Skip to main content
Log in

Optimal Control for a Linear Quadratic Problem with a Stochastic Time Scale

  • STOCHASTIC SYSTEMS
  • Published:
Automation and Remote Control Aims and scope Submit manuscript

Abstract

We consider a linear-quadratic control problem where a time parameter evolves according to a stochastic time scale. The stochastic time scale is defined via a stochastic process with continuously differentiable paths. We obtain an optimal infinite-time control law under criteria similar to the long-run averages. Some examples of stochastic time scales from various applications have been examined.

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.

Institutional subscriptions

Similar content being viewed by others

REFERENCES

  1. Veraart, A.E.D. and Winkel, M., Time Change/Encyclopedia of Quantitative Finance, New York: Wiley, 2010, pp. 1878–1881.

    Google Scholar 

  2. Kobayashi, K., Stochastic calculus for a time-changed semimartingale and the associated stochastic differential equations, J. Theor. Probab., 2011, vol. 24, no. 3, pp. 789–820.

    Article  MathSciNet  Google Scholar 

  3. Borovkova, S. and Schmeck, M.D., Electricity price modeling with stochastic time change, Energy Econ., 2017, vol. 63, pp. 51–65.

    Article  Google Scholar 

  4. Capra, W.B. and Muller, H.G., An accelerated-time model for response curves, JASA, 1997, vol. 92, no. 437, pp. 72–83.

    Article  MathSciNet  Google Scholar 

  5. Heath, D. and Platen, E., Understanding the implied volatility surface for options on a diversified index, Asia-Pac. Financ. Mark., 2004, vol. 11, no. 1, pp. 55–77.

    Article  Google Scholar 

  6. Ray, D. and Bossaerts, P., Positive temporal dependence of the biological clock implies hyperbolic discounting, Front. Neurosci., 2011, vol. 5, p. 2.

    Article  Google Scholar 

  7. Uneyama, T., Miyaguchi, T., and Akimoto, T., Relaxation functions of the Ornstein-Uhlenbeck process with fluctuating diffusivity, Phys. Rev. E, 2019, vol. 99, no. 3, p. 032127.

    Article  Google Scholar 

  8. Ye, Z.S. and Xie, M., Stochastic modelling and analysis of degradation for highly reliable products, Appl. Stochastic Models Bus. Ind., 2015, vol. 31, no. 1, pp. 16–32.

    Article  MathSciNet  Google Scholar 

  9. Poulsen, D.R., Davis, J.M., and Gravagne, I.A., Optimal control on stochastic time scales, IFAC-PapersOnLine, 2017, vol. 50, no. 1, pp. 14861–14866.

    Article  Google Scholar 

  10. Lamperski, A. and Cowan, N.J., Time-changed linear quadratic regulators, 2013 Eur. Control Conf. (ECC), IEEE, 2013, pp. 198–203.

  11. Tang, S., General linear quadratic optimal stochastic control problems with random coefficients: linear stochastic Hamilton systems and backward stochastic Riccati equations, SIAM J. Control Optim., 2003, vol. 42, no. 1, pp. 53–75.

    Article  MathSciNet  Google Scholar 

  12. Liptser, R.S. and Shiryaev, A.N., Statistics of Random Processes: I. General Theory, Berlin: Springer, 2001.

    Book  Google Scholar 

  13. Øksendal, B., When is a stochastic integral a time change of a diffusion?, J. Theor. Probab., 1990, vol. 3, no. 2, pp. 207–226.

    Article  MathSciNet  Google Scholar 

  14. Shaliastovich, I. and Tauchen, G., Pricing Implications of Stochastic Volatility, Business Cycle Time Change and Non-Gaussianity. Working Paper, Duke Univ., 2005.

  15. Howison, S. and Lamper, D., Trading Volume in Models of Financial Derivatives, Appl. Math. Financ., 2001, vol. 8, no. 2, pp. 119–135.

    Article  Google Scholar 

  16. Lanoiselee, Y., Moutal, N., and Grebenkov, D.S., Diffusion-limited reactions in dynamic heterogeneous media, Nature Commun., 2018, vol. 9, no. 1, pp. 1–16.

    Article  Google Scholar 

  17. Liptser, R.S. and Shiryaev, A.N., Statistics of Random Processes: II. Applications, Berlin: Springer, 2001.

    Book  Google Scholar 

  18. Davis, M.H.A., Linear Estimation and Stochastic Control, London: Chapman and Hall, 1977. Translated under the title: Lineinoe otsenivanie i stokhasticheskoe upravlenie, Moscow: Nauka, 1984.

