math0701580/GM.tex
1: % 31.07.2006: New version: approximations etc. (Carlo)
2: % 24.08.2006: Changed introduction (Carlo)
3: % 26.08.2006: Abstract rewritten (Carlo)
4: % 31.08.2006: minor corrections (Carlo)
5: % 11.10.2006: added figures of Sergei (Carlo)
6: % 07.06.2007: changed title and introduction (CM)
7: % xx.06.2008: referees' corrections
8: 
9: \documentclass[a4paper,11pt]{article}
10: \usepackage{amsfonts,amsmath,amsthm}
11: \usepackage{graphicx}
12: %\usepackage{showkeys}
13: \usepackage{color}
14: 
15: \newcommand{\A}{\mathcal{A}}
16: \newcommand{\E}{\mathbb{E}}
17: \newcommand{\erre}{\mathbb{R}}
18: 
19: \newcommand{\cp}[2]{\langle#1,#2\rangle}
20: \newcommand{\ds}{\displaystyle}
21: \newcommand{\tr}{\mathop{\mathrm{Tr}}\nolimits}
22: 
23: \newtheorem{prop}{Proposition}[section]
24: \newtheorem{thm}[prop]{Theorem}
25: \newtheorem{coroll}[prop]{Corollary}
26: \newtheorem{lemma}[prop]{Lemma}
27: \theoremstyle{definition}
28: \newtheorem{rmk}[prop]{Remark}
29: 
30: \newsymbol\lesssim 132E
31: 
32: \newcommand{\carlo}[1]{\textcolor{red}{#1}}
33: 
34: \title{On controlled linear diffusions with delay in a model of
35:   optimal advertising under uncertainty with memory effects}
36: 
37: \author{Fausto Gozzi\thanks{Facolt\`a di Economia, Libera Universit\`a
38:     degli Studi Sociali ``Guido Carli'', 00198 Roma, Italy. e-mail:
39:     \texttt{fgozzi@luiss.it}.} \and Carlo Marinelli\thanks{Institut
40:     f\"ur Angewandte Mathematik, Universit\"at Bonn, Wegelerstr. 6,
41:     D-53115 Bonn, Germany. URL:
42:     \texttt{http://www.uni-bonn.de/$\sim$cm788}.} \and Sergei
43:   Savin\thanks{Graduate School of Business, Columbia University, New
44:     York, NY 10027, USA. e-mail: \texttt{svs30@columbia.edu}.}}
45: 
46: \date{\normalsize June 7, 2007. Revised July 21, 2008.}
47: 
48: \addtolength{\hoffset}{-1.5cm}
49: \addtolength{\textwidth}{3cm}
50: \addtolength{\voffset}{-1.75cm}
51: \addtolength{\textheight}{3.5cm}
52: 
53: \begin{document}
54: \maketitle
55: 
56: \begin{abstract}
57:   We consider a class of dynamic advertising problems under
58:   uncertainty in the presence of carryover and distributed forgetting
59:   effects, generalizing the classical model of Nerlove and Arrow
60:   \cite{NA}. In particular, we allow the dynamics of the product
61:   goodwill to depend on its past values, as well as previous
62:   advertising levels. Building on previous work (\cite{levico}), the
63:   optimal advertising model is formulated as an infinite dimensional
64:   stochastic control problem. We obtain (partial) regularity as well
65:   as approximation results for the corresponding value function. Under
66:   specific structural assumptions we study the effects of delays on
67:   the value function and optimal strategy. In the absence of carryover
68:   effects, since the value function and the optimal advertising policy
69:   can be characterized in terms of the solution of the associated HJB
70:   equation, we obtain sharper characterizations of the optimal policy.
71:   
72:   \medskip\par\noindent \emph{Keywords:} stochastic control problems
73:   with delay, dynamic programming, infinite dimensional Bellman
74:   equations, optimal advertising.
75: \end{abstract}
76: 
77: 
78: 
79: \section{Introduction}
80: This paper is devoted to the study of a class of optimal control
81: problems for linear stochastic differential equations with delay both
82: in the state and the control term, and is a natural continuation of
83: \cite{levico}. These problems arise in the theory of optimal
84: advertising under uncertainty with memory structures. We approach the
85: problem using stochastic control techniques in infinite dimensions.
86: 
87: In particular, in \cite{levico} we considered a controlled stochastic
88: differential equation (SDE) with delay entering both the state and the
89: control variable as an extension of the dynamic advertising model of
90: Nerlove and Arrow \cite{NA}. The results of \cite{levico} are the
91: following: we construct a controlled infinite dimensional SDE that is
92: equivalent to the controlled SDE with delay, we prove a verification
93: theorem, and we exhibit a simple example for which the Bellman
94: equation associated to the control problem admits a sufficiently
95: regular solution, hence the verification theorem can be applied. 
96: %
97: In the present manuscript, we extend \cite{levico} by
98:   developing several new sets of results.  On the one hand, we provide
99:   qualitative characterization of the first- and second-order
100:   properties of the optimal value function (Section
101:   \ref{sec:reformul}.1). In particular, we show that, under natural
102:   restrictions, the monotonicity of the optimal value function with
103:   respect to the initial goodwill profile still holds even in the
104:   presence of the state and control-related delay terms in the
105:   advertising dynamics (Proposition \ref{prop:Vincr}). In addition, we
106:   establish that the decreasing marginal influence of the attained
107:   goodwill levels on the primal profit components is retained by the
108:   optimal profit function (Proposition \ref{prop:Vconv}). As is well
109:   known, this last property is important in reducing the computational
110:   load required to solve the time- and space-discretized version of
111:   our problem by dynamic programming methods.
112: %
113:   On the other hand, in the view of intractability of the general
114:   variant of our problem, we propose approximation schemes for the
115:   optimal value function (Theorem \ref{prop:approx}) and for the
116:   optimal advertising policy (Propositions 3.11 and 3.12). The latter
117:   result is of particular importance since it suggests a
118:   computationally feasible approach to constructing asymptotically
119:   optimal advertising trajectories.
120: %
121:   In addition, we provide a complete characterization of the optimal
122:   advertising policy in the case when the cost function is quadratic
123:   and the reward function is linear in goodwill level (Section
124:   3.3). For a specific instance of this case we conduct a numerical
125:   study aimed at demonstrating the importance of proper accounting of
126:   the delay effects in calculating the optimal advertising policy.
127: 
128:   Finally, we are able to provide sharper characterization of the
129:   optimal policies in the case when the influence of advertising on
130:   the goodwill evolution is instantaneous and the delay effects are of
131:   the state-only type (Section \ref{sec:ex}). The key result in this
132:   section is Theorem \ref{thm:L2} which formulates sufficient
133:   conditions ensuring that the optimal advertising policy is of the
134:   feedback type. In particular, in the case of linear cost function
135:   the optimal control takes a particularly simple ``bang-bang'' form
136:   (Corollary \ref{cor:bang}).
137: 
138: % Using the setting of \cite{levico}, here we prove qualitative
139: % properties of the value function (monotonicity, concavity,
140: % regularity), and obtain approximation results for the value function
141: % and the optimal strategy.  In particular, the value function is
142: % characterized as the limit of a sequence of solutions of approximating
143: % ``regularized'' Bellman equations. The reason to look for such results
144: % is that the Bellman equation associated to the control problem cannot
145: % be solved directly, i.e. it is not covered by any of the available
146: % techniques (see \cite{DZ02}, \cite{FT1}, \cite{GolGoz}), with the
147: % possible exception of the viscosity approach, which nonetheless
148: % requires an ad hoc treatment and does not allow one to construct
149: % optimal strategies.  A study of the effects of delay with respect to a
150: % model without delay is in general not feasible without making specific
151: % structural assumptions. This is due to the fact that a controlled SDE
152: % with delay is essentially an infinite dimensional object, for which
153: % hardly anything can be computed explicitly. However, the special case
154: % of quadratic cost and linear reward, for which the Bellman equation
155: % admits an explicit solution, provides us with a realistic example
156: % where the effects of delay can be observed. Moreover, in the case of
157: % delay in the state only, we show that in some examples of interest one
158: % can give qualitative characterizations of the optimal strategy.
159: 
160: Optimal control problems for stochastic systems with delay in the
161: state term admit alternative, more traditional treatments: for
162: instance, see \cite{elsanosi} and \cite{larssen} for a more direct
163: application of the dynamic programming principle without appealing to
164: infinite dimensional analysis, and \cite{KSlibro} for the
165: linear-quadratic case. However, we would like to point out that none
166: of the methods just mentioned apply to the control of stochastic
167: differential equations with delay in the control term.
168: 
169: Analysis of advertising policies has always been occupying a
170: front-and-center place in the marketing research. The sheer size of
171: the advertising market (over \$143 billion in the US in 2005
172: \cite{admarket}) and the strong body of evidence of systematic
173: over-advertising by firms across many industries (see e.g.
174: \cite{overad1}, \cite{overad4}, \cite{overad2}, \cite{overad3})
175: has caused a renewal of attention to the proper accounting for the
176: so-called ``carryover'' or ``distributed lag'' advertising effects.
177: The term ``carryover'' designates an empirically observed
178: advertising feature under which the advertising influence on product
179: sales or goodwill level is not immediate, but rather is spread over
180: some period of time: according to a survey of recent empirical
181: ``carryover'' research by Leone \cite{leone}, delayed advertising effects
182: can last between 6 and 9 months in different settings.
183: 
184: On the theoretical front, pioneering work of \cite{VW} and \cite{NA}
185: has paved the way for the development of a number of models dealing
186: with the optimal distribution of advertising spending over time in
187: both monopolistic and competitive settings. A comprehensive review of
188: the state of the advertising control literature in \cite{sethi} points
189: out that the majority of these models operate under deterministic
190: assumptions and do not capture some of the most essential
191: characteristics of real-world advertising phenomena. On the empirical
192: side, one of the first and most important substreams of advertising
193: literature was formed by the papers focused on the studies of
194: distributed advertising lag (see e.g. \cite{bass2},
195: \cite{bassparsons}, \cite{griliches}). An important early empirical
196: result was obtained by Bass and Clark \cite{bassclark}, who
197: established that the initially adopted models with monotone decreasing
198: lags (see \cite{Koyck}) are often inferior in their explanatory power
199: to the models with more general lag distributions.
200: 
201: Despite the wide and growing empirical literature on the measurement
202: of carryover effects, there are practically no analytical studies that
203: incorporate distributed lag structure into the optimal advertising
204: modeling framework in the stochastic setting. The only papers dealing
205: with optimal dynamic advertising with distributed lags we are aware of
206: are \cite{bassparsons} (which provides a numerical solution to a
207: discrete-time deterministic example), and \cite{hartl}, \cite{HS}
208: (which applies a version of the maximum principle in the deterministic
209: setting). The creation of models which incorporate the treatment of
210: ``carryover'' effects in the stochastic settings have long been
211: advocated in the advertising modeling literature (see e.g.
212: \cite{hartl}, \cite{sethi} and references therein).
213: 
214: As mentioned above, in this work we study a class of stochastic models
215: deriving from that of Nerlove and Arrow \cite{NA}, incorporating both
216: the advertising lags as well as distributed ``churn''
217: (\cite{sethi-VWcomp}), or ``forgetting'', effects.  More precisely,
218: we formulate an optimization program that seeks to maximize the
219: goodwill level at a given time $T>0$ net of the cumulative cost of
220: advertising until $T$. This optimization problem is studied using
221: techniques of stochastic optimal control in infinite dimensions, using
222: the modeling approach of \cite{levico}: in particular, we specify the
223: goodwill dynamics in terms of a controlled stochastic delay
224: differential equation (SDDE), that can be rewritten as a stochastic
225: differential equation (without delay) in a suitable Hilbert space.
226: This allows us to associate to the original control problem for the
227: SDDE an equivalent infinite dimensional control problem for the
228: ``lifted'' stochastic equation.
229: 
230: The paper is organized as follows: in section \ref{sec:formul} we
231: formulate the optimal advertising problem as an optimal control
232: problem for an SDE with delay, and we recall the equivalence result of
233: \cite{levico}. In section \ref{sec:reformul} we prove the above
234: mentioned results about the value function and approximate strategies
235: in the general case, together with a detailed discussion of the effect
236: of delays in a specific situation. Section \ref{sec:ex} treats the
237: case of distributed forgetting in the absence of advertising
238: carryover.
239: 
240: Let us conclude this introduction fixing notation and recalling some
241: notions that will be needed. Given a lower semicontinuous convex
242: function $f:E\to\bar{\erre}:=\erre\cup\{+\infty\}$ on a Hilbert space
243: $E$ with inner product $\cp{\cdot}{\cdot}$, we denote its conjugate by
244: $f^*(y):=\sup_{x\in E}(\cp{x}{y}-f(x))$. Recall also that $D^-
245: f^*(y)=\arg\max_{x\in E} (\cp{x}{y}-f(x))$, where $D^-$ stands for the
246: subdifferential operator (see e.g. \cite{barbu-v}, p.~103).
247: %
248: Throughout the paper, $X$ will be the Hilbert space defined as
249: $$
250: X = \erre \times L^2([-r,0],\erre),
251: $$
252: with inner product
253: $$
254: \langle x,y\rangle = x_0y_0 + \int_{-r}^0 x_1(\xi)y_1(\xi)\,d\xi
255: $$
256: and norm
257: $$
258: |x| = \left( |x_0|^2 + \int_{-r}^0 |x_1(\xi)|^2\,d\xi\right)^{1/2},
259: $$
260: where $r>0$, $x_0$ and $x_1(\cdot)$ denote the $\erre$-valued and the
261: $L^2([-r,0],\erre)$-valued components, respectively, of the generic
262: element $x$ of $X$. Given $f:X\to\erre$, $k\in\{0,1\}$, we shall
263: denote by $\partial_k f$, $D^-_k f$, respectively, the partial
264: derivative and the subdifferential of $f$ with respect to the $k$-th
265: component.  We shall use mollifiers in a standard way: for $\zeta \in
266: C^\infty(\erre^d,\erre_+)$, equal to zero for $|x|>1$ and such that
267: $\int_{\erre^d}\zeta(x)\,dx=1$, we shall set
268: $\zeta_\lambda(x)=\lambda^{-d}\zeta(\lambda^{-1}x)$ for $\lambda\neq 0$.
269: %
270: By $a \lesssim b$ we mean that there exists a constant $N$ such that
271: $a \leq Nb$. If $N$ depends on some parameter of interest $p$, we
272: shall write $N(p)$ and $a \lesssim_p b$.
