diffusion inventory system. Math. Oper. Res. 11, 126–133 (1984)

    Google Scholar 

  34. Amman, H.: Numerical methods for linear-quadratic models. In: Handbook of Computational Economics, vol. 1, pp. 587–618. North-Holland, Amsterdam (1996)

    Google Scholar 

  35. Bellman, R.: An Introduction to the Theory of Dynamic Programming. Rand, Santa Monica (1953)

    MATH  Google Scholar 

  36. Crandall, M.G., Ishii, H., Lions, P.: User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. 27(1), 1–67 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  37. Paxton, C.: Consumption and income seasonality in Thailand. J. Polit. Econ. 101(1), 39–72 (1993)

    Article  Google Scholar 

  38. Lutz, G.A.: Business Cycle Theory. Oxford University Press, Oxford (2002)

    Google Scholar 

  39. Mnif, M., Sulem, A.: Optimal risk control and dividend policies under excess of loss reinsurance. Stochastics 77(5), 455–476 (2005)

    MATH  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors thank anonymous referees for their comments and suggestions which helped significantly to improve the presentation of the paper.

The research of the first two authors was supported by the Austrian Science Fund, grant P22449-N18.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julia Eisenberg.

Additional information

Communicated by Vladimir Veliov.

Appendix

Appendix

1.1 HJB Equation—Viscosity Solution

Proof of Theorem 3.1

The proof of Theorem 3.1 consists of two steps: the super- and subsolution proofs. The following lemma presents the subsolution direction; the supersolution proof does not include subtleties and follows standard techniques, see Mnif and Sulem [39]. □

Lemma A.1

The function V(t,x) given by (1) is a viscosity subsolution to (13).

Proof

We need to show, that for every \(\bar{\psi}\in C^{(1,1)}[0,T]\times[0,\infty[\):

at every \((\bar{t},\bar{x})\in\,]0,T[\times]0,\infty[\) which is a (strict) maximizer of \(V^{*}-\bar{\psi}\) on [0,T]×[0,∞[ with \(V^{*}(\bar {t},\bar{x})=\bar{\psi}(\bar{t},\bar{x})\).

As usual, the subsolution proof is done by way of contradiction. Suppose there are some \((\bar{t},\bar{x})\) and \(\bar{\psi}\) with the properties stated before but with

(14)

for some ξ>0. Consider the function \(\psi(t,x)=\bar{\psi}(t,x)+\frac{(x-\bar{x})^{2}+(t-\bar {t})^{2}}{\bar{t}^{2}+\bar{x}^{2}}\xi\).

Then it holds that \(V^{*}(\bar{t},\bar{x})=\bar{\psi}(\bar{t},\bar {x})=\psi(\bar{t},\bar{x}),\bar{\psi}_{x}(\bar{t},\bar{x})=\psi _{x}(\bar{t},\bar{x})\) and \(\bar{\psi}_{t}(\bar{t},\bar{x})=\psi _{t}(\bar{t},\bar{x})\), which gives

Since ψ(t,x) is continuously differentiable in both t and x and μ t is continuous, there is \(\delta\in\,]0,\frac{\sqrt{\bar{t}^{2}+\bar{x}^{2}}}{2}[\) such that

for \((t,x)\in B_{\delta}(\bar{t},\bar{x})\). We obtain \(V^{*}(t,x)\leq\bar{\psi}(t,x)=\psi(t,x)-\frac{\delta^{2}}{\bar {t}^{2}+\bar{x}^{2}}\xi\) for \((t,x)\in\partial B_{\delta}(\bar{t},\bar {x})\).

Let now \(\varepsilon=\frac{1}{2}\frac{\delta^{2}}{\bar{t}^{2}+\bar {x}^{2}}\xi\), then on \(B_{\delta}(\bar{t},\bar{x})\) we have

(15)

while for \((t,x)\notin B_{\delta}(\bar{t},\bar{x})\) we have V (t,x)≤ψ(t,x)−2ε.

Now let \((t_{n},x_{n})\to(\bar{t},\bar{x})\) such that \(V(t_{n},x_{n})\to V^{*}(\bar{t},\bar{x})\) and assume (w.l.o.g.) that \((t_{n},x_{n})\in B_{\delta}(\bar{t},\bar{x})\) for all \(n\in\mathbb{N}\).

