Skip to main content
Log in

On the Infinite-Horizon Optimal Control of Age-Structured Systems

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

Abstract

The paper presents necessary optimality conditions of Pontryagin’s type for infinite-horizon optimal control problems for age-structured systems with state- and control-dependent boundary conditions. Despite the numerous applications of such problems in population dynamics and economics, a “complete” set of optimality conditions is missing in the existing literature, because it is problematic to define in a sound way appropriate transversality conditions for the corresponding adjoint system. The main novelty is that (building on recent results by Aseev and the second author) the adjoint function in the Pontryagin principle is explicitly defined, which avoids the necessity of transversality conditions. The result is applied to several models considered in the literature.

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. We do not mention some publications, where transversality conditions are introduced ad hoc or based on non-sound arguments (see the recent paper [7] for more information). There are some exceptions, out of which we mention [2], where, however, the dynamics and the boundary conditions are local.

  2. The last part of the assumption means that, for every compact sets \(Y\), \(Z\), \(\bar{U} \subseteq U\), \(\bar{V} \subseteq V\), and \(T > 0\), there exists a constant \(L\) such that, for each of the functions listed above (take \(g(t,a,y,z,u,v)\) as a representative)

    $$\begin{aligned} |g(t,a,y_1,z_1,u_1,v_1) - g(t,a,y_2,z_2,u_2,v_2)| \le L (|y_1-y_2| + |z_1-z_2| + |u_1-u_2| + |v_1-v_2|), \end{aligned}$$

    for every \((t,a,y_i,z_i,u_i,v_i) \in [0,T] \times [0,\omega ] \times Y \times Z \times \bar{U} \times \bar{V}\), \(i = 1, 2\).

References

  1. Webb, G.F.: Theory of Nonlinear Age-Dependent Population Dynamics. Marcel Dekker, New York (1985)

    MATH  Google Scholar 

  2. Barucci, E., Gozzi, F.: Technology adoption and accumulation in a vintage capital model. J. Econ. 74(1), 1–30 (2001)

    Article  MATH  Google Scholar 

  3. Boucekkine, R., De la Croix, D., Licandro, O.: Vintage capital growth theory: three breakthroughs. Front. Econ. Glob. 11, 87–116 (2001)

    Article  Google Scholar 

  4. Feichtinger, G., Tragler, G., Veliov, V.M.: Optimality conditions for age-structured control systems. J. Math. Anal. Appl. 288, 47–68 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  5. Aseev, S.M., Kryazhimskii, A.V.: The pontryagin maximum principle and transversality conditions for a class of optimal control problems with infinite time horizons. SIAM J. Control Optim. 43, 1094–1119 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  6. Aseev, S.M., Veliov, V.M.: Needle variations in infinite-horizon optimal control. Contemp. Math. 619, 1–17 (2014)

    Article  MathSciNet  Google Scholar 

  7. Krastanov, M.I., Ribarska, N.K., Tsachev, Ts.Y.: A pontryagin maximum principle for infinite-dimensional problems. SIAM J. Control Optim. 49(5), 2155–2182 (2011)

  8. Anita, S.: Analysis and Control of Age-Dependent Population Dynamics. Kluwer, Dordrecht (2000)

    Book  MATH  Google Scholar 

  9. Feichtinger, G., Hartl, R.F., Kort, P.M., Veliov, V.M.: Anticipation effects of technological progress on capital accumulation: a vintage capital approach. J. Econ. Theory 126(1), 143–164 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  10. Prskawetz, A., Tsachev, Ts., Veliov, V.M.: Optimal education in an age-structured model under changing labor demand and supply. Macroecon. Dyn. 16, 159–183 (2012)

  11. Prskawetz, A., Veliov, V.M.: Age-specific dynamic labor demand and human capital investment. J. Econ. Dyn. Control 31(12), 3741–3777 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  12. Aseev, S.M., Veliov, V.M.: Maximum principle for infinite-horizon optimal control problems with dominating discount. Dyn. Contin. Discrete Impuls. Syst. B 19, 43–63 (2012)

    MATH  MathSciNet  Google Scholar 

  13. Carlson, D.A., Haurie, A.B., Leizarowitz, A.: Infinite Horizon Optimal Control: Deterministic and Stochastic Systems. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  14. Faggian, S.: Applications of dynamic programming to economic problems with vintage capital. Dyn. Contin. Discrete Impuls. Syst. A 15, 527–553 (2008)