273: 
274: 
275: \section{The model}\label{sec:formul}
276: We consider a monopolistic firm preparing the market introduction of a
277: new product at some time $T$ in the future. In defining the state
278: descriptor for a firm to follow we use the Nerlove-Arrow framework and
279: consider the product's ``goodwill stock'' $y(t)$, $0\leq s \leq t \leq
280: T$. The firm directly influences the rate of advertising spending
281: $z(t)$ to induce the following trajectory for the goodwill stock:
282: \begin{equation}
283: \label{eq:SDDE}
284: \left\{\begin{array}{l}
285: dy(t) = \ds \left[a_0 y(t) + \int_{-r}^0 a_1(\xi)y(t+\xi)\,d\xi
286:         + b_0 z(t) +
287:             \int_{-r}^0b_1(\xi)z(t+\xi)\,d\xi\right]dt \\[10pt]
288: \ds \qquad\qquad  + \sigma\, dW_0(t), \quad s \leq t\leq T \\[10pt]
289: y(s)=x_0; \quad y(s+\xi)=x_1(\xi),\;
290: z(s+\xi)=\delta(\xi),\;\; \xi\in[-r,0],
291: \end{array}\right.
292: \end{equation}
293: where the Brownian motion $W_0$ is defined on a filtered probability
294: space $(\Omega,\mathcal{F},\mathbb{F}=(\mathcal{F}_t)_{t\geq 0},%
295: \mathbb{P})$, with $\mathbb{F}$ being the completion of the filtration
296: generated by $W_0$. We assume that the advertising spending rate
297: $z(t)$ is constrained to remain in the set $\mathcal{U}$, the space of
298: $\mathbb{F}$-adapted processes taking values in a compact
299:   interval $U\subseteq\erre_+$. In addition, we assume that the
300: following conditions hold:
301: \begin{itemize}
302: \item[(i)] $a_0 \leq 0$;
303: \item[(ii)] $a_1(\cdot) \in L^2([-r,0],\erre)$;
304: \item[(iii)] $b_0 \geq 0$;
305: \item[(iv)] $b_1(\cdot) \in L^2([-r,0],\erre_+)$;\label{iv}
306: \item[(v)] $x_0 \geq 0$;
307: \item[(vi)] $x_1(\cdot) \geq 0$, with $x_1(0)=x_0$;
308: \item[(vii)] $\delta(\cdot) \geq 0$.
309: \end{itemize}
310: %
311: Here $a_0$ and $a_1(\cdot)$ describe the process of goodwill
312: deterioration when the advertising stops, and $b_0$ and $b_1(\cdot)$
313: provide the characterization of the effect of the current and the past
314: advertising rates on the goodwill level. The values of $x_0$,
315: $x_1(\cdot)$ and $\delta(\cdot)$ reflect the ``initial'' goodwill and
316: advertising trajectories.  Note that we recover the model of Nerlove
317: and Arrow from (\ref{eq:SDDE}) in the deterministic setting
318: ($\sigma=0$) in the absence of delay effects
319: ($a_1(\cdot)=b_1(\cdot)=0$).
320: 
321: Setting $X \ni x:=(x_0,x_1(\cdot))$ and denoting by $y^{s,x,z}(t)$,
322: $t\in[0,T]$, a solution of (\ref{eq:SDDE}), we define the objective
323: functional
324: \begin{equation}
325: \label{eq:obj-orig}
326: J(s,x;z) =
327: \E\left[\varphi_0(y^{s,x,z}(T))-\int_s^T h_0(z(t))\,dt \right],
328: \end{equation}
329: where $\varphi_0:\erre\to\erre$ and $h_0:\erre_+\to\erre_+$ are
330: measurable utility and cost functions, respectively, satisfying the
331: conditions
332: \begin{equation}\label{eq:hphipol}
333: |\varphi_0(x)| \leq K(1+|x|)^m
334: \qquad \forall x\in\erre,
335: \end{equation}
336: and
337: \begin{equation}
338:   \label{eq:schip}
339:   |h(x)| \leq K \qquad \forall x \in U,
340: \end{equation}
341: for some $K>0$ and $m\geq 0$. In the sequel we shall often move the
342: superscripts $s,x,z$ to the expectation sign, with obvious meaning of
343: the notation.
344: %
345: Let us also define the value function $V$ for this problem as follows:
346: $$
347: V(s,x) = \sup_{z\in\mathcal{U}} J(s,x;z).
348: $$
349: We shall say that $z^*\in\mathcal{U}$ is an optimal strategy if it
350: is such that
351: $$
352: V(s,x)=J(s,x;z^*).
353: $$
354: %
355: The problems we will deal with are the maximization of the objective
356: functional $J$ over all admissible strategies $\mathcal{U}$, and the
357: characterization of the value function $V$ and of the optimal strategy
358: $z^*$.
359: 
360: Throughout the paper we will always assume that the assumptions of
361: this section hold true. In particular the constants $T$, $m$ and $K$
362: are fixed from now on.
363: 
364: 
365: \subsection{An equivalent infinite dimensional Markovian
366:   representation}
367: We shall recall a representation result (proposition \ref{prop:equiv}
368: below) proved in \cite{levico}, generalizing a corresponding
369: deterministic result due to Vinter and Kwong \cite{VK}.
370: 
371: \noindent
372: Let us define an operator $A:D(A)\subset X\to X$ as
373: follows:
374: \begin{eqnarray*}
375: A: (x_0,x_1(\xi)) &\mapsto& \Big(a_0x_0 + x_1(0),a_1(\xi)x_0
376:                            -{dx_1(\xi)\over d\xi}\Big) \quad%
377: \textrm{a.e.}\ \xi\in [-r,0], \\
378: D(A) &=& \left\{ x \in X:
379:        x_1 \in W^{1,2}([-r,0];\erre),\; x_1(-r)=0\right\}.
380: \end{eqnarray*}
381: Moreover, define the bounded linear control operator $B:U \to X$ as
382: \begin{equation}
383: B: u \mapsto \Big(b_0u,b_1(\cdot)u\Big),
384: \end{equation}
385: and finally the operator $G:\erre \to X$ as $G: x_0 \mapsto (\sigma
386: x_0, 0)$. Sometimes it will be useful to identify the operator $B$
387: with the element $(b_0,b_1)\in X$.
388: %
389: \begin{prop}\label{prop:equiv}
390:   Let $Y(\cdot)$ be the weak solution of the abstract evolution
391:   equation
392: \begin{equation}
393: \label{eq:abstract}
394: \left\{\begin{array}{l}
395: dY(t) = (A Y(t)+Bz(t))\,dt + G\,dW_0(t) \\[8pt]
396: Y(s) = \bar{x} \in X,
397: \end{array}\right.
398: \end{equation}
399: with arbitrary initial datum $\bar{x} \in X$ and control
400: $z\in\mathcal{U}$. Then, for $t\geq r$, one has,
401: $\mathbb{P}$-a.s.,
402: $$
403: Y(t) = M(Y_0(t),Y_0(t+\cdot),z(t+\cdot)),
404: $$
405: where
406: \begin{eqnarray*}
407: M: X \times L^2([-r,0],\erre) &\to& X \\
408: (x_0,x_1(\cdot),v(\cdot)) &\mapsto& (x_0,m(\cdot)),
409: \end{eqnarray*}
410: $$
411: m(\xi) := \int_{-r}^\xi a_1(\zeta) x_1(\zeta-\xi)\,d\zeta
412: + \int_{-r}^\xi b_1(\zeta) v(\zeta-\xi)\,d\zeta.
413: $$
414: %
415: Moreover, let $\{y(t),\; t\geq -r\}$ be a continuous solution of the
416: stochastic delay differential equation (\ref{eq:SDDE}), and $Y(\cdot)$
417: be the weak solution of the abstract evolution equation
418: (\ref{eq:abstract}) with initial condition
419: $$
420: \bar{x} = M(x_0,x_1,\delta(\cdot)).
421: $$
422: Then, for $t\geq 0$, one has, $\mathbb{P}$-a.s.,
423: $$
424: Y(t) = M(y(t),y(t+\cdot),z(t+\cdot)),
425: $$
426: hence $y(t)=Y_0(t)$, $\mathbb{P}$-a.s., for all $t\geq 0$.
427: \end{prop}
428: %
429: Using this equivalence result, we can now give a Markovian
430: reformulation on the Hilbert space $X$ of the problem of maximizing
431: (\ref{eq:obj-orig}), as in \cite{levico}.
432: In particular, denoting by $Y^{s,\bar{x},z}(\cdot)$ a mild solution of
433: (\ref{eq:abstract}), (\ref{eq:obj-orig}) is equivalent to
434: \begin{equation}
435: \label{ex1jbis}
436: J(s,x;z) =
437: \E\left[ \varphi(Y^{s,\bar{x},z}(T))+\int_s^T h(z(t))\,dt \right],
438: \end{equation}
439: with the functions $h:U\to\erre$ and $\varphi:X\to\erre$ defined by
440: \begin{eqnarray*}
441: h(z) &=& -h_0(z) \\
442: \varphi(x_0,x_1) &=& \varphi_0(x_0).
443: \end{eqnarray*}
444: Hence also $V(s,x)=\sup_{z\in \mathcal{U}} J(s,x;z)$.
445: 
446: 
447: \section{The case of delay in the state and the control term}
448: \label{sec:reformul}
449: The aims of this section are the following: to prove regularity
450: properties of the value function, to develop an approximation scheme
451: for the value function and the optimal strategy, and to illustrate in
452: a numerical example the effects of the delay structures in our model.
453: In particular, we prove that, under natural assumptions, the value
454: function is continuous in both arguments, and monotone concave with
455: respect to the initial goodwill profile. As already remarked, this
456: property is essential in order to obtain computationally tractable
457: discrete-time and discrete-state-space dynamic programming versions of
458: our problem.
459: 
460: Moreover, since we cannot guarantee that the Bellman equation
461: associated to our control problem admits a solution in general (nor do
462: we have any information about its uniqueness and regularity), it is of
463: primary interest to obtain approximation schemes for the optimal value
464: function and for the optimal advertising policy. The latter result is
465: of particular importance since it suggests a computationally feasible
466: approach to constructing asymptotically optimal advertising
467: trajectories.
468: 
469: In addition, in the last subsection we provide a complete
470: characterization of the optimal advertising policy in the case when
471: the cost function is quadratic and the reward function is linear in
472: goodwill level, and for a specific instance of this case we
473: conduct a numerical study aimed at demonstrating the importance of
474: proper accounting of the delay effects in calculating the optimal
475: advertising policy.
476: 
477: \subsection{Qualitative properties of the value function}
478: Let us first show that the value function is finite.
479: \begin{prop}\label{prop:Vfin}
480:   There exists a
481:   constant $N=N(T,m,K)$ such that $|V(s,x) |\leq N(1+|x|)^m$ for all
482:   $s\in[0,T]$, $x\in X$.
483: \end{prop}
484: \begin{proof}
485:   The estimate from below simply follows by taking a constant
486:   deterministic control. For the estimate from above we have,
487:   recalling that $h(x) \leq 0$ for all $x\in U$,
488:   \begin{eqnarray*}
489:     V(s,x) &\leq& \sup_{z\in\mathcal{U}} \E_{s,x}^z \Big[
490:     \int_s^T h(z(t))\,dt + |\varphi(Y(T))| \Big]\\
491:     &\leq& K \sup_{z\in\mathcal{U}} \E_{s,x}^z (1+|Y(T)|)^m.
492:   \end{eqnarray*}
493:   Moreover, we have
494:   \begin{align*}
495:     \E|Y(T)|^m &\lesssim |e^{(T-s)A}x|^m 
496:          + \E\bigg|\int_0^{T-s} e^{(T-s-r)A}Bz(r)\,dr\bigg|^m
497:     + \E\bigg|\int_0^{T-s} e^{(T-s-r)A}G\,dW(r)\bigg|^m\\
498:     &\lesssim_T M_T^m |x|^m + |B|^mM_T^m \bar{z}^m + \E|W_A(T)|^m,
499:   \end{align*}
500:   where $\bar{z}:=\max\{z:\,z\in U\}$, $W_A(t):=\int_0^{t-s}
501:   e^{(t-s-r)A}G\,dW(r)$, and
502:   $M_T:=\sup_{t\in[0,T]}|e^{tA}|$. Recalling that $GG^*$ is of trace
503:   class, hence
504:   \[
505:   \E|W_A(T)|^2 = \mathrm{Tr}\,\int_0^{T-s} e^{tA}GG^*e^{tA^*}\,dt <
506:   \infty
507:   \]
508:   (see e.g. \cite{DP-K}, Proposition 2.2), i.e. $W_A(T)$ is a
509:   well-defined Gaussian random variable on $X$, we also get that
510:   $\E|W_A(T)|^m<\infty$. The proof is completed observing that the
511:   upper bound on $\E|Y(T)|^m$ is uniform over $z$.
512: \end{proof}
513: %
514: %\begin{rmk}
515: %  Assuming that admissible strategies are Markovian of the type
516: %  $z(t)=f(t,Y(t))$ with $f(t,x)$ Lipschitz in $x$ uniformly over $t$
517: %  and $|f(t,x)|<K(1+|x|)$ with $K$ independent of $f$, we do not need
518: %  to assume that $U$ nor $h$ are bounded, and a polynomial growth
519: %. Here is a sketch of the argument: if we
520: %  write the state equation (\ref{eq:abstract}) in mild form,
521: %  \[
522: %  Y(t) = e^{tA}x + \int_0^t e^{(t-s)A}Bf(s,Y(s))\,ds + W_A(t),
523: %  \]
524: %  then we get, using elementary inequalities,
525: %  \[
526: %  \E\sup_{t\in[0,T]}|Y(t)|^m \leq K_1\Big( 1 + |x|^m + T
527: %  \E\sup_{t\in[0,T]}|Y(t)|^m + \E\sup_{t\in[0,T]}|W_A(t)|^m \Big),
528: %  \]
529: %and we obtain the required result for a sufficiently small $T$
530: %recalling that, e.g. by Theorem 2.9 of \cite{DP-K}, there exists a
531: %constant $C=C(m,T)$ such that $\E\sup_{t\in[0,T]} |W_A(t)|^m \leq C$.
532: %The result for general $T$ follows by considering the equation on
533: %partitions $[0,t_0]$, $[t_0,2t_0]$, etc. with $t_0$ small enough.
534: %\end{rmk}
535: 
536: 
537: We establish now some qualitative properties of the value function
538: that do not require studying an associated Bellman equation. The
539: following simple result, typical of control problems with linear
540: dynamics, asserts that the value function inherits the concavity with
541: respect to the space variable from the reward and cost functions.
542: %
543: \begin{prop}\label{prop:Vconv}
544:   If $\varphi$ and $h$ are concave,
545:   then the value function $V(s,x)$ is proper concave with respect to
546:   $x$.