Let \(C^{n}\in\mathcal{C}(t_{n},x_{n})\), \(X^{C^{n}}\) be the corresponding wealth starting in (t n ,x n ) and \(\tau^{*}=\tau_{n}\wedge\bar{T}\) (with some \(\bar{T}\) such that \(t_{n}<\bar{T}\leq T\) for all n) where

At first we observe that \(X^{C^{n}}\) can only have downward jumps in the x direction and that V(t,x) is increasing in x. Because of continuity of μ t , jumps in the wealth process are due to jumps in the consumption process, and we have \(X_{s}^{C^{n}}-X_{s-}^{C^{n}}=-\Delta C^{n}_{s}\).

Suppose τ =τ n , i.e. stopping because of leaving \(B_{\delta }(\bar{t},\bar{x})\). Then either we hit the boundary continuously or leave the ball due to a jump at time τ in which case \(X_{\tau ^{*}-}\in B_{\delta}(\bar{t},\bar{x})\). From the above estimates we get \(V(\tau^{*},X_{\tau^{*}-}^{C_{n}})\leq\psi (\tau^{*},X_{\tau^{*}-}^{C_{n}})-2\varepsilon I_{\{X_{\tau^{*}-}=X_{\tau ^{*}}\}}\).

If \(\tau^{*}=\bar{T}\) then \(X_{\tau^{*}}^{C_{n}}\), as well as \(X_{\tau ^{*}-}^{C_{n}}\), is still inside the ball and we have \(V(\tau^{*},X_{\tau^{*}-}^{C_{n}})\leq\psi(\tau^{*},X_{\tau^{*}-}^{C_{n}})\). In total, we arrive at

In the above formula C n,c denotes the continuous part of strategy C n.

If \(X_{s}^{C_{n}}\neq X_{s-}^{C^{n}}\), we have that \(X_{s}^{C^{n}}- X_{s-}^{C^{n}}=-\Delta C_{s}^{n}\) and we can write

Combining the last expression with the continuous part of C n and using e βtψ x (t,x) on \(B_{\delta}(\bar{t},\bar {x})\), we arrive at

Finally, using \(\psi_{t}(s,X_{s}^{C^{n}})+\mu_{s} \psi_{x}(s,X_{s}^{C^{n}})\leq -\varepsilon\) from (15) we get

which is the same as

Since also \(\psi(t_{n},x_{n})\to V^{*}(\bar{t},\bar{x})\) if n→∞, we can choose n large enough such that \(0\leq\psi(t_{n},x_{n})-V(t_{n},x_{n})\leq\frac{(\tau^{*}-t_{n}) \varepsilon +2\varepsilon I_{\{\tau_{n}=\tau^{*}\wedge X_{\tau^{*}-}=X_{\tau^{*}}\}}}{2}\). We get

(16)

Now, before we can state that (16) is a contradiction to (12), we have to discuss the case τ =τ n =t n , where an immediate lump-sum consumption leads to

$$x_n=X_{\tau^*-}^{C^n}>X_{\tau^*}^{C^n} \notin B_\delta(\bar{t},\bar {x}). $$

We notice that property (P6) and the construction of the value function show that in this case V is continuously differentiable around the point \((\bar{t},\bar{x})\) and \(V_{x}(\bar{t},\bar{x})=e^{-\beta \bar{t}}\). Therefore V =V around \((\bar{t},\bar{x})\), which furthermore yields that \(\psi_{x}(\bar{t},\bar{x})=e^{-\beta \bar{t}}\). Since the test function ψ is continuously differentiable in x, inequality (15) cannot be true and states a contradiction to (14).

Thus we have, when stating (14), that there exists an area around \((\bar{t},\bar{x})\) inside which it is not optimal to consume a lump sum from the wealth. Consequently a strategy C n, for playing a role in the dynamic programming principle for n large enough such that (t n ,x n ) are inside this non-paying area, has the feature that \(X_{t}^{C^{n}}\) can leave \(B_{\delta}(\bar{t},\bar {x})\) through a jump not before leaving the non-paying area continuously.

Therefore τ >t n , which completes the proof and we can conclude that V(t,x) is a viscosity solution to (13). □

Proof of Lemma 3.1

Since V(t,x)≥e βt x for (t,x)∈[0,T]×[0,∞[ (you can always pay out everything and quit by consuming a small constant rate such that X t+<0), we also have V (t,x)≥e βt x and V (t,x)≥e βt x. From e βT x=V(T,x)≥V (T,x) we get V (T,x)=e βT x.