    MATH  MathSciNet  Google Scholar 

  15. Faggian, S., Gozzi, F.: Dynamic programming for infinite horizon boundary control problems of PDE’s with age structure. J. Math. Econ. 46(4), 416–437 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  16. Iannelli, M.: Mathematical Theory of Age-Structured Population Dynamics. Giardini Editori, Pisa (1994)

    Google Scholar 

  17. Brokate, M.: Pontryagin’s principle for control problems in age-dependent population dynamics. J. Math. Biol. 23, 75–101 (1985)

    Article  MATH  MathSciNet  Google Scholar 

  18. Gripenberg, G., Londen, S.O., Staffans, O.: Volterra Integral and Functional Equations. Cambridge University Press, Cambridge (1990)

    Book  MATH  Google Scholar 

  19. Veliov, V.M.: Optimal control of heterogeneous systems: basic theory. J. Math. Anal. Appl. 346, 227–242 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  20. Feichtinger, G., Veliov, V.M.: On a distributed control problem arising in dynamic optimization of a fixed-size population. SIAM J. Optim. 18, 980–1003 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  21. Simon, C., Skritek, B., Veliov, V.M.: Optimal immigration age-patterns in populations of fixed size. J. Math. Anal. Appl. 405(1), 71–89 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  22. Krastev, V.Y.: Arrow-type sufficient conditions for optimality of age-structured control problems. Cent. Eur. J. Math. 11(6), 1094–1111 (2013)

    MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

This research was supported by the Austrian Science Foundation (FWF) under Grant No. I 476-N13.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to V. M. Veliov.

Appendix

Appendix

Below we give the proofs of Lemmas 4.1, 5.15.4.

Proof of Lemma 4.1

For a fixed \(t > 0\), the integrand in (27) can be estimated using (28), (A2) and (A3) as follows:

$$\begin{aligned} |\psi (\theta )\, R(\theta ,t)|&= \left| \int _0^{\omega } [g_yX(\theta +s,s) \varPhi (\theta ) + g_z(\theta ,s)] \mathrm{d}s\, R(\theta ,t) \right| \nonumber \\&\le \int _0^{\omega } \left[ \rho (\theta +s) \,\vert X(\theta +s,s) \vert \, \vert \varPhi (\theta ) \vert + \rho (\theta ) \right] \mathrm{d}s \, \vert R(\theta ,t) \vert \le \lambda (\theta , t). \end{aligned}$$
(45)

The first claim of Lemma 4.1 follows from the integrability of \(\lambda (\cdot ,t)\). The local boundedness of \(\hat{\zeta }\) follows from the local boundedness of \(\psi \) and \(\int _t^\infty \lambda (\theta ,t) \mathrm{d}\theta \).

Now consider \(\hat{\xi }\). Denote \(\xi _1(t,a) := \int _a^{\omega } g_y X(t-a+s,s) \mathrm{d}s X^{-1}(t,a)\), which is the first term in (26), including the multiplication with \(X^{-1}\). First, we show that \(\xi _1\) is Lipschitz continuous along the characteristic lines \(t-a = \mathrm{const}\):

$$\begin{aligned}&\hat{\xi }_1(t+\varepsilon ,a+\varepsilon ) - \hat{\xi }_1(t,a) = \dots \\&\quad \int _{a+\varepsilon }^{\omega } g_yX(t-a+s,s) \mathrm{d}s \, X^{-1}(t+\varepsilon ,a+\varepsilon )\\&\qquad -\int _a^{\omega } g_y X(t-a+s,s)\mathrm{d}s \, X^{-1}(t,a) = \dots \\&\quad \int _{a+\varepsilon }^{\omega } \! g_yX(t-a+s,s) \mathrm d s \left[ X^{-1}(t+\varepsilon ,a+\varepsilon ) \! - \! X^{-1}(t,a) \right] \\&\qquad + \! \int _a^{a+\varepsilon } \! \! g_yX(t-a+s,s) \mathrm d s \, X^{-1}(t,a). \end{aligned}$$

The Lipschitz continuity follows from the Lipschitz continuity of \(X^{-1}\) along the characteristic lines and the local boundedness of \(g_y X\) and \(X^{-1}\). Applying the differentiation \(\fancyscript{D}\) to \(\xi _1\) and using the expression (14) for \(\fancyscript{D}X^{-1}\), we obtain

$$\begin{aligned} - \fancyscript{D}\xi _1(t,a)&= \int _a^{\omega } g_y X(t-a+s,s) \mathrm{d}s\, (X^{-1}(t,a) F(t,a)) + g_y(t,a) \\&= \xi _1(t,a) F(t,a) + g_y(t,a). \end{aligned}$$

The proof that the second term, \(\xi _2\), in the definition of \(\hat{\xi }\) in (22) is Lipschitz continuous along the characteristic lines \(t-a = \mathrm{const}\) and satisfies the equation \(-\fancyscript{D}\xi _2(t,a) = \xi _2(t,a) F(t,a) + \hat{\zeta }(t) H(t,a)\) is similar and therefore omitted. Then, \(\hat{\xi }= \xi _1 + \xi _2\) belongs to the space \(\fancyscript{A}(D)\) and satisfies (22).