547: \end{prop}
548: \begin{proof}
549: Properness follows by the previous proposition. Moreover, let $x^1$,
550: $x^2 \in X$. Then
551: \begin{eqnarray*}
552: \lambda V(s,x^1) + (1-\lambda)V(s,x^2) &=&
553: \lambda \sup_{z\in\mathcal{U}}
554: \E\Big[\int_s^T h(z(t))\,dt + \varphi(y^{s,x^1,z}(T))\Big] \\
555: && + (1-\lambda) \sup_{z\in\mathcal{U}}
556: \E\left[\int_s^T h(z(t))\,dt + \varphi(y^{s,x^2,z}(T))\right] \\
557: &=& \sup_{z^1,z^2\in\mathcal{U}}
558: \E\bigg[\int_s^T [\lambda h(z^1(t))+(1-\lambda)h(z^2(t))]\,dt \\
559: && \phantom{\sup_{z^1,z^2\in\mathcal{Z}}\E\bigg[}
560: +\lambda\varphi(y^{s,x^1,z^1}(T)) + (1-\lambda)\varphi(y^{s,x^2,z^2}(T))\bigg].
561: \end{eqnarray*}
562: Since $\mathcal{U}$ is a convex set and $h$ is concave, then
563: $z_\lambda:=\lambda z^1 + (1-\lambda) z^2$ is admissible for any
564: choice of $z^1$, $z^2\in\mathcal{U}$, and one has
565: \begin{equation}
566:   \label{eq:hc}
567:   h(z_\lambda(s)) \geq \lambda h(z^1(s))+(1-\lambda)h(z^2(s)).
568: \end{equation}
569: Moreover, by linearity
570: of the state equation, it is easy to prove that
571: $$
572: y^{s,x_\lambda,z_\lambda}(T) = \lambda y^{s,x^1,z^1}(T) +
573: (1-\lambda) y^{s,x^2,z^2}(T),
574: $$
575: hence, by the concavity of $\varphi$,
576: \begin{equation}
577:   \label{eq:phic}
578:   \varphi(y^{s,x_\lambda,z_\lambda}(T)) \geq
579:   \lambda \varphi(y^{s,x^1,z^1}(T)) +
580:   (1-\lambda) \varphi(y^{s,x^2,z^2}(T)).
581: \end{equation}
582: Therefore, as a consequence of (\ref{eq:hc}) and (\ref{eq:phic}), we obtain
583: \begin{eqnarray*}
584: \lambda V(s,x^1) + (1-\lambda)V(s,x^2)
585: &\leq&
586: \sup_{z_\lambda\in\mathcal{U}} \E\bigg[\int_s^T h(z_\lambda(t))\,dt +
587: \varphi(y^{s,x_\lambda,z_\lambda}(T)) \bigg] \\
588: &\leq&
589: \sup_{z\in\mathcal{U}} \E\bigg[\int_s^T h(z(t))\,dt +
590: \varphi(y^{s,x_\lambda,z}(T)) \bigg] \\
591: &=& V(s,\lambda x^1+(1-\lambda)x^2),
592: \end{eqnarray*}
593: which proves the claim.
594: \end{proof}
595: %
596: As a consequence of the previous propositions we obtain the following
597: regularity result. Of course it would be ideal to obtain a result
598: guaranteeing that $V \in C^{0,1}([0,T]\times X)$, so that a
599: verification theorem could be proved. Unfortunately we have not been
600: able to obtain such result. We shall prove though that $V$ is locally
601: Lipschitz continuous in the $X$-valued variable.
602: 
603: \begin{coroll}
604:   Under the hypotheses of Proposition
605:   \ref{prop:Vconv}, the value function $V(s,x)$ is locally Lipschitz continuous with
606:   respect to $x\in X$. Moreover, the subgradient $\partial V(s,x)$
607:   with respect to $x$ exists for all $x\in X$ and is locally bounded.
608: \end{coroll}
609: \begin{proof}
610:   The first assertion comes from the fact that a concave locally
611:   bounded function is continuous in the interior of its effective
612:   domain (see e.g. Theorem 2.1.3 in \cite{barbu-v}) and $V(s,x)$ is
613:   finite for all $x\in X$.  Corollary 2.4 in \cite{ET} and the fact
614:   that $D(\phi)^\circ \subset D(\partial \phi)$ for any concave
615:   function $\phi$, where $A^\circ$ denotes the interior of a set $A$,
616:   imply that $V(s,x)$ is locally Lipschitz in $x$, so the assertion on
617:   $\partial V$ follows.
618: \end{proof}
619: %
620: Since $V(s,x)\equiv V(s,x_0,x_1)$ is a concave function of $x_0$ for
621: fixed $s$ and $x_1$, one can also say that $V$ is twice differentiable
622: almost everywhere with respect to $x_0$, as it follows by the
623: Busemann-Feller theorem. A similar statement is not true regarding
624: differentiability with respect to $x_1$, as the Alexandrov theorem is in
625: general no longer true in infinite dimensions.
626: %
627: We now prove that the value function is continuous with respect to
628: the time variable. It is possible to prove local Lipschitz
629: continuity of $V(s,\cdot)$ without appealing to concavity, but
630: assuming local Lipschitz continuity of $\varphi$.
631: \begin{prop} The value function $V(s,x)$ is continuous in $s$. Moreover, if
632: $|\varphi_0(x)-\varphi_0(y)| \leq K(1+R)^m|x-y|$ for all
633:   $|x|$, $|y| \leq R$, then the function $V(s,x)$ is locally Lipschitz
634:   continuous with respect to $x$. Furthermore, there exists a constant
635:   $N=N(K,m)$ such that
636:   \begin{equation}
637:     \label{eq:Vlip}
638:   |V(s,x)-V(s,y)| \leq N(1+R)^m |x-y|
639:   \end{equation}
640:   for all $|x|$, $|y| \leq R$.
641: \end{prop}
642: \begin{proof}
643: Recalling that the difference of two suprema is less or equal to the
644: supremum of the difference, we have
645: \begin{eqnarray*}
646: |V(s,x) - V(s,y)| &\leq&
647: \sup_{z\in\mathcal{U}}|J(s,x;z)-J(s,y;z)|\\
648: &\leq& \sup_{z\in\mathcal{U}} \E|\varphi(Y^{z,s,x}) - \varphi(Y^{z,s,y})|,
649: \end{eqnarray*}
650: and, by Cauchy-Schwarz' inequality,
651: \begin{eqnarray*}
652: && \E|\varphi(Y^{z,s,x}(T)) - \varphi(Y^{z,s,y}(T))| \leq \\
653: && \qquad K\Big(\E(1 + |Y^{z,s,x}(T)|^m + |Y^{z,s,y}(T)|^m)^2\Big)^{1/2}
654:    \Big(\E|Y^{z,s,x}(T) - Y^{z,s,y}(T)|^2\Big)^{1/2}.
655: \end{eqnarray*}
656: Arguing as in the proof of Proposition \ref{prop:Vfin}, there exists a
657: constant $N_1=N_1(K,m)$ such that $\E|Y^{z,s,x}(T)|^m \leq N_1(1+|x|^m)$.
658: Furthermore, a simple calculation reveals that
659: $|Y^{z,s,x}(t)-Y^{z,s,y}(t)|=|e^{(t-s)A}(x-y)|$, hence
660: $$
661: \E|Y^{z,s,x}(T) - Y^{z,s,y}(T)|^2 \leq M_T^2 |x-y|^2,
662: $$
663: and the second claim is proved.
664: %
665: Let us now prove that $V(s,x)$ is continuous in $s$ for a fixed $x$.
666: Let $s_n \uparrow s$ be a given sequence (the case $s_n \downarrow s$
667: is completely similar).
668: Bellman's principle yields
669: \begin{equation}
670:   \label{eq:bp}
671: V(s_n,x) = \sup_{z\in\mathcal{U}} \E\left[
672: \int_{s_n}^s h(z(r))\,dr + V(s,Y^{z,s_n,x}(s)) \right],
673: \end{equation}
674: and choosing a $1/n$-optimal strategy $z_n$ in (\ref{eq:bp}),
675: %i.e. such
676: %that $|V(s_n,x)-J(s_n,x;z_n)|\leq 1/n$,
677: we have
678: \begin{equation}
679: \label{eq:tano}
680: \begin{array}{rcl}
681: \ds \limsup_{n\to\infty}
682: V(s_n,x)-V(s,x) &\leq&
683: \ds \limsup_{n\to\infty} \E\int_{s_n}^s h(z_n(r))\,dr\\[10pt]
684: && \ds + \limsup_{n\to\infty} \E|V(s,Y^{z_n,s_n,x}(s))-V(s,x)|.
685: \end{array}
686: \end{equation}
687: The first term on the right-hand side in (\ref{eq:tano}) is zero
688: because $h$ is bounded on $U$. Let us show the also the second term on
689: the right-hand side of (\ref{eq:tano}) is zero: in fact we have
690: \begin{eqnarray*}
691: \E|Y^{z_n,s_n,x}(s) - x|^2 &\leq& K \Big(
692: \E\int_{s_n}^s |e^{(s-r)A}h(z_n(r))|^2\,dr
693: + \E\Big|\int_0^{s-s_n} e^{(s-s_n-r)A}G\,dW(r)\Big|^2 \Big) \\
694: &\leq& K_1(s-s_n) + q_n \to 0
695: \end{eqnarray*}
696: as $n\to\infty$, hecause $h$ is bounded on $U$ and the stochastic
697: convolution is a Gaussian random variable with covariance operator
698: going to $0$ as $n\to\infty$. Therefore we also have $Y^{z_n,s_n,x}(s)
699: \to x$ in probability. By (\ref{eq:Vlip}), $V(s,x)$ is
700: continuous in $x$ uniformly with respect to $s$, hence
701: $$
702: V(s,Y^{z_n,s_n,x}(s)) \to V(s,x)
703: $$
704: in probability by the continuous mapping theorem. Moreover, recalling
705: that $|V(s,x)|\leq N(1+|x|^m)$ and $\E\sup_{t\in[0,T]}
706: |Y(t)|^m<\infty$, we have
707: \begin{equation}
708: \label{eq:porcon}
709:   \E|V(s,Y^{z_n,s_n,x}(s)) - V(s,x)|^m < \infty,
710: \end{equation}
711: hence Vitali's theorem implies
712: $$
713: \E|V(s,Y^{z_n,s_n,x}(s)) - V(s,x)| \to 0,
714: $$
715: hence $V(s_n,x)-V(s,x) \to 0$ as $n\to \infty$. Furthermore, taking
716: any $z_0 \in U$, we have from (\ref{eq:bp}),
717: $$
718: V(s,x)-V(s_n,x) \leq V(s,x) - (s_n-s)h(z_0) - \E V(s,Y^{z_0,s_n,x}),
719: $$
720: which goes to zero as $n \to \infty$ by (\ref{eq:porcon}), and the
721: claim is proved.
722: \end{proof}
723: %
724: \begin{rmk}
725:   Notice that, by the Rademacher theorem in infinite dimensions, the
726:   previous proposition implies that the value function $V(s,x)$ is
727:   differentiable in a dense subset of $X$. The local Lipschitz
728:   continuity of $V$ also implies that (Clarke's) generalized gradient
729:   of $V(s,x)$ with respect to $x$ is defined everywhere on $X$.
730: \end{rmk}
731: %
732: For the following proposition, which establishes a monotonicity
733: property of the value function, we need to define the natural ordering
734: in $X$: we shall write $x^1\geq x^2$ if $x^1_0 \geq x^2_0$ and $x^1_1
735: \geq x^2_1$ almost everywhere. Similarly, $x^1>x^2$ if the previous
736: inequalities hold with the strict inequality sign.
737: %
738: \begin{prop}\label{prop:Vincr}
739: If $a_1 \geq 0$ and $\varphi_0$ is increasing, then the value
740:   function $V(s,x)$ is increasing with respect to $x$ in the sense just defined.
741: \end{prop}
742: \begin{proof}
743:   The proof is completely analogous to that of proposition
744:   \ref{prop:V-incr} below if we prove that $A$ generates a positivity
745:   preserving semigroup. This is indeed the case: in fact, a direct
746:   calculation shows that $A$ is the adjoint of the operator $\A$
747:   defined in section \ref{sec:ex}. Since $a_1\geq 0$, $\A$
748:   generates a positivity preserving semigroup $S(t)$. It is well known
749:   that $A$ is the generator of the adjoint semigroup $S(t)^*$. Let
750:   $x$, $y$ be arbitrary positive elements of $X$. Then
751: $$
752: 0 \leq \cp{S(t)x}{y} = \cp{x}{S(t)^*y}.
753: $$
754: By the arbitrariness of $x$ and $y$, $S(t)^*$ is positivity preserving.
755: \end{proof}
756: 
757: 
758: \subsection{Approximating the value function and the optimal strategy}
759: Let us now consider the Bellman equation on $X$ associated to the problem of
760: maximizing (\ref{ex1jbis}), which can be written as
761: \begin{equation}
762: \label{eq:HJB0}
763: \left\{\begin{array}{l}
764: \ds v_t + {1\over 2}\tr (GG^*v_{xx}) + \cp{A x}{v_x} + H_0(v_x) = 0,
765: \quad 0 \leq t \leq T\\[8pt]
766: v(T) = \varphi,
767: \end{array}\right.
768: \end{equation}
769: where $H_0(p) = \sup_{z\in U} (\cp{Bz}{p}+h(z))$.
770: 
771: The main problem with (\ref{eq:HJB0}) is that it is not solvable with
772: any of the techniques currently available, with the possible exception
773: of the theory of viscosity solutions. In particular, as of now, one
774: cannot characterize the value function as the (unique) solution, in a
775: suitable sense, of equation (\ref{eq:HJB0}). 
776: %
777: As a consequence, we cannot obtain an optimal strategy for the
778: optimization problem at hand. As a (partial) remedy we develop a
779: method to approximate the value function and to construct suboptimal
780: feedback strategies that are asymptotically optimal, in the sense of
781: Proposition \ref{prop:asop} below.  Let us also briefly recall that,
782: if we know a priori that a smooth solution to the Bellman equation
783: exists, then we can apply the verification theorem proved in
784: \cite{levico}, which in turns allows to obtain precise
785: characterizations of the optimal strategy (see subsection
786: \ref{subsec:exex} and section \ref{sec:ex}).
787: 
788: \smallskip
789: 
790: Let us begin proving some approximation results for the value
791: function $V$. Let $\varepsilon\in]0,1]$ and define $G_\varepsilon:
792: \erre^2\to X$ as
793: $$
794: G_\varepsilon = \left[\begin{array}{cc}
795:                          \sigma_0 & 0 \\
796:                          0        & \varepsilon b_1
797:                       \end{array}
798:                 \right].