    MATH  Google Scholar 

  19. Kwakernaak, H. and Sivan, R., Linear Optimal Control Systems, New York–London–Sydney–Toronto: Wiley-Interscience, 1972. Translated under the title: Lineinye optimal’nye sistemy upravleniya, Moscow: Nauka, 1977.

    MATH  Google Scholar 

  20. Palamarchuk, E.S., On invariance of optimal control of linear economic system under simultaneous scaling of its parameters, Sb. Tsentr. Ekon. Mat. Inst., 2018, no. 2. https://cemi.jes.su/s111111110000084-5-1.

  21. Chen, H., A Brownian model of stochastic processing networks, in Stochastic Modeling and Optimization. With Applications in Queues, Finance, and Supply Chains, Yao, D.D, Zhang, H., and Zhou, X.Y., Eds., New York: Springer, 2003, pp. 171–192.

  22. Øksendal, B., Stochastic Differential Equations. An Introduction with Applications, Heidelberg–New York: Springer-Verlag, 2000. Translated under the title: Stokhasticheskie differentsial’nye uravneniya. Vvedenie v teoriyu i prilozheniya, Moscow: Mir–AST, 2003.

    MATH  Google Scholar 

  23. Jeanblanc, M., Yor, M., and Chesney, M., Mathematical Methods for Financial Markets, New York: Springer, 2009.

    Book  Google Scholar 

  24. Loeve, M., Probability Theory II , New York: Springer, 1978.

    Book  Google Scholar 

  25. Freris, N.M., Borkar, V.S., and Kumar, P.R., A model-based approach to clock synchronization, Proc. 48th IEEE Conf. Decis. Control (CDC), New York: IEEE, 2009, pp. 5744–5749.

  26. Xia, W., Pricing exotic power options with a Brownian-time-changed variance gamma process, Commun. Math. Financ., 2017, vol. 6, no. 1, pp. 21–60.

    Google Scholar 

  27. Csaki, E., Iterated logarithm laws for the square integral of a Wiener process, in The First Pannonian Symposium on Mathematical Statistics, Revesz, P., Schmetterer, L., and Zolotarev, V.M., Eds., New York: Springer, 1981, pp. 42–53.

  28. Dufresne, D., The distribution of a perpetuity, with applications to risk theory and pension funding, Scand. Actuarial J., 1990, vol. 1990, no. 1, pp. 39–79.

    Article  MathSciNet  Google Scholar 

  29. Lanoiselee, Y. and Grebenkov, D.S., A model of non-Gaussian diffusion in heterogeneous media, J. Phys. A: Math. Theor., 2018, vol. 51, no. 14, p. 145602.

    Article  MathSciNet  Google Scholar 

  30. Ditlevsen, S. and Lansky, P., Estimation of the input parameters in the Feller neuronal model, Phys. Rev. E, 2006, vol. 73, no. 6, p. 061910.

    Article  MathSciNet  Google Scholar 

  31. Belkina, T.A. and Palamarchuk, E.S., On stochastic optimality for a linear controller with attenuating disturbances, Autom. Remote Control, 2013, vol. 74, no. 4, pp. 628–641.

    Article  MathSciNet  Google Scholar 

  32. Lapeyre, B., A priori bound for the supremum of solutions of stable stochastic differential equations, Stochastics, 1989, vol. 28, no. 3, pp. 145–160.

    MathSciNet  MATH  Google Scholar 

  33. Wang, J., A law of the iterated logarithm for stochastic integrals, Stochastic Process. Appl., 1993, vol. 47, no. 2, pp. 215–228.

    Article  MathSciNet  Google Scholar 

  34. Belkina, T.A., Kabanov, Yu.M., and Presman, E.L., On a stochastic optimality of the feedback control in the LQG-problem, Theory Probab. Appl., 2004, vol. 48, no. 4, pp. 592–603.

    Article  MathSciNet  Google Scholar 

Download references

Funding

This work was prepared within the framework of the HSE University Basic Research Program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. S. Palamarchuk.