Assume that V (T,x)>e βT x, then there exists η>0 with V (T,x)≥2η+e βT x.

Now choose a sequence (t n ,x n )→(T,x) such that V(t n ,x n )→V (T,x). There is some n 0>0 such that for nn 0 we have

(17)

Let \(C^{n}\in\mathcal{C}(t_{n},x_{n})\) and define \(\tau_{n}=\inf\{t\geq t_{n}\mid X_{t}^{C^{n}}<0\}\wedge T\). Since C n is admissible, we have \(\Delta C_{t_{n}}^{n}\leq x_{n}\), there is no lump-sum payment leading to ruin. Furthermore, by the definition of t n , we have τ n t n →0 if n→∞. Now, fix some ε>0 and choose n large enough such that

Taking a supremum over strategies C n we get \(V(t_{n},x_{n})\leq e^{-\beta t_{n}} x_{n}+\varepsilon\), which contradicts (17) since ε is arbitrary and if (t n ,x n )→(T,x) we have \(e^{-\beta t_{n}} x_{n}\to e^{-\beta T} x\). □

Proof of Theorem 3.2

Assume there is \((\hat{t},\hat{x})\in[0,T]\times\mathbb{R}_{+}\) such that \(u(\hat{t},\hat{x})-v(\hat{t},\hat{x})>0\). We assume, w.l.o.g., \((\hat{t},\hat{x})\in\{(t,x){:}\ t_{j}< t <t_{j-1}, \gamma_{j,i}(t)< x<\gamma_{j,i+1}(t)\}=:A\) with t j <t j−1, γ j,i <γ j,i+1<∞ defined as in Sect. 2 and in Remark 2.2. Note that because of (P5) we do not need to consider {(t,γ j,i (t)): t j <t<t j−1}∪{(t j ,x): γ j,i (t j )<x<γ j,i+1(t j )}. If u(t,γ j,i (t))−v(t,γ j,i (t))>0, then there is h>0 with (t,x+h)∈A with u(t,x+h)−v(t,x+h)>0; the same holds for t=t j .

We further assume that the comparison principle is already shown for the intervals [t l ,t l−1[ with l∈{j−1,…,m}, i.e. u(t,x)≤v(t,x) on \([t_{j-1},T]\times\mathbb{R}_{+}\). Note that Lemma 3.1 yields u(x,T)=v(x,T) for all \(x\in \mathbb{R}_{+}\).

Define v k=kv for k>1. It is easy to check that \(k\tilde{v}\) is still a supersolution with lsc envelope kv. Choose k>1 such that \(u(\hat{t},\hat{x})-v^{k}(\hat{t},\hat{x})>0\). Due to Lemma 2.1, we obtain the following inequality:

It is clear that u(t,x)−v k(t,x)≤0 for \(x\ge\frac{\lambda (0)}{k-1}e^{\beta t_{j-1}}=:\eta\). We can choose k in such a way that ηγ j,i+1(t) for all t∈[t j ,t j−1[ and define

$$A:=\bigl\{(t,x){:}\ t_j< t< t_{j-1},\gamma_{j,i}(t)< x<\gamma_{j,i+1}(t)\bigr\} . $$

Define further

$$M:=\sup_{(t,x)\in A}\bigl\{u(t,x)-v^k(t,x)\bigr\}. $$

From above we know that M<λ(0)<∞ and obtain \(0<u(\hat{t},\hat{x})-v^{k}(\hat{t},\hat{x})\le M\). Since uv k is continuous on A, there is (t ,x )∈A with \(u(t^{*},x^{*})-v^{k}(t^{*},x^{*})>\frac{M}{2}>0\). Further, we let H:={(t,x,s,y): (t,x),(s,y)∈A, yx≥0, ts≥0} and for ξ>0:

In the following, we mean by u(t,γ j,i+1) and v k(s,γ j,i+1) the corresponding lsc envelopes u (t,γ j,i+1) and kv (s,γ j,i+1), respectively. Then it holds that

Note that \((t,x,s,\gamma_{j,i}(s)), (t,\gamma_{j,i+1}(t),s,y), (t_{j},x,s,y)\in\bar{H}\) only if t=s and x=γ j,i (s), t=s and y=γ j,i+1(t) or s=t j which yields

$$f_\xi\bigl(t,x,s,\gamma_{j,i}(s)\bigr),f_\xi \bigl(t,\gamma_{j,i+1}(t),s,y\bigr), f_\xi(t_j,x,s,y)<0. $$