799: $$
800: Let $W_1$ be a standard real Wiener process independent of $W_0$, set
801: $W=(W_0,W_1)$, and denote by $\tilde{\mathbb{F}}$ the filtration
802: generated by $W$. Let $\tilde{\mathcal{U}}$ be the set of
803: $\tilde{\mathbb{F}}$-adapted processes taking values in $U$.
804: %
805: \par\noindent
806: %
807: Consider the following approximating SDE on $X$:
808: \begin{equation}
809: \label{eq:eps}
810: dY(t) = [A Y(t) + Bz(t)]\,dt + G_\varepsilon\,dW(t),
811: \quad Y(s)=\bar{x},
812: \end{equation}
813: where $z\in\tilde{\mathcal{U}}$ and $0 \leq s \leq t \leq T$.
814: \par\noindent
815: %
816: For a fixed $z$, let $Y$ and $Y_\varepsilon$ be, respectively,
817: solutions of (\ref{eq:abstract}) and of (\ref{eq:eps}).  Moreover, let
818: us define $\varphi_\varepsilon(x)=(\varphi_{0\varepsilon}(x),0)$,
819: $\varphi_{0\varepsilon}=\tilde{\varphi}_{0\varepsilon}\ast\zeta_\varepsilon$,
820: $\tilde{\varphi}_{0\varepsilon}(x) =
821: \varphi_0(x)\chi_{[-1/\varepsilon,1/\varepsilon]}(x)$, and
822: $h_\varepsilon(x)=(-h_{0\varepsilon}(x),0)$,
823: $h_{0\varepsilon}(x)=h_0\ast\zeta_\varepsilon(x)$. In particular
824: $\varphi_{0\varepsilon}\in C^2_c(\erre)$, $h_{0\varepsilon}\in
825: C^2(\erre)$.  Finally, the approximate objective function and value
826: function are defined as
827: $$
828: J_{\varepsilon_1,\varepsilon_2}(s,x;z) =
829: \E_{s,\bar{x}}^z\Big[\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))
830: + \int_s^T h_{\varepsilon_2}(z(t))\,dt \Big],
831: \qquad
832: V_{\varepsilon_1,\varepsilon_2}(s,x) =
833: \sup_{z\in\tilde{\mathcal{U}}} J_{\varepsilon_1,\varepsilon_2}(s,x;z).
834: $$
835: In the following we shall set
836: $\varepsilon=(\varepsilon_1,\varepsilon_2)$ and
837: $\lim_{\varepsilon\to 0}:= \lim_{\varepsilon_2\to
838: 0}\lim_{\varepsilon_1\to 0}$. Moreover, for $R>0$ we set
839: $C_R=\left\{x\in X: \;|x_0|\le R \right\}$.
840: %
841: \begin{thm}\label{prop:approx}
842:   One has $V_\varepsilon(s,x) \to V(s,x)$ as $\varepsilon \to 0$
843:   uniformly over $s\in[0,T]$, $x\in C_R$, for all $R>0$.
844: \end{thm}
845: \begin{proof}
846: Setting $\eta_{\varepsilon_1}(t)=Y_{\varepsilon_1}(t)-Y(t)$, one has
847: $$
848: \eta_{\varepsilon_1}(t) = \varepsilon_1 \int_s^t e^{(t-r)A}B_1\,dW_1(r)
849: = :\varepsilon_1 \eta(T),
850: $$
851: with $B_1:\erre\to X$, $B_1:x \to (0,b_1(\cdot)x)$,
852: and (suppressing the subscripts on the expectation sign for simplicity)
853: \begin{eqnarray}
854: |J_\varepsilon(s,x;z)-J(s,x;z)| &\leq&
855: |\E[\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T)) - \varphi(Y(T))]|\nonumber\\
856: && + \left|\E\Big[\int_s^T h_{\varepsilon_2}(z(t))\,dt
857:               - \int_s^T h(z(t))\,dt\Big]\right|\nonumber\\
858: &\leq& |\E[\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))
859:            - \varphi_{\varepsilon_2}(Y(T))]|
860: + |\E[\varphi_{\varepsilon_2}(Y(T)) - \varphi(Y(T))]|\nonumber\\
861: \label{eq:Je}
862: && + \left|\E\Big[\int_s^T h_{\varepsilon_2}(z(t))\,dt
863:               - \int_s^T h(z(t))\,dt\Big]\right|
864: \end{eqnarray}
865: Since
866: $$
867: \E|Y_{\varepsilon_1}(T)-Y(T)| =
868: \varepsilon_1 \E\Big|\int_s^t e^{(t-r)A}B_1\,dW_1(r)\Big| \to 0,
869: $$
870: then $Y_{\varepsilon_1}(T) \to Y(T)$ in probability uniformly over
871: $x\in X$ as $\varepsilon_1\to 0$, and by the continuous mapping
872: theorem $\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T)) \to
873: \varphi_{\varepsilon_2}(Y(T))$ in probability. Moreover
874: $\varphi_{\varepsilon_2}(Y(T)) \to \varphi(Y(T))$ in probability
875: uniformly over $x\in C_R$, for all $R>0$, as $\varepsilon_2\to 0$
876: because $\varphi_{0\varepsilon_2}(x)\to\varphi_0(x)$ $dx$-a.e. in
877: $\erre$.
878: %
879: Let us now prove that $\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))$ is
880: uniformly integrable with respect to $\varepsilon$. First let us
881: observe, as it is immediate to show, that there exists $\bar{K}$,
882: independent of $\varepsilon_2$, such that
883: $\varphi_{\varepsilon_2}(x)\leq\bar{K}(1+|x|)^m$.
884: Then we can write
885: \begin{eqnarray}
886: \sup_{\varepsilon\in]0,1]^2} \E|\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))| &\leq&
887: \sup_{\varepsilon_1\in]0,1]} \bar{K}\E(1+|Y(T)+\eta_{\varepsilon_1}(T)|)^m\nonumber \\
888: &\leq& K_1 + K_2 \sup_{\varepsilon_1\in]0,1]}
889:                  (\E|Y(T)|^m + {\varepsilon_1}^m\E|\eta(T)|^m)\nonumber\\
890: &=& \label{eq:ui1}
891:  K_1 + K_2 (\E|Y(T)|^m + \E|\eta(T)|^m) < \infty,
892: \end{eqnarray}
893: where we used twice the inequality $|x+y|^m\leq 2^m(|x|^m+|y|^m)$ and
894: Burkholder-Davis-Gundy's inequality.
895: Furthermore,
896: \begin{equation}
897: \label{eq:ui2}
898: \sup_{\varepsilon\in]0,1]^2} \E[|\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))|\chi_A]
899: \leq \E[(K_1 + K_2(|Y(T)|^m + |\eta(T)|^m))\chi_A] \to 0
900: \end{equation}
901: as $\mathbb{P}(A)\to 0$, because $K_1 + K_2(|Y(T)|^m + |\eta(T)|^m$
902: has finite expectation.  Then (\ref{eq:ui1}) and (\ref{eq:ui2}) imply
903: that $\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T))$ is uniformly
904: integrable (see e.g. \cite{kall}, lemma 3.10), hence
905: $$
906: | \E\varphi_{\varepsilon_2}(Y_{\varepsilon_1}(T)) - \E\varphi(Y(T)) | \to 0
907: $$
908: as $\varepsilon\to 0$ (see e.g. \cite{kall}, proposition 3.12).
909: %
910: \par\noindent
911: %
912: Similarly, since
913: $$
914: |h_{\varepsilon_2}(z(t,\omega)) - h(z(t,\omega))| \leq
915: K_1 + K_2 |z(t,\omega)|^m
916: $$
917: for all $t\in[0,T]$, $\omega\in\Omega$ and $\E\int_0^T |z(t)|^m <
918: \infty$ (because $U$ is compact), by the dominated convergence theorem
919: we have
920: $$
921: \E\int_0^T |h_{\varepsilon_2}(z(t)) - h(z(t)|\,dt \to 0
922: $$
923: as $\varepsilon_2 \to 0$.
924: 
925: In view of (\ref{eq:Je}) we have thus proved that
926: $|J_\varepsilon(s,x;z)-J(s,x;z)| \to 0$ uniformly over $s\in[0,T]$,
927: $x\in C_R$, for all $R>0$ and $z\in\tilde{\mathcal{U}}$, hence also
928: that $V_\varepsilon(s,x) \to V(s,x)$.
929: \end{proof}
930: %
931: %
932: If the cost function $h_0$ is continuous, one can use a different
933: regularization, without requiring compactness of $U$.
934: \begin{prop}
935:   If $h$ is continuous, then the assertion of theorem \ref{prop:approx}
936:   holds.
937: \end{prop}
938: \begin{proof}
939:   Let $h_{\varepsilon,\delta}$ be the sup-inf convolution of $h$ (in
940:   the sense of \cite{LasLio}), that is
941:   $$
942:   h_{\varepsilon,\delta}(x) = \sup_{z\in X}\inf_{y\in X}
943:      \Big(\frac{|z-y|^2}{2\varepsilon} - \frac{|z-x|^2}{2\delta} + h(y) \Big), \qquad 0<\delta<\varepsilon.
944:   $$
945:   It is known that $h_{\varepsilon,\delta}$ is differentiable with
946:   continuous derivative, that $\inf_{x \in X} h(x) \le
947:   h_{\varepsilon,\delta}(x) \le h(x)$ for all $x \in X$ and that
948:   $\lim_{\varepsilon,\delta \to 0^+}h_{\varepsilon,\delta}(x) \rightarrow h(x)$
949: uniformly over $x\in C_R$ (see \cite{LasLio}). Setting
950: $h_{\varepsilon_2}=h_{\varepsilon_2,\varepsilon_2/2}$, proposition
951: \ref{prop:Vfin} and the dominated convergence theorem yield
952:   $$
953:   \E\int_0^T h_{\varepsilon_2}(z(t))\,dt
954:   \stackrel{\varepsilon_2\downarrow 0}{\longrightarrow}
955:   \E\int_0^T h(z(t))\,dt,
956:   $$
957:   which implies that the third term on the right-hand side of
958:   (\ref{eq:Je}) converges to 0 as $\varepsilon \to 0$.
959: \end{proof}
960: %
961: %
962: Theorem \ref{prop:approx} (or its variant), together with the
963: following result, allow one to approximate the value function $V(s,x)$
964: in terms of the solutions of a sequence of Bellman equations.
965: \begin{prop}
966:   Assume that the hypotheses of theorem \ref{prop:approx} are
967:   verified. Assume moreover that $h_0$ is strictly convex and
968:   $\varphi_0$ is concave.  Then the approximate value function
969:   $V_\varepsilon$ is the unique mild solution (in the sense of
970:   \cite{FT05}) of the Bellman equation
971:   $$
972:   v_t + \frac12 \tr(G_{\varepsilon_1} G^*_{\varepsilon_1} v_{xx})
973:   + \cp{Ax}{v_x} + H_{0\varepsilon_2}(v_x) = 0,
974:   \qquad v(T)=\varphi_{\varepsilon_2},
975:   $$
976:   where $H_{0\varepsilon_2}(p)=\sup_{z\in U}(\cp{Bz}{p}+h_{\varepsilon_2}(z))$.
977: \end{prop}
978: \begin{proof}
979: Setting
980: $$
981: \tilde{B}_{\varepsilon_1}:\erre\to\erre^2, \qquad
982: \tilde{B}_{\varepsilon_1} = \left[\begin{array}{c} b_0/\sigma_0 \\
983: \varepsilon_1^{-1}
984:                               \end{array}
985:                         \right],
986: $$
987: the approximating equation (\ref{eq:eps}) can be rewritten as
988: \begin{equation}
989: \label{eq:epss}
990: dY(t) = [AY(t) + G_{\varepsilon_1} \tilde{B}_{\varepsilon_1} z(t)]\,dt
991: + G_{\varepsilon_1}\,dW(t),
992: \quad Y(s)=\bar{x}.
993: \end{equation}
994: The state equation (\ref{eq:epss}), hence also (\ref{eq:eps}), is
995: covered by the FBSDE approach to semilinear PDEs in Hilbert spaces
996: (see e.g. \cite{FT-corso}). In order to prove the statement, we
997: shall verify that hypothesis 7.1 in \cite{FT05} holds true. In
998: particular, $G_{\varepsilon_1}$ is Hilbert-Schmidt because $b_1\in
999: L^2([-r,0],\erre_+)$; $\varphi_{\varepsilon_2}$ is Lipschitz because
1000: $\varphi_{0\varepsilon_2} \in C^2_c(\erre)$;
1001: $|\tilde{B}_{\varepsilon_1} z|_X$ is bounded for $z\in U$ because
1002: $b_1\in L^2([-r,0],\erre)$ and $U$ is compact; finally, since $h_0$
1003: is proper and positive, it is immediate to find
1004: $\varepsilon_0\in]0,1]$, $C\geq 0$ such that $h_{\varepsilon_2}(x)
1005: \geq -C$ and $\inf_U h_{\varepsilon_2} \leq C$, for all positive
1006: $\varepsilon_2<\varepsilon_0$.
1007: 
1008: Since $h$ is strictly convex, for a sufficiently small
1009: $\varepsilon_2$ also $h_{\varepsilon_2}$ is strictly convex.
1010: Therefore we have
1011: $$
1012: g_{\varepsilon_2}(p)= \arg\max_{z\in
1013: U}(\cp{Bz}{p}+h_{\varepsilon_2}(z))=
1014: (h_{\varepsilon_2}')^{-1}(B^*p).
1015: $$
1016: The claim now follows from \cite{FT05}, theorem 7.2 provided we
1017: prove that the closed loop equation
1018: \begin{equation}
1019: \label{eq:CLE} dY(t) = [AY(t) + B g(v_x(t,Y(t))]\,dt +
1020: G_{\varepsilon_1}\,dW(t), \quad Y(s)=\bar{x}.
1021: \end{equation}
1022: admits a solution. In fact this follows as in Theorem 7.2 of
1023: \cite{FT1}.
1024: \end{proof}
1025: %
1026: \begin{rmk}
1027:   In fact the convexity of $U$ implies that $H_{0\varepsilon_2}\in
1028:   C^1(\erre)$, and hence that $V_\varepsilon \in C^{0,1}([0,T],X)$, as
1029:   in corollary \ref{cor:Vdiff} below.