Additional information

Translated by V. Potapchouck

APPENDIX

Proof of the Lemma. First, the assertions in the Lemma are proved for the case of Eq. (10) with \(A-BR^{-1}B^\mathrm {T}\Pi =-\kappa I\); \(\kappa >0 \) is a constant, and \(I \) is the identity matrix. Then it is shown that the underlying properties of the process \(X^{*}_t\), \(t\to \infty \), in Eq. (10) do not change for an arbitrary exponentially stable matrix \( A-BR^{-1}B^\mathrm {T}\Pi \). Consider the process \( X^{*}_t=\hat X_t\), \(t\geqslant 0 \), with the dynamics

$$ d\hat X_t=-\kappa \alpha _tI \hat X_t+\sqrt {\alpha _t}GdW_t , \quad \hat X_0=0 . $$
Denoting the \(i \)th component of the process \(\hat X_t \), \(i=1,\dots ,n \), by \(\hat X_{it} \), one can readily obtain the representation \(\hat X_{it}=c_{i}M_{it}\left (\left \langle M_{it}\right \rangle +1\right )^{-1/2} \), where the martingale
$$ {M_{it}={c_{i}}^{-1}\int \limits _0^t \sqrt {\alpha _s}\exp {\left \{\int \limits _0^s \kappa \alpha _v dv\right \}}\left (\thinspace \sum _{j=1}^{d} G_{ij}dW_{js}\right )} $$
has the quadratic variation \(\langle M_{it}\rangle =\exp {\big \{2\kappa \int \nolimits _0^t \alpha _v dv\big \}}-1 \); further, \(c_{i}=\big ((2\kappa )^{-1}\sum _{j=1}^{d}G^2_{ij}\big )^{1/2}\), where the \(G_{ij} \) are the entries of the matrix \(G \) (\(i=1,\dots ,n \), \(j=1,\dots ,d \)); \(W_{jt} \) is the \(j \)th component of the Wiener process \(W_t \) (\(j=1,\dots ,d \)). It was established in [32, Lemma 2.3] that
$$ {\Big \|M_{it}\big (\langle M_{it}\rangle +1\big )^{-1/2}\Big \|\leqslant N(\omega )\sqrt {\ln \ln \big (\langle M_{it}\rangle +e^{e}\big )}},$$
where \(N(\omega )\geqslant 0\) is a.s. a finite random variable. Therefore, for the process \(\|\hat X_t\|^2\) with probability \(1 \) one has the relation
$$ \|\hat X_t\|^2\leqslant cN^2(\omega )\ln \left (\int \limits _0^t \alpha _v dv+e\right ) .$$
(A.1)
Then it follows from the estimate (A.1) and the Jensen inequality that
$$ \mathrm {E}\|\hat X_t\|^2\leqslant \tilde c\ln \left (\mathrm {E}\int \limits _0^t\alpha _v dv+e\right ) , $$
(A.2)
where \(c\) and \(\tilde c \) in (A.1) and (A.2) denote some positive constants whose particular values are inessential and can vary from formula to formula. It can readily be noticed that relation (13) is an obvious consequence of the above representation for the components of \( \hat X_t\) and of the law of iterated logarithm for martingales; see, e.g., [33]. Further, we introduce the process \(Z_t=X^{*}_t-\hat X_t \) with the dynamic equation
$$ dZ_t=\alpha _t\left (A-BR^{-1}B^\mathrm {T}\Pi \right ) Z_tdt+\alpha _t\left (A-BR^{-1}B^\mathrm {T}\Pi +\kappa I\right )\hat X_t , \quad Z_0=x,$$
which has the solution \(Z_t=\Phi (t,0)x+\int \nolimits _0^t\Phi (t,s) \alpha _s(A-BR^{-1}B^{\mathrm {T}}\Pi +\kappa I)\hat X_s ds \). In view of the upper bound for \(\|\Phi (t,s)\| \) (see the comment on (11)) and the Cauchy–Schwarz inequality for the process \(Z_t \), we have the estimate
$$ \begin {aligned} \|Z_t\|^2&\leqslant 2\kappa _0^2{ \exp {\left \{-2\kappa \int \limits _0^t \alpha _v dv\right \}}}\|x\|^2 \\ &\qquad {}+c\thinspace { \exp {\left \{-\kappa \int \limits _0^t \alpha _v dv\right \}}}\int \limits _0^t \alpha _s{\thinspace \exp {\left \{\kappa \int \limits _0^s \alpha _v dv\right \}}\|\hat X_s\|^2 ds}. \end {aligned} $$
(A.3)
Applying (A.1) to (A.3) gives the relation
$$ \|Z_t\|^2\leqslant 2\kappa _0^2{\thinspace \exp {\left \{-2\kappa \int \limits _0^t \alpha _v dv\right \}}}\|x\|^2+cN^2(\omega ) \ln \left (\int \limits _0^t \alpha _v dv+e\right ).$$
Taking the expectation on both sides in this relation combined with the Jensen inequality leads to the estimate \(\mathrm {E}\|Z_t\|^2\leqslant \tilde c+\tilde c \ln \left (\mathrm {E}\int \nolimits _0^t \alpha _v dv+e\right ) \); this then implies (12) in the lemma being proved. Inequality (13) is also easy to obtain in a well-known way; see, e.g., the proof in [31, Theorem 2], if one notices that \(h_t=\ln \left (\int \nolimits _0^t \alpha _v dv+e\right )\) is a nondecreasing function. Then dividing (A.3) by \(h_t \) in the subsequent estimation of the integral on the right-hand side (while using the result for \(\|\hat X_t\|^2\)) gives a bounded limit. The proof of the lemma is complete.\(\quad \blacksquare \)