Let M ξ =sup H f ξ . Since f ξ is continuous on H, there is \((t_{\xi},x_{\xi}, s_{\xi},y_{\xi})\in\bar{H}\) such that M ξ =f ξ (t ξ ,x ξ ,s ξ ,y ξ ). Since (t ,x )∈A, it holds that γ j,i (t )<x <γ j,i+1(t ), from which it follows that

We obtain directly: M ξ >0 for

In particular, we have \(\liminf_{\xi\to\infty}M_{\xi}\ge \frac{M}{2}>0\). With the usual techniques, see for example [3], one can show that there is ξ 1>0 such that (t ξ ,x ξ ,s ξ ,y ξ )∈H∂H for ξ>max{ξ 1,ξ 0}.

Note that by definition of (t ξ ,x ξ ,s ξ ,y ξ ) it holds that

$$ f_\xi(t_\xi,x_\xi,s_\xi,x_\xi)+f_\xi(t_\xi,y_\xi,s_\xi,y_\xi )\le2f_\xi(t_\xi,x_\xi,s_\xi,y_\xi). $$
(18)

Choose a sequence ξ n →∞ such that \((t_{\xi_{n}},x_{\xi _{n}},s_{\xi_{n}},y_{\xi_{n}})\to(\bar{t},\bar{x},\bar{s},\bar{y})\).

Let \(z_{n}:= (\xi_{n}(t_{\xi_{n}}-s_{\xi_{n}})+1 ) (\xi_{n}(y_{\xi _{n}}-x_{\xi_{n}}+t_{\xi_{n}}-s_{\xi_{n}})+1 )\). Then using (P2) and (18) we obtain

(19)

Here \(\lim_{n\to\infty}\xi_{n}(y_{\xi_{n}}-x_{\xi_{n}}+t_{\xi _{n}}-s_{\xi_{n}})=\infty\) implies \(\lim_{n\to\infty}\xi _{n}(y_{\xi_{n}}-x_{\xi_{n}})^{3}<0\), which is a contradiction. Thus, we conclude \(\lim_{n\to\infty} (t_{\xi_{n}}-s_{\xi_{n}})= \lim_{n\to\infty} (y_{\xi_{n}}-x_{\xi_{n}})=0\), and consequently \(\bar{x}=\bar{y}\), \(\bar{t}=\bar{s}\). On the other hand, \(\lim_{n\to\infty}\xi_{n}(y_{\xi_{n}}-x_{\xi_{n}}+t_{\xi_{n}}-s_{\xi_{n}})\le \frac{2k}{1+k}\), because otherwise we would again obtain \(\lim_{n\to\infty}\xi_{n}(y_{\xi_{n}}-x_{\xi_{n}})^{3}<0\). Note that \(\bar{x}\) is bounded away from \(\gamma_{j,i+1}(\bar{t})\) and \(\gamma_{j,i}(\bar{t})\).

Define the functions

These functions are continuously differentiable in t and in x. Furthermore, \(u(t,x)e^{\beta s_{\xi_{n}}}-\psi(t,x)\) attains its maximum at \((t_{\xi_{n}},x_{\xi_{n}})\) and \(v^{k}(s,y)e^{\beta t_{\xi_{n}}}-\phi(s,y)\) attains its minimum at \((s_{\xi_{n}},y_{\xi _{n}})\). Thus, \(\psi(t,x)e^{-\beta s_{\xi_{n}}}\) and \(\phi(s,y)e^{-\beta t_{\xi_{n}}}\) are test functions for u(t,x) and v k(s,y), respectively. Due to the arguments after relation (19), we have

$$ \lim_{n\to\infty}\psi_x(t_{\xi_n},x_{\xi_n})= \lim_{n\to\infty}\phi_y(s_{\xi_n},y _{\xi_n})\ge\frac{k+1}{2}>1. $$

Thus, there is \(N\in\mathbb{N}\) such that for n>N it holds that

Therefore, it holds by Definition 3.2 of viscosity sub- and super-solutions that:

Subtracting the above inequalities, rearranging the terms and letting ξ n →∞ yields the following relation:

On the other hand, we know that

which is a contradiction. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Eisenberg, J., Grandits, P. & Thonhauser, S. Optimal Consumption Under Deterministic Income. J Optim Theory Appl 160, 255–279 (2014). https://doi.org/10.1007/s10957-013-0320-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-013-0320-x

Keywords

Navigation