1030: \end{rmk}
1031: 
1032: The above approximations do not give a way to construct approximately
1033: optimal strategies for the original problem. In fact, it is well known
1034: that the problem of constructing approximately optimal controls from
1035: the knowledge of an approximate value function is very hard, and in
1036: general unsolved. However, it is possible to construct a (suboptimal)
1037: feedback control for which we have some error control, in the sense
1038: defined below. For a map $f:[0,T]\times X \to U$ such that the
1039: equation
1040: $$
1041: dY(t) = AY(t)\,dt + Bf(t,Y(t))\,dt + G\,dW(t),\qquad Y(s)=x,
1042: $$
1043: admits a mild solution $Y(t)$, let us set $u_f(t)=f(t,Y(t))$ and
1044: $V^f(s,x)= J(s,x;u_f)$. Similarly we define $V_\varepsilon^f(s,x)$.
1045: Let us suppose that we can obtain a feedback law $f:[0,T]\times X
1046: \to U$, which is approximately optimal for the regularized problem,
1047: and let us write $V_\varepsilon \approx V_\varepsilon^f$ to mean
1048: that the two values differ by a small constant. Moreover, recall
1049: that $V \approx V_\varepsilon \approx V_\varepsilon^f$.
1050: \begin{prop}
1051:   Let $f(t,x)$ be Lipschitz in $x$ uniformly over $t$. Then
1052:   $V_\varepsilon^f(s,x) \to V^f(s,x)$ as $\varepsilon\to 0$.
1053: \end{prop}
1054: %
1055: \begin{proof}
1056: Denote by $Y^f$ and $Y^f_\varepsilon$, respectively, the
1057: solutions of the equations
1058: \begin{eqnarray*}
1059:   dY^f(t) &=& AY^f(t)\,dt + Bf(t,Y^f(t))\,dt + G\,dW(t) \\
1060:   dY^f_\varepsilon(t) &=& AY^f_\varepsilon(t)\,dt
1061:                           + Bf(t,Y^f_\varepsilon(t))\,dt
1062:                           + G_\varepsilon\,dW(t),
1063: \end{eqnarray*}
1064: with $Y^f(s)=Y^f_\varepsilon(s)=x$. Let us assume, without loss of
1065: generality, $s=0$.  Let us show that $Y^f_\varepsilon(t) \to Y^f(t)$
1066: in $L^1(\Omega,\mathbb{P})$, hence in probability, for all
1067: $t\in[0,T]$: by variation of constants we have
1068: \begin{eqnarray*}
1069: |Y^f_\varepsilon(t) - Y^f(t)| &\leq&
1070: \int_0^t |e^{(t-s)A}B(f(s,Y^f_\varepsilon(s))-f(s,Y^f(s)))|\,ds\\
1071: && + \varepsilon \Big|\int_0^t e^{(t-s)A}B_1\,dW_1(s)\Big|\\
1072: &\leq& \int_0^t m(s)|Y^f_\varepsilon(s) - Y^f(s)|\,ds
1073:        + \varepsilon \Big|\int_0^t e^{(t-s)A}B_1\,dW_1(s)\Big|,
1074: \end{eqnarray*}
1075: where $m(s)=|e^{(t-s)A}|\,|B|\,|f|_{\mathrm{Lip}}$. Taking expectation
1076: on both sides and recalling that the stochastic convolution has finite
1077: mean, Gronwall's lemma yields
1078: \begin{equation}
1079:   \label{eq:gro}
1080: \E|Y^f_\varepsilon(t) - Y^f(t)| \leq \varepsilon N e^{\int_0^T m(s)\,ds}
1081: \to 0
1082: \end{equation}
1083: as $\varepsilon \to 0$, hence $Y^f_\varepsilon(t) \to Y^f(t)$ in
1084: probability for all $t\in[0,T]$.
1085: %
1086: \par\noindent
1087: %
1088: By the same arguments used in the proof of theorem \ref{prop:approx}
1089: we obtain that
1090: \begin{equation}
1091:   \label{eq:lim1}
1092: \lim_{\varepsilon_2\to 0}\lim_{\varepsilon_1\to 0}
1093: \E\varphi_{\varepsilon_2}(Y^f_{\varepsilon_1}(T)) = \E\varphi(Y^f(T)).
1094: \end{equation}
1095: Similarly,
1096: \begin{eqnarray*}
1097: |h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))-h(f(t,Y^f(t)))|
1098: &\leq&
1099: |h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))
1100:     - h_{\varepsilon_2}(f(t,Y^f(t)))| \\
1101: && + |h_{\varepsilon_2}(f(t,Y^f(t)))-h(f(t,Y^f(t)))|
1102: \end{eqnarray*}
1103: and
1104: $$
1105: h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))
1106: \to h_{\varepsilon_2}(f(t,Y^f(t)))
1107: $$
1108: in probability as $\varepsilon_1\to 0$ for all $t\in[0,T]$, because
1109: $Y^f_{\varepsilon_1}(t) \to Y^f(t)$ in probability and
1110: $h_{\varepsilon_2}\circ f(t,\cdot)$ is continuous. Furthermore,
1111: $h_{\varepsilon_2}(f(t,Y^f(t))) \to h(f(t,Y^f(t)))$ $\mathbb{P}$-a.s.
1112: for all $t\in[0,T]$ as $\varepsilon_2\to 0$ because $h_{\varepsilon_2}
1113: \to h$ a.e. on $\erre$.  Since
1114: $|h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))| \leq \sup_{x\in
1115:   U}h_{\varepsilon_2}(x) < \infty$, then
1116: $h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))$ is uniformly
1117: integrable with respect to $\varepsilon_1$, hence
1118: \begin{equation}
1119:   \label{eq:lim2}
1120: \E\int_0^T \Big|
1121: h_{\varepsilon_2}(f(t,Y^f_{\varepsilon_1}(t)))                                 - h_{\varepsilon_2}(f(t,Y^f(t)))
1122: \Big|\,dt \to 0
1123: \end{equation}
1124: as $\varepsilon_1\to 0$.
1125: Finally,
1126: \begin{equation}
1127:   \label{eq:lim3}
1128: \E\int_0^T \Big|h_{\varepsilon_2}(f(t,Y^f(t)))\,dt
1129: - h(f(t,Y^f(t)))\Big|\,dt \to 0
1130: \end{equation}
1131: as $\varepsilon_2\to 0$ by the dominated convergence theorem, taking
1132: into account that
1133: $$
1134: \E \int_0^T |h_{\varepsilon_2}(f(t,Y^f(t)))-h(f(t,Y^f(t)))|\,dt
1135: < \infty,
1136: $$
1137: because $f$ is bounded, as follows by the compactness of $U$.  The
1138: claim now follows by (\ref{eq:lim1}), (\ref{eq:lim2}) and
1139: (\ref{eq:lim3}).
1140: \end{proof}
1141: %
1142: The previous proposition does not obviously allow one to say that
1143: $u(t)=f(t,Y(t))$ is an approximately optimal feedback map for the
1144: original problem, as $f$ itself in general depends on $\varepsilon_1$,
1145: $\varepsilon_2$.
1146: %
1147: The next proposition gives quantitative estimates on
1148: $|V^f(s,x)-V^f_\varepsilon(s,x)|$.
1149: %
1150: \begin{prop}     \label{prop:asop}
1151:   Assume that $f(t,x)$ is Lipschitz in $x$ uniformly over $t$,
1152:   and that $\varphi$, $h$ are Lipschitz continuous. Then there exist
1153:   constants $N=N(|f|_\mathrm{Lip})$ and $\delta=\delta(\varepsilon_2)$
1154:   such that
1155: $$
1156: |V^f(s,x)-V^f_\varepsilon(s,x)| \leq N\varepsilon_1 + \delta(\varepsilon_2)
1157: $$
1158: with $\lim_{\varepsilon_2\to 0} \delta(\varepsilon_2) = 0$.
1159: \end{prop}
1160: %
1161: \begin{proof}
1162: Let us write
1163: \begin{eqnarray*}
1164: |V^f-V^f_\varepsilon| &\leq&
1165: \E| \varphi_{\varepsilon_2}(Y^f_{\varepsilon_1}(T)) - \varphi(Y^f(T))| \\
1166: && + \E\int_0^T \Big|h_{\varepsilon_2}(f(t,Y^f(t)))\,dt                           - h(f(t,Y^f(t)))\Big|\,dt \\
1167: &=& I_1+I_2,
1168: \end{eqnarray*}
1169: and
1170: \begin{eqnarray*}
1171: I_1 &\leq& |\E[\varphi_{\varepsilon_2}(Y^f_{\varepsilon_1}(T))
1172:            - \varphi_{\varepsilon_2}(Y^f(T))]|
1173: + |\E[\varphi_{\varepsilon_2}(Y^f(T)) - \varphi(Y^f(T))]|\\
1174: &=& I_{11} + I_{12}.
1175: \end{eqnarray*}
1176: Then
1177: $$
1178: I_{11} \leq |\varphi_{\varepsilon_2}|_{\mathrm{Lip}}
1179: \E|Y^f_{\varepsilon_1}(T) - Y^f(T)| \leq
1180: |\varphi|_{\mathrm{Lip}} \, \varepsilon_1 e^{\int_0^T m(s)\,ds},
1181: $$
1182: where we used the fact that mollification does not increase the
1183: Lipschitz constant.
1184: We also have
1185: \begin{eqnarray*}
1186: I_{12} &\leq& \E\Big[
1187: |\varphi_{0\varepsilon_2}(Y_0^f(T)) - \varphi_0(Y_0^f(T))|;
1188: |Y_0^f(T)| \leq \varepsilon_2^{-1}
1189: \Big]\\
1190: && + \E\Big[
1191: |\varphi_{0\varepsilon_2}(Y_0^f(T)) - \varphi_0(Y_0^f(T))|;
1192: |Y_0^f(T)| > \varepsilon_2^{-1}
1193: \Big]\\
1194: &\leq& \delta_1(\varepsilon_2) + \delta_2(\varepsilon_2),
1195: \end{eqnarray*}
1196: where
1197: $$
1198: \delta_1(\varepsilon_2) = \sup_{|x|\leq1/\varepsilon_2}
1199: |\varphi_{0\varepsilon_2}(x) - \varphi_0(x)| < \infty,
1200: $$
1201: as $\varphi_{0\varepsilon_2}$ converges to $\varphi_0$ uniformly
1202: on compact sets, and $\delta_2(\varepsilon_2)$ is defined as follows:
1203: there exist $K_1$, $K_2\geq 0$ such that
1204: $$
1205: \E\Big[
1206: |\varphi_{0\varepsilon_2}(Y_0^f(T)) - \varphi_0(Y_0^f(T))|;
1207: |Y_0^f(T)| > \varepsilon_2^{-1}
1208: \Big]
1209: \leq
1210: \E\Big[ K_1+K_2|Y_0^f(T)|^m; |Y_0^f(T)| > \varepsilon_2^{-1} \Big],
1211: $$
1212: and
1213: \begin{eqnarray*}
1214: Y^f(T) &\leq& e^{TA}x
1215: + \int_0^T e^{(T-t)A}Bf(t,Y^f(t))\,dt
1216: +  \int_0^T e^{(T-t)A}G\,dW(t) \\
1217: &\leq& e^{TA}x
1218: + \int_0^T e^{(T-t)A}R\,dt
1219: + \int_0^T e^{(T-t)A}G\,dW(t) =: \mu_2 + Z_1.
1220: \end{eqnarray*}
1221: Similarly,
1222: $$
1223: Y^f(T) \geq e^{TA}x
1224: + \int_0^T e^{(T-t)A}r\,dt
1225: + \int_0^T e^{(T-t)A}G\,dW(t) =: \mu_1 + Z_1,
1226: $$
1227: where $\mu_1$, $\mu_2\in X$, $U \subseteq [r,R]$, and
1228: $Z_1$ is a centered $X$-valued Gaussian random variable.
1229: Denoting by $Z$ the $\erre$-valued components of $Z_1$, we have
1230: $$
1231: Y^f_0(T) - Z \leq (\mu_2)_0,
1232: \qquad
1233: Y^f_0(T) - Z \geq (\mu_1)_0,
1234: $$
1235: hence $|Y^f_0(T) - Z| \leq |(\mu_1)_0| \vee |(\mu_2)_0|=:\mu$, or
1236: equivalently $|Y_0^f(T)| \leq \mu + |Z|$.  In particular, $Z$ is a
1237: centered Gaussian random variable. Then
1238: \begin{eqnarray*}
1239: \E\Big[ K_1+K_2|Y_0^f(T)|^m; |Y_0^f(T)| > \varepsilon_2^{-1} \Big]
1240: &\leq& K_1 \mathbb{P}(|Z|+\mu > \varepsilon_2^{-1}) \\
1241: && + K_2\E\Big[ (|Z|+\mu)^m; |Z|+\mu > \varepsilon_2^{-1} \Big]\\
1242: &=:& \delta_2(\varepsilon_2).
1243: \end{eqnarray*}
1244: Note that $\delta_2(\varepsilon_2) \to 0$ as $\varepsilon_2\to 0$
1245: since $\E(|Z|+\mu)^m < \infty$.
1246: 
1247: We have
1248: \begin{eqnarray*}
1249: I_2 &\leq& \Big|
1250: h_{\varepsilon_2}(f(\cdot,Y_{\varepsilon_1}))
1251: - h_{\varepsilon_2}(f(\cdot,Y))
1252: \Big|_{L^1_T}
1253: + \Big|
1254: h_{\varepsilon_2}(f(\cdot,Y))
1255: - h(f(\cdot,Y))
1256: \Big|_{L^1_T}\\
1257: &=& I_{21}+I_{22},
1258: \end{eqnarray*}
1259: where $L^1_T$ stands for $L^1(\Omega\times[0,T],d\mathbb{P}\times
1260: dt)$.  Recalling again that mollification does not increase the
1261: Lipschitz constant, we also have
1262: $$
1263: I_{21} \leq |h|_{\mathrm{Lip}} |f|_{\mathrm{Lip}}
1264: |Y_{\varepsilon_1} - Y|_{L^1_T}
1265: \leq |h|_{\mathrm{Lip}} |f|_{\mathrm{Lip}}
1266: \varepsilon_1 N \int_0^T e^{\int_0^t m(s)\,ds}\,dt.
1267: $$
1268: Finally, using again the uniform convergence on compact sets of
1269: mollified continuous functions,
1270: \[
1271: I_{22} \leq T \delta_3(\varepsilon_2).