Proof of the Theorem. For \(U\in \mathcal {U}\), we write a representation for the difference of cost functionals,

$$ J^{(\alpha )}_T(U^{*})-J^{(\alpha )}_T(U)=2x^\mathrm {T}_T\Pi X^{*}_T-\int \limits _0^T \alpha _t\left (x^\mathrm {T}_tQx_t+u^\mathrm {T}_tRu_t\right )dt-2\int \limits _0^T \sqrt {\alpha _t}x^\mathrm {T}_t\Pi GdW_t , $$
(A.4)
where the variables are \(x_t=X^{*}_t-X_t \) and \(u_t=U^{*}_t-U_t \) with \(dx_t=\alpha _t A x_t dt+\alpha _t B u_t dt \), \(x_0=0 \). Since the pair \((A,C) \) is observable, it follows that there exists a matrix \(F \) such that the matrix \(A+FC \) is exponentially stable. Then
$$ {\|x_t\|\leqslant c \exp {\left \{-\bar \kappa \int \limits _0^t \alpha _v dv\right \}}\int \limits _0^t \exp {\left \{\bar \kappa \int \limits _0^{s} \alpha _v dv\right \}}\alpha _s\big (\|Cx_s\|+\|u_s\|\big )ds}$$
for some constant \(\bar \kappa >0 \). After squaring and applying the Cauchy–Schwarz inequality as well as the conditions \(Q=C^{\mathrm {T}}C\), \(R>0 \), we have
$$ {\|x_t\|^2\leqslant \tilde c \exp {\left \{-\bar \kappa \int \limits _0^t \alpha _v dv\right \}}\int \limits _0^t \exp {\left \{\bar \kappa \int \limits _0^{s} \alpha _v dv\right \}}\alpha _s\left (x^\mathrm {T}_sQx_s+u^{\mathrm {T}}_sRu_s\right )ds}. $$
Further, using integration by parts, we show that
$$ {\int \limits _0^T \|x_t\|^2 ds\leqslant \tilde c\bar \kappa ^{-1} \int \limits _0^T\alpha _t\left (x^\mathrm {T}_tQx_t+u^{\mathrm {T}}_tRu_t\right )dt}. $$
Accordingly, we obtain the estimate
$$ {\|x_T\|^2+\int \limits _0^T \|x_s\|^2 ds\leqslant c_0\int \limits _0^T \alpha _t\left (x^\mathrm {T}_tQx_t+u^\mathrm {T}_tRu_t\right )dt}$$
with \(T>0 \) and some constant \(c_0>0 \). Then, considering the elementary inequality \(2ab\leqslant a^2\bar c+b^2/\bar c\), which holds for any numbers \(a,b \) and \(\bar c>0 \), the expression on the right-hand side of (A.4) is estimated in the form
$$ J^{(\alpha )}_T(U^{*})-J^{(\alpha )}_T(U)\leqslant c_1 \|X^{*}_T\|^2-c_2\int \limits _0^T \alpha _t\|x_t\|^2 dt-2\int \limits _0^T \sqrt {\alpha _t}x^\mathrm {T}_t\Pi GdW_t ,$$
(A.5)
where \(c_1,c_2>0 \) are some constants. After taking the expectation on both sides in (A.5) and dividing by \({\mathrm {E}}(\int \nolimits _0^T \alpha _t dt)\), in the limit as \(T\to \infty \) we use the result (A.2) in the Lemma; this leads to the relation
$$ \limsup _{T \to \infty }\left ({ \mathrm {E}J^{(\alpha )}_T(U^{*})}\Bigg /{ \mathrm {E}\left (\displaystyle \int \limits _0^T\alpha _t dt\right )}\right )\leqslant \limsup _{T \to \infty }\left ({ \mathrm {E}J^{(\alpha )}_T(U)}\Bigg /{ \mathrm {E}\left (\displaystyle \int \limits _0^T\alpha _t dt\right )}\right ).$$
In the pathwise analysis of (A.5), introducing the notation
$$ {M_T=-2\int \limits _0^T \sqrt {\alpha _t}x^\mathrm {T}_t\Pi GdW_t, } $$