The definition (27) of \(\hat{\zeta }\) has the form (9), with \(T = \infty \). We know that \(K(t,s) = 0\) for \(t \not \in [s,s+\omega ]\). Moreover, from (45) and (A3) we obtain that the integral in (9) with \(T = \infty \) is locally bounded in \(t\). Then, we may apply the implication in the end of Sect. 3.1, which claims that

$$\begin{aligned} \hat{\zeta }(t) = \psi (t) + \int _t^\infty \hat{\zeta }(\theta ) \, K(\theta ,t) \mathrm{d}\theta = \psi (t) + \int _t^{t+\omega } \hat{\zeta }(\theta ) \, K(\theta ,t) \mathrm{d}\theta . \end{aligned}$$

Inserting the expressions (20) and (28) for \(\psi \) and \(K\), respectively, we obtain that

$$\begin{aligned} \hat{\zeta }(t) \!&= \! \left[ \int _0^\omega g_y X(t+x,x) \mathrm{d}x \!+\! \int _t^{t+\omega } \hat{\zeta }(\theta ) \, HX(\theta ,\theta -t) \mathrm{d}\theta \right] \varPhi (t) \!+\! \int _0^{\omega } g_z(t,a) \mathrm{d}a \\&= \hat{\xi }(t,0) \, \varPhi (t) + \int _0^{\omega } g_z(t,a), \end{aligned}$$

that is, (23) is fulfilled by \((\hat{\xi },\hat{\zeta })\). The proof is complete. \(\square \)

Proof of Lemma 5.1

We represent

$$\begin{aligned} \Delta _\tau (u_\alpha )&= \int _{\tau -\alpha }^\tau \int _0^\omega [(g(t,a,y_{\alpha },z_{\alpha },u_{\alpha }) - g(t,a,z_{\alpha },u_{\alpha }))\nonumber \\&+ (g(t,a,z_{\alpha },u_{\alpha }) - g(t,a,u_{\alpha })) \\&+ \, (g(t,a,u_{\alpha })- g(t,a))] \mathrm{d}a \mathrm{d}t \, =: \, I_1 + I_2 + I_3. \end{aligned}$$

We remind that \(\Vert \Delta y\Vert _{L_\infty (D_\tau )} + \Vert \Delta z\Vert _{L_\infty ([0,\tau ])} \le c_0 \,\alpha \) (see (38)). Moreover, \(u_\alpha (t,a) \not = \hat{u}(t,a)\) only on the set \(B(\tau ,b;\alpha )\) (where \(u_\alpha (t,a) = u\)), and \(\tilde{y}_\alpha (t,a) = \hat{y}(t,a)\) except on a set of measure \(2 \alpha ^2\). Using, in addition, that \(g\) is Lipschitz continuous with respect to \((y,z)\) in the domain where \((y_\alpha ,z_\alpha )\) and \((\hat{y}, \hat{z})\) take values for \((t,a) \in D_\tau \), we apparently have \(|I_1| = o(\alpha ^2)\). For \(I_3\) we have

$$\begin{aligned} I_3 = \iint _{B(\tau ,b;\alpha )} (g(t,a,u)- g(t,a)) \mathrm{d}a \mathrm{d}t = \alpha ^2 (g(\tau ,b,u)- g(\tau ,b)) + o(\alpha ^2) \end{aligned}$$

due to the Lebesgue property of \((\tau ,b)\); see the beginning of “Part 2 for \(u\)” in Sect. 5. Finally,

$$\begin{aligned} I_2 =&\int _{\tau -\alpha }^\tau \left( \int _0^\omega g_z(t,a,u_\alpha (t,a))\, \Delta z(t) \mathrm{d}a + o(\alpha ;t) \right) \mathrm{d}t \\ =&\int _{\tau -\alpha }^\tau \int _0^\omega g_z(t,a)\, \Delta z(t) \mathrm{d}a \mathrm{d}t + o(\alpha ^2), \end{aligned}$$