1272: \qedhere
1273: \]
1274: \end{proof}
1275: 
1276: 
1277: 
1278: \subsection{An example with explicit solutions}     \label{subsec:exex}
1279: In this subsection we study in detail the optimal advertising problem
1280: with linear reward and quadratic cost. In particular, we shall assume
1281: $h(z)=-\beta z_0^2$ and $\varphi(x) = \gamma x_0$, with $\beta$,
1282: $\gamma>0$. In \cite{levico} we proved that a solution (in integral
1283: sense) of the HJB equation (\ref{eq:HJB0}) is of the type
1284: $$
1285: v(t,x) = \cp{w(t)}{x} + c(t), \qquad t \in [0,T], \; x \in X,
1286: $$
1287: where $w=(w_0,w_1):[0,T]\to X$ and $c:[0,T]\to\erre$ are given by
1288: \begin{equation}\label{eq:HJBf3}
1289: \left\{
1290: \begin{array}{ll}
1291: \ds w_0'(t) + a_0w_0(t) + \int_{-r}^0 a_1(\xi)w_1(t,\xi)\,d\xi = 0, &
1292: t \in [0,T[ \\[8pt]
1293: w_0(T)=\gamma,\\[8pt]
1294: w_1(t,\xi) = w_0(t-\xi)\chi_{[0,T]}(t-\xi),\\[8pt]
1295: \displaystyle c(t) = \int_t^T \frac{(\cp{B}{w(s)}^+)^2}{4\beta}\,ds,
1296: & t \in [0,T].
1297: \end{array}\right.
1298: \end{equation}
1299: Moreover, the optimal strategy is
1300: \begin{equation}
1301: \label{eq:22}
1302: z^*(t) = \frac{\cp{B}{v_x(t)}^+}{2\beta} =
1303: \frac{\cp{B}{w(t)}^+}{2\beta}, \qquad t \in [0,T]
1304: \end{equation}
1305: (see \cite{levico} for more details).
1306: 
1307: We extend now the analysis of this specific situation. Let us begin
1308: with a rather explicit characterization of the optimal trajectory,
1309: which could be numerically approximated simply by solving a linear ODE
1310: with delay.
1311: %
1312: In particular, let $w=(w_0,w_1)$ be the solution of (\ref{eq:HJBf3}).
1313: Then, setting $z^*(t)={1\over 2\beta}\cp{B}{w(t)}^+$, the optimal
1314: trajectory is the $\erre$-valued component $Y_0$ of the (mild)
1315: solution of the abstract SDE
1316: $$ dY(t)=[AY(t) + Bz^*(t)]\,dt + G\,dW(t), $$
1317: which is given by
1318: $$
1319: Y(t) = e^{tA}Y(0) + \int_0^t e^{(t-s)A}Bz^*(s)\,ds
1320: + \int_0^t e^{(t-s)A}G\,dW(s).
1321: $$
1322: In particular $Y$ is a $X$-valued Gaussian process with mean
1323: and covariance operator
1324: $$
1325: \mu_t = e^{tA}Y(0) + \int_0^t e^{(t-s)A}Bz^*(s)\,ds ,
1326: \quad
1327: Q_t = \int_0^t e^{(t-s)A}GG^*e^{(t-s)A^*}\,ds,
1328: $$
1329: respectively.
1330: It follows that $Y_0$ is a Gaussian process itself with mean
1331: \begin{eqnarray}
1332: \E Y_0(t) &=& \cp{\mu_t}{e_1}_X =
1333: \left\langle e^{tA}Y(0),e_1\right\rangle_X
1334: + \left\langle \int_0^t e^{(t-s)A}Bz^*(s)\,ds,e_1\right\rangle_X \nonumber\\
1335: \label{eq:opttraj}
1336: &=& \left\langle Y(0),e^{tA^*}e_1\right\rangle_X
1337: + \int_0^t\left\langle Bz^*(s),e^{(t-s)A^*}e_1\right\rangle_X ds,
1338: \end{eqnarray}
1339: where $e_1=(1,0)\in X$.\\
1340: %
1341: Since $Y(0)$ is given as in Proposition \ref{prop:equiv} and
1342: $Bz^*(\cdot)$ is also easy to compute ($z^*$ is one dimensional and
1343: $B$ is just multiplication by a fixed vector in $X$), we are left with
1344: the problem of computing $e^{tA^*}e_1$. However, as one can prove by a
1345: direct calculation, the semigroup $e^{tA^*}$ is given by
1346: $$
1347: e^{tA^*}(x_0,x_1(\cdot)) = \Big(\phi(t),\phi(t+\xi)|_{\xi \in [-r,0]}\Big),
1348: $$
1349: where $\phi(\cdot)$ solves the linear ODE with delay
1350: \begin{equation}
1351: \label{eq:sgdelay}
1352: \left\{\begin{array}{l}
1353: \ds {d\phi(t)\over dt} =
1354: a_0 \phi(t) + \int_{-r}^0a_1(\xi)\phi(t+\xi)\,d\xi,
1355: \quad 0\leq t\leq T \\[10pt]
1356: \phi(0)=x_0; \quad \phi(\xi)=x_1(\xi) \;\; \forall\xi\in[-r,0].
1357: \end{array}\right.
1358: \end{equation}
1359: %
1360: Therefore $e^{tA^*}e_1$ is given by $(\phi(t),\phi(t+\xi)|_{\xi \in
1361:   [-r,0]})$, where $\phi$ solves (\ref{eq:sgdelay}) with initial
1362: condition $x_0=1$, $x_1(\cdot)=0$. Such $\phi$ can be computed
1363: numerically by discretizing (\ref{eq:sgdelay}), and then $\E Y_0(t)$
1364: can be obtained by approximating the integrals in (\ref{eq:opttraj})
1365: with finite sums.
1366: 
1367: Analogously one can write the variance of the optimal trajectory in
1368: such a way that it can be easily approximated by numerical methods. In
1369: particular, one has
1370: \begin{eqnarray}
1371: \mathrm{Var}\,Y_0(t) &=& \cp{Q_te_1}{e_1}
1372: = \left\langle
1373: \int_0^t e^{(t-s)A}GG^*e^{(t-s)A^*}\,ds\, e_1,e_1
1374: \right\rangle \nonumber\\
1375: &=& \int_0^t
1376: \left\langle e^{(t-s)A}GG^*e^{(t-s)A^*} e_1,e_1 \right\rangle ds \nonumber\\
1377: &=& \int_0^t
1378: \left\langle Ge^{(t-s)A^*}e_1,Ge^{(t-s)A^*}e_1 \right\rangle ds \nonumber\\
1379: &=& \int_0^t \left|Ge^{(t-s)A^*}e_1\right|^2\,ds.\label{eq:optvar}
1380: \end{eqnarray}
1381: Setting $\psi(s)=e^{(t-s)A^*}e_1$, which can be approximated as
1382: indicated before, one finally has
1383: $$
1384: \mathrm{Var}\,Y_0(t) = \sigma^2 \int_0^t \psi(s)^2\,ds.
1385: $$
1386: 
1387: \medskip
1388: 
1389: One can also perform simple comparative statics on the
1390: value function. For instance we can compute explicitly its sensitivity
1391: with respect to the (maximal) delay $r$:
1392: \begin{eqnarray*}
1393: \frac{\partial V}{\partial r}(t,x;r) &=&
1394: {\partial \over \partial r}\cp{w_1(t)}{x_1}_{L^2([-r,0])} +
1395: \frac{\partial c}{\partial r}(t;r) \\
1396: &=& w_1(t,-r)x_1(-r)
1397: + {1 \over 2\beta}
1398: \int_t^T \cp{B}{w(s)}
1399: {\partial\over\partial r}\left(\int_{-r}^0 b_1(\xi)w_1(s,\xi)\,d\xi\right)\,ds \\
1400: &=& w_1(t,-r)x_1(-r) + {b_1(-r) \over 2\beta}
1401: \int_t^T \cp{B}{w(s)} w_1(s,-r)\,ds,
1402: \end{eqnarray*}
1403: where we have used the fact that $\cp{B}{w(t)}^+=\cp{B}{w(t)}$.
1404: \par\noindent
1405: Note that in the above expression everything can be computed
1406: explicitly, as soon as we fix the delay kernel $b_1$.
1407: %
1408: Let us consider, as an example, the special case of
1409: $b_1(\xi)=b_1\chi_{[-r,0]}(\xi)$, where on the right-hand side, with a
1410: slight abuse of notation, $b_1$ is a positive constant. One has
1411: \begin{eqnarray*}
1412: \frac{\partial V}{\partial r}(t,x;r) &=& w_1(t,-r)x_1(-r) +
1413: {b_1 \over 2\beta} \int_t^T w_1(s,-r)
1414: \Big(b_0w_0(s) + b_1\int_{-r}^0w_1(s,\xi)\,d\xi\Big)\,ds.
1415: \end{eqnarray*}
1416: %
1417: Furthermore, if we consider the special case of delay in the control
1418: only, that is $a_1(\cdot)=0$, we obtain, after some calculations,
1419: \begin{eqnarray*}
1420: \frac{\partial V}{\partial r}(t,x;r) &=&
1421: \gamma e^{a_0(T-t+r)}x_1(-r)
1422: - \frac{b_1}{4\beta a_0}\gamma^2e^{a_0r}%
1423: \Big(b_0-{b_1\over a_0}(1-e^{a_0r})\Big)%
1424: (1-e^{2a_0(T-t)}),
1425: \end{eqnarray*}
1426: for $t \in [r,T]$.
1427: 
1428: 
1429: %\subsubsection{Delay in the control only}
1430: 
1431: \medskip
1432: 
1433: In the special case of $a_1(\cdot)\equiv 0$ an explicit solution of
1434: (\ref{eq:HJBf3}) is easily obtained. This solvability in closed form
1435: then ``propagates'' to other quantities of interest. In fact, note
1436: that (\ref{eq:HJBf3}) reduces to
1437: \begin{equation}%\label{eq:HJBf5}
1438: \left\{
1439: \begin{array}{l}
1440: \ds w_0'(t) + a_0w_0(t) = 0 \\[8pt]
1441: w_0(T)=\gamma,
1442: \end{array}\right.
1443: \end{equation}
1444: yielding
1445: $$ w_0(t) = \gamma e^{(T-t)a_0}, $$
1446: and therefore
1447: $$
1448: w_1(t,\xi) = \gamma e^{(T-(t+\xi))a_0}\,\chi_{[0,T]}(t+\xi)
1449: \quad
1450: c(t) = \int_t^T {\cp{B}{w(s)}^2 \over 4\beta}\,ds.
1451: $$
1452: That is, the last three formulae explicitly give a solution of the
1453: HJB equation (\ref{eq:HJB0}) in our specific case.
1454: 
1455: As a consequence we can also determine the unique optimal feedback control
1456: $z^*$ as follows:
1457: \begin{equation}
1458: \label{eq:27}
1459: z^*(t) = {\cp{B}{v_x}^+ \over 2\beta} = {\cp{B}{w(t)}^+ \over 2\beta}
1460: = {\gamma e^{(T-t)a_0} \over 2\beta} \left[%
1461: b_0 + \int_{-r}^0 b_1(\xi)e^{-a_0\xi}\chi_{[0,T]}(t+\xi)\,d\xi \right].
1462: \end{equation}
1463: 
1464: The optimal trajectory can be characterized in a completely similar
1465: way as above, with the difference that now we can explicitly write:
1466: $$
1467: e^{tA^*}e_1 = (e^{a_0t},e^{a_0t+\xi}|_{\xi\in[-r,0]}),
1468: $$
1469: hence simplifying (\ref{eq:opttraj}) in the present case. Even
1470: simpler is the expression for the variance of the optimal trajectory,
1471: which can be obtained by (\ref{eq:optvar}):
1472: $$
1473: \mathrm{Var}\,Y_0(t) = \sigma^2 \int_0^t e^{2a_0(t-s)}\,ds
1474: =\frac{\sigma^2}{2a_0} (e^{2a_0t}-1).
1475: $$
1476: 
1477: \begin{figure}[t]
1478: \includegraphics[width=\textwidth]{Fig1.eps}
1479: \caption{Optimal advertising policy in four different ``churn'' settings}
1480: \label{fig:1}
1481: \end{figure}
1482: 
1483: Sharp characterizations of the optimal advertising trajectory as
1484: well as the resulting expected profit functions in the case of
1485: linear reward and quadratic cost function allow for interesting
1486: observations regarding the importance of the proper accounting for
1487: the memory effects in planning the advertising campaign. Figure
1488: \ref{fig:1} displays the optimal advertising spending rates
1489: $z^*(t)$, as expressed by (\ref{eq:22}), with
1490: $a_1(\xi)=\hat{a}_1e^{-|\xi|/\delta_a}$ and
1491: $b_1(\xi)=\hat{b}_1e^{-|\xi|/\delta_b}$ in four different settings:
1492: $a_1 = b_1 = 0$ (``no churn''), $a_1=-5$, $b_1=0$ (``goodwill
1493: churn''), $a_1=0$, $b_1=5$ (``advertising churn''), $a_1=-5$,
1494: $b_1=5$ (``goodwill-advertising churn''). Note that in the absence
1495: of churn, the optimal advertising trajectory, as implied by
1496: (\ref{eq:27}), is a monotone function of time with $z^*(t)=\gamma
1497: b_0/(2\beta)$. While the advertising rates are similar in all
1498: settings in the beginning of the pre-launch period as well as right
1499: before the product launch time $T$, the details of advertising
1500: policies differ dramatically in the middle of the pre-launch period.
1501: For example, in the presence of a strong ``goodwill churn'' the
1502: optimal advertising trajectory takes a characteristic ``impulse''
1503: shape, while in the strong ``advertising churn'' setting the optimal
1504: advertising spending quickly builds up a strong goodwill level in
1505: the middle of the pre-launch region, slowing down significantly
1506: right before the product launch. When the presence of both types of
1507: ``churn'' is pronounced, the optimal advertising policy is
1508: represented by a set of advertising sprees with rapidly growing
1509: intensity. Figure \ref{fig:2} illustrates how the strong influence
1510: of memory effects on the shape of optimal advertising policies
1511: translates into performance differences between the optimal
1512: advertising policies and the policies which neglect the presence of
1513: advertising delays. In this figure we plot the relative difference
1514: $$
1515: \frac{V(0,x)-V^0(0,x)}{V(0,x)}
1516: $$
1517: between the optimal expected profit function $V(0,x)$ and the expected
1518: profit value $V^0(0,x)$ obtained by applying, for $t\in [0,T]$, the
1519: advertising policy $z^0(t)=\gamma b_0e^{(T-t)a_0}/(2\beta)$ optimal in
1520: the absence of memory effects (i.e. in the setting $a_1=b_1=0$).  This
1521: relative difference is plotted as a function of the amplitude of the
1522: ``goodwill churn'' term $a_1$ (Figure \ref{fig:2}a), and as a function
1523: of the amplitude of the ``advertising churn'' term $b_1$ (Figure
1524: \ref{fig:2}b). The initial goodwill conditions were selected as
1525: $x_0=10$ and $x_1(\xi)= x_0 e^{-|\xi|}$ for $\xi \in [-r,0]$ and the
1526: advertising history $z_0(\xi)$ was set equal to $0$ for
1527: $\xi\in[-r,0]$. We observe that the relative loss of efficiency
1528: associated with the use of the ``memoryless'' policy $z^0(t)$ can be
1529: quite significant -- in the examples we use it exceeds 5\% and can be
1530: as high as 20\% in settings with strong ``churn'' effects.