we write an estimate for (A.5) in the form
$$ J^{(\alpha )}_T(U^{*})\leqslant J^{(\alpha )}_T(U)+\mathcal {R_T} ,$$
where \(\mathcal {R}_T=-c_3\langle M_T\rangle +M_T \) for some constant \(c_3>0 \); \(\langle M_T\rangle =\int \nolimits _0^T \|\sqrt {\alpha _t}G^{\mathrm {T}}\Pi x_t\|^2 dt \) is the quadratic variation of \(M_T \). Note that \(\limsup _{T\to \infty }g_T\mathcal {R}_T\leqslant 0\) a.s. for any monotone function \(g_T\) with the property \(g_T>0 \) and \(g_T\to \infty \) as \(T\to \infty \) (see [34]); in particular, \(g_T=\left (\int \nolimits _0^T \alpha _t dt\right )^{-1}\) (see also Assumption \(\mathcal {A} \)). It is also obvious that (13) implies \(\|X^{*}_T\|^2\left (\int \nolimits _0^T \alpha _t dt\right )^{-1} \to 0\) a.s. as \(T\to \infty \). Therefore, with probability \(1 \) we have the relation
$$ \limsup _{T \to \infty }\left ({J^{(\alpha )}_T(U^{*})}\Bigg /{\left (\int \limits _0^T\alpha _t dt\right )}\right )\leqslant \limsup _{T \to \infty }\left ({J^{(\alpha )}_T(U)}\Bigg /{\left (\int \limits _0^T\alpha _t dt\right )}\right ). $$
By the Itô formula, in a standard manner we establish that
$$ {J^{(\alpha )}_T(U^{*})=x^\mathrm {T}\Pi x-(X^{*}_T)^{\mathrm {T}}\Pi X^{*}_T+\mathrm {tr}\thinspace (G^\mathrm {T}\Pi G)\int \limits _0^T \alpha _t dt+2\int \limits _0^T \sqrt {\alpha _t}(X^{*}_t)^\mathrm {T}\Pi GdW_t}. $$
Then the value of the criterion in (6) for \(U^{*}\) is equal to
$$ {\lim _{T\to \infty }\left \{ \mathrm {E}J^{(\alpha )}_T(U^{*}) \mathrm {E}\left (\int \limits _0^T\alpha _t dt\right )^{-1}\right \}=\mathrm {tr}\thinspace (G^\mathrm {T}\Pi G)}.$$
Applying the iterated logarithm law to the martingale \( \mathcal {M}_T=\int \nolimits _0^T \sqrt {\alpha _t}(X^{*}_t)^{\mathrm {T}}\Pi GdW_t\) yields the estimate \(\|\mathcal {M}_T\|\leqslant c \sqrt {\langle \mathcal {M}_T\rangle \ln \ln \langle \mathcal {M}_T\rangle } \) for large \(T \), and the use of (13) allows one to pass to the ineqaulity \(\langle \mathcal {M}_T\rangle \leqslant \tilde c \left ( \int \nolimits _0^T \alpha _t dt \right )\ln \left ( \int \nolimits _0^T \alpha _t dt\right )\). Consequently, \( \|\mathcal {M}_T\| \left (\int \nolimits _0^T \alpha _t dt\right )^{-1} \to 0 \) as \(T\to \infty \) a.s.; then
$$ {\lim _{T\to \infty }\left \{J^{(\alpha )}_T(U^{*})\left (\int \limits _0^T\alpha _t dt\right )^{-1}\right \}=\mathrm {tr}\thinspace (G^\mathrm {T}\Pi G) }$$
with probability \( 1\). The proof of the theorem is complete. \(\quad \blacksquare \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Palamarchuk, E.S. Optimal Control for a Linear Quadratic Problem with a Stochastic Time Scale. Autom Remote Control 82, 759–771 (2021). https://doi.org/10.1134/S0005117921050027

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0005117921050027

Keywords

Navigation