where here and below \(o(\alpha ;t)/\alpha \) converges to zero uniformly in \(t\) in the interval of interest (in this case, it is \([0,\tau ]\)). For \(\Delta z\) we have

$$\begin{aligned} \Delta z(t)&= \int _0^\omega \left( H^\sharp (t,a,y_\alpha (t,a),u_\alpha (t,a)) - H^\sharp (t,a)\right) \mathrm{d}a \nonumber \\&= \int _0^\omega \left( H^\sharp (t,a,u_\alpha (t,a)) - H^\sharp (t,a)\right) \mathrm{d}a + o(\alpha )\end{aligned}$$
(46)
$$\begin{aligned}&= \int _{b-\alpha }^b \left( H^\sharp (t,a,u) - H^\sharp (t,a)\right) \mathrm{d}a + o(\alpha ). \end{aligned}$$
(47)

Inserting this in the expression for \(I_2\) and changing the order of integration, we obtain

$$\begin{aligned} I_2&= \iint _{B(\tau ,b;\alpha )} \int _0^\omega g_z(t,s) \mathrm{d}s \,(H^\sharp (t,a,u)- H^\sharp (t,a)) \mathrm{d}a \mathrm{d}t + o(\alpha ^2) \\&= \alpha ^2 \int _0^\omega g_z(\tau ,s) \mathrm{d}s \,(H^\sharp (\tau ,b,u)- H^\sharp (\tau ,b)) + o(\alpha ^2), \end{aligned}$$

where for the last equality we use the Lebesgue property of \((\tau ,b)\); see the beginning of “Part 2 for \(u\)” in Sect. 5.

Summing the obtained expressions for \(I_1\), \(I_2\), and \(I_3\) we obtain the claim of the lemma. \(\square \)

Proof of Lemma 5.2

We remind the inequalities \(2 \alpha < \tau \), \(2 \alpha < b\), and \(2 \alpha < \omega - b\) posed for \(\alpha \) in “Part 2 for \(u\)” in Sect. 5. Observe that \(\Delta y(\tau ,a) = 0\) for all \(a\) except for \(a \in [0,\alpha ]\cup [b-\alpha ,b+\alpha ]\). Therefore, we consider the integral in the formulation of the lemma separately on these two intervals.

Beginning with \([0,\alpha ]\), we represent

$$\begin{aligned}&\int _0^\alpha \hat{\xi }(\tau ,a) \, \Delta y(\tau ,a) \mathrm{d}a \\&\quad = \! \int _0^{\alpha } \left[ \hat{\xi }(\tau -a,0) + \int _0^a \fancyscript{D}\hat{\xi }(\tau -a+s,s) \mathrm{d}s\right] \\&\qquad \times \Bigg [\Delta y(\tau -a,0) \\&\qquad + \int _0^a \fancyscript{D}\Delta y(\tau -a+s,s)\mathrm{d}s\Bigg ] \mathrm{d}a. \end{aligned}$$

We remind that \(\Vert \Delta y\Vert _{L_\infty (D_\tau )} \le c_0 \,\alpha \). For any \(t \ge 0\) and \(s \in [0,\alpha ]\), we have \(u_\alpha (t,s) = \hat{u}(t,s)\), hence \(|\fancyscript{D}\Delta y(t,s)| = |F(t,s)\, \Delta y(t,s)| \le c \,\alpha \). Moreover, due to Lemma 4.1 and the local boundedness of \(F\), \(\hat{\xi }\), \(\hat{\zeta }\), \(H\), and \(g_y\), we have \(\Vert \fancyscript{D}\hat{\xi }\Vert _{L_\infty (D_\tau )} \le c\). Hence,

$$\begin{aligned} \int _0^\alpha \hat{\xi }(\tau ,a) \, \Delta y(\tau ,a) \mathrm{d}a = \int _0^{\alpha } \hat{\xi }(\tau -a,0)\, \Delta y(\tau -a,0) \mathrm{d}a + o(\alpha ^2). \end{aligned}$$

Since \(\Delta y(\tau -a,0) = \varPhi (\tau -a) \, \Delta z(\tau -a)\), using representation (47) and changing the integration variable, we obtain that

$$\begin{aligned} \int _0^\alpha \hat{\xi }(\tau ,a) \, \Delta y(\tau ,a) \mathrm{d}a&= \iint _{B(\tau ,b;\alpha )} \hat{\xi }(t,0) \,\varPhi (t) \, (H^\sharp (t,a,u)\nonumber \\&-H^\sharp (t,a)) \mathrm{d}a \mathrm{d}t + o(\alpha ^2). \end{aligned}$$

Due to the Lebesgue point property of \((\tau ,b)\), the expression in the right-hand side equals the second term in the right-hand side in the assertion of the lemma.