1531: 
1532: \begin{figure}[t]
1533: \includegraphics[width=\textwidth]{Fig2.eps}
1534: \caption{Relative performance gap of the ``memoryless'' advertising policy}
1535: \label{fig:2}
1536: \end{figure}
1537: 
1538: 
1539: 
1540: \section{The case of delay in the state term only}
1541: \label{sec:ex}
1542: 
1543: In this section we consider a model for the dynamics of goodwill with
1544: forgetting, but without lags in the effect of advertising expenditure
1545: (carryover), i.e. with $b_1(\cdot)=0$ in (\ref{eq:SDDE}). An analysis
1546: of this model was sketched in \cite{levico}, where only an abstract
1547: existence result was given. Here we present a more refined result (see
1548: theorem \ref{thm:L2} below) and obtain some qualitative properties of
1549: the value function, together with a characterization of the optimal
1550: strategies in terms of the value function in two specific cases. In
1551: particular, theorem \ref{thm:L2} formulates sufficient conditions
1552: ensuring that the value function is the unique solution (in a suitable
1553: sense) of the associated Bellman equation and that the optimal
1554: advertising policy is of the feedback type. Let us recall once again
1555: that such a situation is not possible in the more general case
1556: discussed in the previous section. Moreover, in the case of linear
1557: cost function, the optimal control takes a particularly simple
1558: ``bang-bang'' form (Corollary \ref{cor:bang}).
1559: 
1560: Let us also mention that for stochastic control problem with delay
1561: terms in the state variable one can apply both the approach of
1562: Hamilton-Jacobi equations in $L^2$ spaces developed by Goldys and
1563: Gozzi \cite{GolGoz}, and the forward-backward SDE approach of
1564: Fuhrman and Tessitore \cite{FT1}.  We follow here the first approach,
1565: showing that both the value function and the optimal advertising
1566: policy can be characterized in terms of the solution of a Bellman
1567: equation in infinite dimensions.
1568: 
1569: We assume, for the sake of simplicity, that the goodwill evolves
1570: according to the following equation, where the distribution of the
1571: forgetting factor is concentrated on a point:
1572: \begin{equation}
1573: \label{eq:ex1}
1574: \left\{\begin{array}{l}
1575: dy(t) = \ds [a_0 y(t) + a_1y(t-r)
1576:     + b_0 z(t)]\,dt + \sigma\,dW_0(t), \quad 0\leq s\leq t\leq T \\[10pt]
1577: y(s)=x_0; \quad y(s+\xi)=x_1(\xi), \;\; \xi\in[-r,0].
1578: \end{array}\right.
1579: \end{equation}
1580: %
1581: The following standard infinite dimensional Markovian reformulation of
1582: this dynamics will turn out to be useful.
1583: 
1584: Let us define the operator $\A:D(\A)\subset X \to X$ as
1585: $$
1586: \A:(x_0,x_1(\cdot)) \mapsto (a_0x_0+a_1x_1(-r),x_1'(\cdot)),
1587: \qquad
1588: D(\A) = \erre\times W^{1,2}([-r,0];\erre).
1589: $$
1590: It is well-known (see e.g. \cite{DZ96}) that $\A$ is the generator of
1591: a strongly continuous semigroup $S(t)$ on $X$. More precisely, one has
1592: $$
1593: S(t)(x_0,x_1) = (u(t),u(t+\xi)|_{\xi\in[-r,0]}),
1594: $$
1595: where $u(\cdot)$ is the solution of the deterministic delay equation
1596: \begin{equation}
1597: \label{eq:sgA}
1598: \left\{\begin{array}{l}
1599: \ds {du(t)\over dt} = a_0 u(t) + a_1 u(t-r), \quad 0\leq t\leq T \\[10pt]
1600: u(0)=x_0; \quad u(\xi)=x_1(\xi), \;\; \xi\in[-r,0].
1601: \end{array}\right.
1602: \end{equation}
1603: %
1604: Furthermore, set $\bar{z}=(\sigma^{-1}b_0 z,z_1(\cdot))$, with
1605: $z_1(\cdot)$ a fictitious control taking values in
1606: $L^2([-r,0],\erre)$, and
1607: define $G:X \to X$ as
1608: $$
1609: G: (x_0,x_1(\cdot)) \mapsto (\sigma x_0,0).
1610: $$
1611: Let $W_1$ be a cylindrical Wiener process taking
1612: values in $L^2([-r,0],\erre)$, so that $W=(W_0,W_1)$ is an $X$-valued
1613: cylindrical Wiener process.
1614: 
1615: Chojnowska-Michalik \cite{choj78} proved the following equivalence result.
1616: \begin{lemma}
1617: \label{lem:equiv}
1618:   Let $Y=(Y_0,Y_1)$ be the unique mild solution of the following stochastic
1619:   evolution equation on $X$:
1620: \begin{equation}
1621: \label{eq:ex1a}
1622: \left\{\begin{array}{l}
1623: dY(t) = (\A Y(t) + G\bar{z}(t))\,dt + G\,dW(t),
1624:               \quad 0\leq s \leq t \leq T,\\[10pt]
1625: Y(s) = x.
1626: \end{array}\right.
1627: \end{equation}
1628: Then $Y_0(t)$ solves the stochastic delay equation (\ref{eq:ex1}).
1629: \end{lemma}
1630: %
1631: Define $h:X\to\erre$ and $\varphi:X\to\erre$ as
1632: \begin{eqnarray*}
1633: h(x_0,x_1) &=& -h_0(\sigma b_0^{-1}x_0) \\
1634: \varphi(x_0,x_1) &=& \varphi_0(x_0).
1635: \end{eqnarray*}
1636: Then we have, thanks to lemma \ref{lem:equiv},
1637: $$
1638: J(s,x;z) =
1639: \E_{s,x}\left[\varphi(Y(T))+\int_t^T h(\bar{z}(s))\,ds \right],
1640: $$
1641: and
1642: \begin{equation}
1643: \label{ex1j}
1644: V(s,x) = \sup_{\bar{z}\in\mathcal{Z}} J(s,x;\bar{z}),
1645: \end{equation}
1646: where $\mathcal{Z}$ denotes the set of all strategies $z:[0,T]\to
1647: \tilde{U} \times L^2([-r,0],\erre)$ adapted to the filtration
1648: generated by $Y$, and $\tilde{U}$ is the image of $U$ under the action
1649: of the map $x\mapsto \sigma^{-1}b_0x$.
1650: 
1651: \smallskip
1652: 
1653: We can now prove some qualitative properties of the value function.
1654: \begin{prop}\label{prop:V-conv}
1655:   If $\varphi_0$ is concave and $h_0$ is convex, then the value
1656:   function $V(s,x)$ is concave with respect to $x$.
1657: \end{prop}
1658: %
1659: \begin{proof}
1660: Identical to the proof of proposition \ref{prop:Vconv}, thus omitted.
1661: \end{proof}
1662: %
1663: In the following proposition we use the ordering in $X$ defined right before
1664: Proposition \ref{prop:Vincr}.
1665: %
1666: \begin{prop}\label{prop:V-incr}
1667:   Let $a_1 \geq 0$ and $\varphi_0$ be increasing. Then the value
1668:   function $V(s,x)$ is increasing with respect to $x$. Moreover, if
1669:   $\varphi_0$ is strictly increasing, then the value function $V(s,x)$
1670:   is strictly increasing with respect to $x$, and $V(s,x^1)=V(s,x^2)$
1671:   if and only if $x^1=x^2$.
1672: \end{prop}
1673: %
1674: \begin{proof}
1675: Let $x^1\geq x^2$ in the sense just defined. One has
1676: \begin{eqnarray*}
1677: J(s,x^1;z)-J(s,x^2;z) &=&
1678: \E\Big[ \varphi(y^{s,x^1,z}(T)) - \varphi(y^{s,x^2,z}(T)) \Big] \\
1679: &=& \E\Big[ \varphi(e^{\A(T-s)}x^1 + \zeta) -
1680:             \varphi(e^{\A(T-s)}x^2 + \zeta)
1681:       \Big],
1682: \end{eqnarray*}
1683: where $\zeta:=\int_s^T e^{\A(T-t)}G\bar{z}(t)\,dt+%
1684: \int_s^T e^{\A(T-t)}G\,dW(t)$. The assumption $a_1\geq 0$
1685: together with (\ref{eq:sgA}) implies that the semigroup generated by
1686: $\A$ is positivity preserving, i.e.  $x^1\geq x^2$ implies
1687: $e^{\A(T-s)}x^1\geq e^{\A(T-s)}x^2$.  Therefore, by the monotonicity of
1688: $\varphi_0$, one also has $\varphi(e^{\A(T-s)}x^1 + \zeta) \geq
1689: \varphi(e^{\A(T-s)}x^2 + \zeta)$ a.s., hence $J(s,x^1;z) \geq
1690: J(s,x^2;z)$, and finally $V(s,x^1)=\sup_{z\in\mathcal{Z}} J(s,x^1;z)
1691: \geq \sup_{z\in\mathcal{Z}} J(s,x^2;z)=V(s,x^2)$. The other assertions
1692: follow analogously, using (\ref{eq:sgA}).
1693: \end{proof}
1694: %
1695: \begin{rmk}
1696:   In the above proof the positivity preserving property of the
1697:   semigroup $e^{t\A}$ is crucial, and the assumption $a_1\geq 0$ is
1698:   ``sharp'' in the following sense: if $a_1<0$, one can find $x>0$
1699:   such that $e^{t\A}$ \emph{inverts} the sign, i.e. $e^{t\A}x<0$.
1700: 
1701:   Moreover, under the assumptions of the theorem, the value function
1702:   is increasing with respect to the real valued component of the
1703:   initial datum. By this we mean that given $x^1 \geq x^2$ with
1704:   $x^1_0>x^2_0$ and $x^1_1=x^2_1$ a.e., then $V(s,x^1)>V(s,x^2)$.
1705:   Therefore one also has $D^-_0V\geq 0$. The subdifferential can be
1706:   replaced by the derivative if we can guarantee that $V$ is
1707:   continuously differentiable with respect to $x_0$. Conditions for
1708:   the continuous differentiability of $V$ with respect to $x$ are
1709:   given in proposition \ref{prop:Vdiff} below.
1710: \end{rmk}
1711: 
1712: In contrast with the general case considered in the previous section,
1713: if delay enters only the state term, then it is possible to uniquely
1714: solve the associated Bellman equation, and thus to characterize the
1715: value function and construct optimal strategies.  The following
1716: result, which relies on \cite{GolGoz}, gives precise conditions for
1717: the above assertions to hold.
1718: %
1719: \begin{thm}\label{thm:L2}
1720:   Assume that $h_0$ is convex, let $H_0:\erre\ni p \mapsto
1721:   \sup_{u\in U}(pb_0u-h_0(u))$, and suppose that $a_0 < -a_1 <
1722:   \sqrt{\gamma^2+a_0^2}$, where $\gamma\,\mathrm{coth}\,\gamma = a_0$,
1723:   $\gamma\in]0,\pi[$. Then the value function $V(s,x)$ coincides
1724:   $\mu$-a.e. with the mild solution in $L^2(X,\mu)$ (in the sense of
1725:   \cite{GolGoz}) of the equation
1726: \begin{equation}
1727: \label{eq:ex1-hjb-expl}
1728: \left\{\begin{array}{l}
1729: \ds \partial_t v
1730: + {1\over 2}\sigma^2 \partial_0^2 v
1731: + (a_0x_0 + a_1x_1(-r)) \partial_0 v
1732: + \int_{-r}^0 x'_1(\xi) \partial_1 v(\xi)\,d\xi
1733: + H_0(\partial_0 v) = 0 \\[10pt]
1734: v(T,x_0,x_1) = \varphi_0(x_0),
1735: \end{array}\right.
1736: \end{equation}
1737: where $\mu$ is a measure of full support on $X$.
1738: %
1739: Moreover, the optimal strategy admits the feedback representation
1740: \begin{equation}
1741: \label{ex1-zopt}
1742: z^*(t) \in \sigma b_0^{-1}D^-_0H( \sigma\partial_0 v(t,Y^*_0(t),Y_1^*(t) ),
1743: \end{equation}
1744: with $H$ defined in (\ref{eq:h}), provided there exists a solution
1745: $Y_0^*(t)$, $Y_1^*(t)$ of the closed-loop differential inclusion
1746: \begin{equation}
1747: \label{eq:di}
1748: \left\{\begin{array}{l}
1749: \ds dY_0(t) \in \left[a_0Y_0(t) + a_1Y_0(t-r) +
1750: \sigma D^-_0H(\sigma \partial_0 v(t,Y^*_0(t),Y_1^*(t))
1751: \right]dt + \sigma\,dW_0(t) \\[10pt]
1752: \ds dY_1(t)(\xi) = \frac{d}{d\xi}Y_1(t)(\xi).
1753: \end{array}\right.
1754: \end{equation}
1755: \end{thm}
1756: %
1757: \begin{proof}
1758: By the usual heuristic application of the dynamic programming principle
1759: one can associate to the control problem (\ref{ex1j}) the following
1760: Hamilton-Jacobi-Bellman equation on $X$:
1761: \begin{equation}
1762: \label{ex1hjb}
1763: \left\{\begin{array}{ll}
1764: \ds v_t + {1\over 2}\tr(GG^*v_{xx}) + \cp{\A x}{v_x} + H_0(v_x) = 0,
1765: \quad 0 \leq t \leq T, \\[10pt]
1766: v(T,x) = \varphi(x),
1767: \end{array}\right.
1768: \end{equation}
1769: which coincides, after some calculations, with (\ref{eq:ex1-hjb-expl}).
1770: %
1771: Note that the Hamiltonian $H_0$ can be regarded as a function of $X$
1772: in $\erre$ and can be equivalently written as
1773: \begin{eqnarray}
1774: H_0(p) = H_0(p_0) &=&
1775: \sup_{z\in \tilde{U}\times L^2([-r,0],\erre)}
1776: \Big(\cp{Gz}{p} + h(z)\Big) \nonumber \\
1777: &=& \sup_{z_0\in \tilde{U}}
1778: \Big(\sigma z_0p_0 - h_0(\sigma b_0^{-1}z_0)\Big).