Now, we consider \(E := \int _{b-\alpha }^{b+\alpha } \hat{\xi }(\tau ,a)\, \Delta y(\tau ,a) \mathrm{d}a\). For \(a\) in the interval of integration

$$\begin{aligned} \Delta y(\tau ,a) = \Delta y(\tau -\alpha ,a-\alpha ) + \int _0^\alpha \fancyscript{D}\Delta y(\tau -x,a-x) \mathrm{d}x, \end{aligned}$$

and the first term in the right-hand side is zero (due to \(a - \alpha \ge 0\)). Then, \(E\) is equal to

$$\begin{aligned} \!&\! \int _{b-\alpha }^{b+\alpha } \! \int _0^\alpha \! \! \hat{\xi }(\tau ,a) \! \left[ F^\sharp (\tau -x,a-x,y_\alpha (\tau -x,a-x),u_\alpha (\tau -x,a-x))\right. \\&\quad \left. - F^\sharp (\tau -x,a-x)\right] \! \mathrm d x \mathrm{d}a \\&= \int _{\tau -\alpha }^{\tau } \int _{b-\tau +t-\alpha }^{b-\tau +t+\alpha } \hat{\xi }(\tau ,\tau -t+s) \left[ F^\sharp (t,s,u_\alpha (t,s)) - F^\sharp (t,s)\right] \mathrm{d}s \mathrm{d}t + o(\alpha ^2), \end{aligned}$$

where we passed to the new variables \(t = \tau - x\) and \(s = a-x\) and used that \(\Vert \Delta y\Vert _{L_\infty (D_\tau )} \le c_0 \,\alpha \). Note that, if \(s < b - \alpha \) or \(s > b\), then the last integrand is zero, since \(u_\alpha (t,s)= \hat{u}(t,s)\). Otherwise \(u_\alpha (t,s) = u\). Then

$$\begin{aligned} E = \iint _{B(\tau ,b;\alpha )} \hat{\xi }(\tau ,\tau -t+s) \, \left[ F^\sharp (t,s,u) - F^\sharp (t,s)\right] \mathrm{d}s \mathrm{d}t + o(\alpha ^2). \end{aligned}$$

Using the Lebesgue property of \((\tau ,b)\) (see the beginning of “Part 2 for \(u\)” in Sect. 5), we obtain the first term in the right-hand side in the assertion of the lemma. \(\square \)

Proof of Lemma 5.3

By the definition of \(v_\alpha \) we have \(|v_\alpha (t) - \hat{v}(t)| \le c \,\alpha \). According to Proposition 3.2, \(\Vert \Delta y(t,\cdot ) \Vert _{L_1([0,\omega ])} \le c \alpha ^2\). Moreover,

$$\begin{aligned} \Delta z(t) = \int _0^{\omega } \alpha H^\sharp _v(t,a)(v-\hat{v}(t)) \mathrm{d}a + o(\alpha ;t). \end{aligned}$$

Then, we obtain

$$\begin{aligned} \Delta _{\tau }(v_{\alpha })&= \int _{\tau -\alpha }^{\tau } \int _0^{\omega } [g(t,a,z_{\alpha }(t),v_\alpha (t)) - g(t,a,v_\alpha (t)) + g(t,a,v_\alpha (t)) \\&- g(t,a) ] \mathrm{d}a \mathrm{d}t + o(\alpha ^2)\\&= \int _{\tau -\alpha }^{\tau } \int _0^{\omega } [g_z(t,a) \Delta z(t) + g(t,a,v_\alpha (t)) - g(t,a)] \mathrm{d}a \mathrm{d}t + o(\alpha ^2) \\&= \int _{\tau -\alpha }^{\tau } \left\{ \int _0^{\omega } g_z(t,s) \mathrm{d}s \int _0^{\omega } H^\sharp _v(t,a) \mathrm{d}a + \int _0^{\omega } g_v(t,a) \mathrm{d}a \right\} \nonumber \\&\times \,\alpha (v-\hat{v}(t)) \mathrm{d}t+ o(\alpha ^2), \end{aligned}$$

which proves the claim of the lemma due to the Lebesgue property of \(\tau \) (see the beginning of “Part 2 for \(v\)” in Sect. 5). \(\square \)

Proof of Lemma 5.4

The proof uses similar arguments as that of Lemma 5.2, and therefore is somewhat shortened. From (3), we have

$$\begin{aligned} \Delta y(t,0) = \varPhi (t,\hat{v}) \Delta z(t) + \alpha \varPhi ^\sharp _v(t)(v-\hat{v}(t)) + o(\alpha ;t). \end{aligned}$$

For \(t=\tau \) and \(a < \alpha \),

$$\begin{aligned} \Delta y(\tau ,a) = \Delta y(\tau -a,0) + \int _{-a}^0 \fancyscript{D}\Delta y(\tau +s,a+s) \mathrm{d}s = \Delta y(\tau -a,0) + o(\alpha ;a), \end{aligned}$$

while for \(a > \alpha \),

$$\begin{aligned} \Delta y(\tau ,a) =&\int _{-\alpha }^0 [F^\sharp (\tau +s,a+s,y_{\alpha },v_{\alpha }) - F^\sharp (\tau +s,a+s)] \mathrm{d}s\\ =&\int _{-\alpha }^0 \alpha F^\sharp _v(\tau +s,a+s)(v-\hat{v}(\tau +s)) \mathrm{d}s + o(\alpha ;a). \end{aligned}$$

This is obtained by splitting the difference in two parts: one for the difference between \(y_{\alpha }\) and \(\hat{y}\), and one for \(v_{\alpha }\) and \(\hat{v}\). Due to Proposition 3.2, the difference with respect to \(y\) can be estimated by \(o(\alpha ;a)\).

Consider the integral \(\int _0^{\omega } \hat{\xi }(\tau ,a) \Delta y(\tau ,a) \mathrm{d}a\) on \([0,\alpha [\). Because \(\vert \Delta y(\tau ,a)\vert \le c \alpha \), and the above representation,

$$\begin{aligned} \int _0^{\alpha } \hat{\xi }(\tau ,a) \Delta y(\tau ,a) \mathrm{d}a = \int _0^{\alpha } \hat{\xi }(\tau -a,0) \Delta y(\tau -a,0) \mathrm{d}a + o(\alpha ^2). \end{aligned}$$

Since \(\tau \) is a Lebesgue point, using the representation for \(\Delta y(\tau ,0)\) and for \(\Delta z(\tau )\) from the proof of the Lemma 5.3 we obtain the expression

$$\begin{aligned} \alpha ^2 \hat{\xi }(\tau ,0) \varPhi (\tau ) \int _0^{\omega } H^\sharp _v(\tau ,a) (v-\hat{v}(\tau )) \mathrm{d}a + \alpha ^2 \hat{\xi }(\tau ,0) \varPhi ^\sharp _v(\tau ) (v-\hat{v}(\tau )) + o(\alpha ^2). \end{aligned}$$

With the representation for \(\Delta y(\tau ,a)\) for \(a > \alpha \), and the absolute continuity of \(\hat{\xi }\) along the characteristic lines, we obtain

$$\begin{aligned} \int _{\alpha }^{\omega } \hat{\xi }(\tau ,a) \Delta y(\tau ,a) \mathrm{d}a =&\int _{\alpha }^{\omega } \int _{\tau -\alpha }^{\tau } \hat{\xi }(t,a-\tau +t) \alpha F^\sharp _v(t,a-\tau +t)(v-\hat{v}(t)) \mathrm{d}t \mathrm{d}a \\&+ o(\alpha ^2) \\ =&\; \alpha \int _{\alpha }^{\omega -\alpha } \int _{\tau -\alpha }^{\tau } \hat{\xi }(t,a) \,F^\sharp _v(t,a) (v-\hat{v}(t)) \mathrm{d}t \mathrm{d}a + o(\alpha ^2)\\ =&\;\alpha \int _{\tau -\alpha }^{\tau } \int _0^{\omega } \hat{\xi }(t,a)\, F^\sharp _v(t,a) (v-\hat{v}(t)) \mathrm{d}a \mathrm{d}t + o(\alpha ^2). \end{aligned}$$

Since \(\tau \) is a Lebesgue point, this implies the claim of the lemma. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Skritek, B., Veliov, V.M. On the Infinite-Horizon Optimal Control of Age-Structured Systems. J Optim Theory Appl 167, 243–271 (2015). https://doi.org/10.1007/s10957-014-0680-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-014-0680-x

Keywords

Mathematics Subject Classification

Navigation