1779: \label{eq:h0}
1780: \end{eqnarray}
1781: In order to apply the results of \cite{GolGoz}, we also need to define
1782: \begin{equation}
1783: \label{eq:h}
1784: H(q)=H(q_0)=H_0(\sigma^{-1}q_0) =
1785: \sup_{z_0\in\tilde{U}} \Big(z_0q_0 - h_0(\sigma b_0^{-1}z_0)\Big).
1786: \end{equation}
1787: %
1788: Since $h_0$ is bounded from below, (\ref{eq:h}) implies that $H$ is
1789: Lipschitz continuous (in $\erre$ and in $X$). Moreover, the assumption
1790: on $a_0$, $a_1$ and assumption (i) of section \ref{sec:formul} imply
1791: that the uncontrolled version of (\ref{eq:ex1a}), i.e.
1792: \begin{equation}
1793: \label{eq:ex1auc}
1794: dY(t) = AY(t)\,dt + G\,dW(t),
1795: \end{equation}
1796: admits a unique non-degenerate invariant measure $\mu$ on $X$ (see
1797: \cite{DZ96}), which is Gaussian with mean zero and covariance
1798: operator $Q_\infty=\int_0^\infty e^{sA}GG^*e^{s^*A}\,ds$. In
1799: particular, the restriction of $\mu$ on the $\erre$-valued component
1800: of $X$ has a density
1801: $\rho(x)=\frac{1}{\nu\sqrt{2\pi}}e^{-|x|^2/2\nu^2}$ for some $\nu>0$.
1802: This implies that $\varphi\in L^2(X,\mu)$: in fact,
1803: $$
1804: \int_X |\varphi(x)|^2\,\mu(dx) =
1805: \int_\erre |\varphi_0(x)|^2\rho(x)\,dx \leq
1806: K \int_\erre (1+|x|)^m e^{-|x|^2/2\nu^2}\,dx < \infty.
1807: $$
1808: Therefore, theorems 3.7 and 5.7 of
1809: \cite{GolGoz} yield the existence and uniqueness of a solution in
1810: $L^2(X,\mu)$ of (\ref{ex1hjb}), or equivalently of
1811: (\ref{eq:ex1-hjb-expl}), which coincides $\mu$-a.e. with the value
1812: function $V$.
1813: %
1814: Finally, observing that the maximum in (\ref{eq:h}) is reached by
1815: $D^-_0H(q_0)$ (setting, if needed, $h_0(x)=+\infty$ for $x\not\in U$),
1816: a slight modification of the proof of theorem 5.7 in \cite{GolGoz}
1817: shows that the optimal strategy is given by $\bar{z}_0^*(t) \in
1818: D^-_0H(\sigma\partial_0 v(t,Y^*_0(t),Y_1^*(t))$, where
1819: $Y^*=(Y_0^*,Y_1^*)$ is a solution (if any) of the stochastic
1820: differential inclusion (\ref{eq:di}). The relation $z^*(t)=\sigma
1821: b_0^{-1}\bar{z}_0^*(t)$ thus completes the proof.
1822: \end{proof}
1823: %
1824: Let us briefly comment on the previous result: the HJB equation
1825: (\ref{eq:ex1-hjb-expl}) is ``genuinely'' infinite dimensional, i.e. it
1826: reduces to a finite dimensional one only in very special cases. For
1827: example, by the results in \cite{lari}, (\ref{eq:ex1-hjb-expl})
1828: reduces to a finite dimensional PDE if and only if $a_0=-a_1$.
1829: However, under this assumption, we cannot guarantee the existence of a
1830: non-degenerate invariant measure for the Ornstein-Uhlenbeck semigroup
1831: associated to (\ref{eq:ex1auc}). Even more extreme would be the
1832: situation of distributed forgetting time: in this case the HJB
1833: equation is finite dimensional only if the term accounting for
1834: distributed forgetting vanishes altogether.
1835: %
1836: Moreover, note that if $a_1$ is negative, i.e. it can be interpreted
1837: as a deterioration factor, the assumption of the theorem says that
1838: $a_1$ cannot be ``much more negative'' than $a_0$. On the other hand,
1839: if $a_1$ is positive, then the improvement effect as measured by $a_1$
1840: cannot exceed the deterioration effect as measured by $|a_0|$.  In
1841: essence, the condition on $a_0$, $a_1$, which is needed to ensure
1842: existence of an invariant measure for equation (\ref{eq:ex1auc}), does
1843: not impose severe restrictions on the dynamics of goodwill.
1844: 
1845: \smallskip
1846: 
1847: If the data of the problem are smoother, a different approach allows
1848: one to obtain regularity of the value function.
1849: \begin{prop}\label{prop:Vdiff}
1850:   Assume that $\varphi_0\in C^1(\erre)$,
1851:   $|\varphi_0'(x)|\leq K(1+|x|)^m$, and $H_0\in C^1(\erre)$.
1852:   Then $V\in C^{0,1}([0,T]\times X)$.
1853: \end{prop}
1854: \begin{proof}
1855:   Follows by the regularity results for solutions of semilinear
1856:   partial differential equations in Hilbert spaces obtained through
1857:   the FBSDE approach. In particular, denoting by $C$ a positive
1858:   constant, boundedness of $U$ implies that $|b_0 z|<C$, $|h_0(z)|
1859:   \leq K(1+|z|)^m < C$, and finally $H_0$ is Lipschitz as follows by
1860:   \begin{eqnarray*}
1861:     |H_0(p)-H_0(q)| &=& |\sup_{z\in U}(\cp{Bp}{z}-h_0(z))
1862:                        - \sup_{z\in U}(\cp{Bp}{z}-h_0(z))| \\
1863:     &\leq& \sup_{z\in U} |\cp{(p-q)}{B^*z}| \leq C|p-q|.
1864:   \end{eqnarray*}
1865:   Then all hypotheses of \cite{FT-corso}, theorem 4.3.1, are
1866:   satisfied, which yields the claim.
1867: \end{proof}
1868: %
1869: %In the following corollary we do not need $h_0$ to satisfy a
1870: %polynomial growth condition.
1871: \begin{coroll}\label{cor:Vdiff}
1872:   Let $\varphi_0$ be as in proposition \ref{prop:Vdiff} and $h_0$
1873:   strictly convex. Then $V\in C^{0,1}([0,T]\times X)$.
1874: \end{coroll}
1875: \begin{proof}
1876:   Since convexity implies continuity in the interior of the domain,
1877:   then $h_0(U)$ is bounded. Extending $h_0$ as $h_0(x)=+\infty$ for
1878:   $x\not\in U$, $h_0$ is clearly $1$-coercive, hence $H$ is convex and
1879:   finite on the whole $\erre$ (\cite{lema}, prop. E.1.3.8). The strict
1880:   convexity of $h_0$ implies that $H_0$ is continuously differentiable
1881:   in the interior of its domain, i.e. on $\erre$ (\cite{lema}, thm.
1882:   E.4.1.1). Then the smoothness of $V$ follows again by
1883:   \cite{FT-corso}, theorem 4.3.1.
1884: \end{proof}
1885: %
1886: %
1887: \begin{rmk}
1888:   We should also mention that in the framework of the FBSDE approach
1889:   to HJB equations (\cite{FT-corso}), if the Hamiltonian $H$ and the
1890:   terminal condition $\varphi$ satisfy some smoothness and boundedness
1891:   conditions, then we do not need the assumption about the existence
1892:   of an invariant measure for the uncontrolled state equation. The
1893:   approach used above (\cite{GolGoz}), while requiring the
1894:   existence of the above mentioned invariant measure, allows for more
1895:   singular data (for instance one could choose $\varphi_0(x)=-M$,
1896:   $M>\!\!>0$, for $x\in\erre_-$, and $\varphi_0(x)\geq 0$ for
1897:   $x\in\erre_+$).
1898: \end{rmk}
1899: 
1900: \medskip
1901: 
1902: In general, obtaining explicit expressions of the value function $V$
1903: trying to solve (\ref{eq:ex1-hjb-expl}) is impossible. However, under
1904: specific assumptions on the model we can obtain stronger
1905: characterizations, at least from a qualitative point of view, of the
1906: value function and/or of the optimal strategy.
1907: %
1908: \begin{coroll}
1909: Assume that $h_0(x)=\beta x^2$ and $U=[0,R]$, $R<\infty$. Then
1910: the optimal strategy is given by
1911: \begin{equation}
1912: \label{eq:L2-ex1}
1913: z^*(t) =
1914: \left\{\begin{array}{ll}
1915: \ds 0, & D_0V^* < 0 \\[8pt]
1916: \ds \frac{b_0D_0V^*}{2\beta}, & 0 \leq D_0V^* \leq
1917:       2b_0^{-1}\beta R \\[8pt]
1918: \ds R, & D_0V^* > 2b_0^{-1}\beta R,
1919: \end{array}\right.
1920: \end{equation}
1921: where $V^*:=V(t,Y_0^*(t),Y_1^*(t))$.
1922: \end{coroll}
1923: %
1924: \begin{proof}
1925: One has
1926: \begin{eqnarray*}
1927: H_0(p) &=& \sup_{0\leq z_0\leq \tilde{R}} \Big(\cp{Gp}{z}+h(z)\Big)
1928: = \sup_{0\leq z_0 \leq \tilde{R}} (\sigma p_0 z_0 - \tilde\beta z_0^2) \\
1929: &=& \ds \frac{(\sigma p_0)^2}{4\tilde\beta}
1930:         I_{\{0 \leq p_0 \leq \frac{2\tilde\beta\tilde{R}}{\sigma} \}}
1931:      + (\sigma p_0 \tilde{R} - \tilde\beta\tilde{R}^2)
1932:         I_{\{ p_0 > {2\tilde\beta\tilde{R} \over \sigma} \}}
1933: \end{eqnarray*}
1934: where $\tilde{R}=\sigma^{-1}b_0R$ and $\tilde\beta:=\sigma^2 b_0^{-2}\beta$.
1935: Therefore
1936: $$
1937: H(q) = q_0^2/4\tilde\beta \, I_{\{0 \leq q_0 \leq 2\tilde\beta\tilde{R} \}}
1938:        + (q_0 \tilde{R} - \tilde\beta\tilde{R}^2) \, I_{\{q_0 > 2\tilde\beta\tilde{R}\}}
1939: $$
1940: and
1941: $$
1942: DH(q) = q_0/2\tilde\beta I_{\{0 \leq q_0 \leq 2\tilde\beta\tilde{R}\}}
1943: + \tilde{R} I_{\{ q_0 > 2\tilde\beta\tilde{R}\}}.
1944: $$
1945: %
1946: Theorem \ref{thm:L2} now yields (\ref{eq:L2-ex1}).
1947: \end{proof}
1948: %
1949: Note that whenever $\varphi_0$ is increasing, we get $D^-_0V^*\geq 0$, hence
1950: the optimal control is either linear in $D_0V^*$ or constant for
1951: $D_0V^*$ over a threshold.
1952: 
1953: \begin{coroll}     \label{cor:bang}
1954: Assume that $h_0(x)=\beta x$. Then the optimal strategy is of
1955: the bang-bang type and is given by
1956: $$
1957: z^*(t) =
1958: \left\{\begin{array}{ll}
1959: \ds 0, & D_0V^* < \sigma b_0^{-2}\beta \\[6pt]
1960: \ds \rho, & D_0V^* = \sigma b_0^{-2}\beta \\[6pt]
1961: \ds R, & D_0V^* > \sigma b_0^{-2}\beta,
1962: \end{array}\right.
1963: $$
1964: where $\rho$ is an arbitrary real number.
1965: \end{coroll}
1966: %
1967: \begin{proof}
1968: Setting $\tilde{R}=\sigma^{-1}b_0R$ and
1969: $\tilde\beta:=\sigma^2 b_0^{-2}\beta$, one has
1970: $$
1971: H_0(p) = \sup_{0\leq z_0 \leq \tilde{R}} (\sigma p_0 z_0 - \tilde\beta z_0)
1972: = (\sigma p_0-\tilde\beta)\tilde{R} \, I_{\{p_0 > \tilde\beta/\sigma\}}
1973: $$
1974: and $H(q) = (q_0-\tilde\beta)\tilde{R} \, I_{\{q_0 > \tilde\beta\}}$,
1975: thus
1976: $$
1977: D^-H(q) = \erre \, I_{\{q_0 = \tilde\beta\}}
1978:           + \tilde{R}\, I_{\{q_0 > \tilde\beta\}}.
1979: $$
1980: Theorem \ref{thm:L2} yields the conclusion.
1981: \end{proof}
1982: 
1983: In general, even specifying a functional form of $h_0$, an explicit
1984: solution of the HJB equation for arbitrary $\varphi_0$ is not
1985: available, hence the above expressions of the optimal control strategy
1986: in terms of the value function and their corresponding qualitative
1987: properties are the ``best'' that one can expect, at least in the cases
1988: we have considered.
1989: 
1990: 
1991: 
1992: \section{Concluding remarks}
1993: 
1994: A number of deterministic advertising models allowing for delay effects have
1995: been proposed in the literature. However, the corresponding problems in the
1996: stochastic setting have not been investigated. One of the reasons is
1997: certainly that a theory of continuous-time stochastic control with delays
1998: has only been developed recently, following two approaches. The first
1999: approach is based on the solution of an associated infinite-dimensional
2000: Hamilton-Jacobi-Bellman equation in spaces of integrable functions (see
2001: \cite{GolGoz}). The other one relies on the analysis of an appropriate
2002: infinite-dimensional forward-backward stochastic differential equation (see
2003: \cite{FT1}). Both approaches, however, cannot be applied to problems with
2004: distributed lag in the effect of advertising.
2005: 
2006: Problems with memory effects in both the state and the control have
2007: been studied first by Vinter and Kwong \cite{VK} (in a deterministic
2008: LQ setting), and by Gozzi and Marinelli \cite{levico} (in the case of
2009: linear stochastic dynamics and general objective function). A general
2010: theory of solvability of corresponding HJB equations is currently not
2011: available, while an infinite-dimensional Markovian reformulation and a
2012: ``smooth'' verification theorem have been proved in \cite{levico}.
2013: 
2014: In this paper we have concentrated on deriving qualitative properties
2015: of the value function (such as convexity, monotonicity with respect to
2016: initial conditions, smoothness). For specific choices of the reward
2017: and cost functions, we obtain more explicit characterizations of value
2018: function and optimal state-control pair.
2019: 
2020: While our work makes a substantial initial step in the analysis of the
2021: stochastic advertising problems with delays, more remains to be done.
2022: Potential extensions of the present work include the analysis of problems
2023: with budget constraints as well as problems of advertising through multiple
2024: media outlets with different delay characteristics.
2025: 
2026: 
2027: \bibliographystyle{amsplain}
2028: \bibliography{ref}
2029: 
2030: 
2031: \end{document}
2032: