1 Introduction

It is well-known that in the classical Black-Scholes market, there is no arbitrage. However, if we include a singular term in the drift of the risky asset, it was first proved by Karatzas and Shreve [11] (Theorem B2, page 329), that arbitrages exist. Subsequently this type of market has been studied by several authors, including Jarrow and Protter [10]. They explain how a singular term in the drift can model a situation where the asset price is partially controlled by a large company which intervenes when the price is reaching a certain lower barrier, in order to prevent it from going below that barrier. They also prove that arbitrages can occur in such situations.

The purpose of our paper is to extend this study in two directions:

First, we introduce jumps in the market. More precisely, we study a jump diffusion market driven by a Brownian motion \(B(\cdot )\) and an independent compensated random measure \(\widetilde{N}(\cdot ,\cdot )\) with an added singular drift term, modelled by a local time of an underlying Lévy process \(Y(\cdot )\). In view of the unstable financial markets we have seen in recent years, and in particular during the economic crises in 2008 and the corona virus crisis this year, we think that jumps are useful in an attempt to obtain more realistic financial market models.

Introducing jumps in the stock price motion goes back to Cox and Ross [5] and to Merton [12].

Second, we assume that the trader only has access to a delayed information flow, represented by the filtration \(\mathcal {F}_{t-\theta }\), where \(\theta >0\) is the delay constant and \(\mathcal {F}_{t}\) is the sigma-algebra generated by both \(\{B(s)\}_{s\le t}\) and \(\{N(s,\cdot )\}_{s\le t}\). This extension is also motivated by the effort to get more realistic market models. Indeed, in all real-life markets there is delay in the information flow available, and traders are willing to pay to get the most recent price information. Especially, when trading with computers even fractions of seconds of delays are important. We represent the singular term by the local time of a given process and show that as long as \(\theta >0\) there is no arbitrage in this market. In fact, we show that this delayed market is viable, in the sense that the value of the optimal portfolio problem with logarithmic utility is finite. However, if the delay goes to 0, the value of the portfolio goes to infinity, at least under some additional assumptions.

We emphasize that our paper deals with delayed information flow, not delay in the coefficients in the model, as for example in the paper by Arriojas et al [2]. There are many papers on optimal stochastic control with delayed information flow, also by us. However, to the best of our knowledge the current paper is the first to discuss the effect of delay in the information flow on arbitrage opportunities in markets with a singular drift coefficient. We will show that by applying techniques from white noise theory we can obtain explicit results. Specifically, our model is the following:

Suppose we have a financial market with the following two investment possibilities:

  • A risk free investment (e.g. a bond or a (safe) bank account), whose unit price \(S_{0}(t)\) at time t is described by

    $$\begin{aligned} {\left\{ \begin{array}{ll} dS_{0}(t)=r(t)S_{0}(t)dt;\quad t \in [0, T],\\ S_{0}(0)=1. \end{array}\right. } \end{aligned}$$
    (1.1)
  • A risky investment, whose unit price S(t) at time t is given by a linear stochastic differential equation (SDE) of the form

    $$\begin{aligned} {\left\{ \begin{array}{ll} dS(t) &{} \!\!=\!S(t^{-})\left[ \mu (t)dt\!+\!\alpha (t)dL_{t}\!+\!\sigma (t)dB(t)\!+\!\int _{\mathbb {R} _{0}}\gamma (t,\zeta )\widetilde{N}(dt,d\zeta )\right] ; \quad t \in [0,T],\\ S(0) &{} >0, \end{array}\right. } \end{aligned}$$
    (1.2)

where \(\mathbb {R}_0=\mathbb {R} {\setminus } \{0\}.\) Here \(B(\cdot )\) and \(\widetilde{N}=N(dt,d\zeta )-\nu (d\zeta )dt\) is a standard Brownian motion and an independent compensated Poisson random measure, respectively, defined on a complete filtered probability space \((\Omega ,\mathcal {F},P)\) equipped with the filtration \(\mathbb {F}=\{\mathcal {F} _{t}\}_{t\ge 0}\) generated by the Brownian motion \(B(\cdot )\) and \(N(\cdot )\). The measure \(\nu \) is the Lévy measure of the Poisson random measure N, and the singular term \(L_{t}=L_{t}(y)\) is represented as the local time at a point \(y\in \mathbb {R}\) of a given \(\mathbb {F}\)-predictable process \(Y(\cdot )\) of the form

$$\begin{aligned} Y(t)=\int _{0}^{t}\phi (s)dB(s)+\int _{0}^{t}\int _{\mathbb {R}_{0}}\psi (s,\zeta )\tilde{N}(ds,d\zeta ), \end{aligned}$$
(1.3)

for some real deterministic functions \(\phi : [0,T] \rightarrow \mathbb {R},\psi :[0,T]\times \mathbb {R}_0\rightarrow \mathbb {R}\) satisfying

$$\begin{aligned} 0<\int _{t}^{T}\Big \{\phi ^{2}(t)+\int _{\mathbb {R}_{0}}\psi ^{2}(t,\zeta )\nu (d\zeta )\Big \}dt<\infty \text { a.s. for all }t\in [0,T]. \end{aligned}$$
(1.4)

The coefficients \(r(t),\mu (t),\alpha (t),\) \(\sigma (t)>0\) and \(\gamma (t,\zeta )>0\) are given bounded \(\mathbb {F}\)-predictable processes, with \(\sigma (t)\) bounded away from 0.

In this market we introduce a portfolio process \(u:[0,T]\times \Omega \rightarrow \mathbb {R}\) giving the fraction of the wealth invested in the risky asset at time t, and a consumption rate process \(c:[0,T]\times \Omega \rightarrow \mathbb {R}^{+}\) giving the fraction of the wealth consumed at time t. We assume that at any time t both u(t) and c(t) are required to be adapted to a given possibly smaller filtration \(\mathbb {G}=\{\mathcal {G}_{t}\}_{t\in [0,T]}\) with \(\mathcal {G} _{t}\subseteq \mathcal {F}_{t}\) for all t. For example, it could be a delayed information flow, with

$$\begin{aligned} \mathcal {G}_{t}=\mathcal {F}_{\max (0,t-\theta )},\quad t\ge 0,\text { for some delay }\theta >0. \end{aligned}$$
(1.5)

This case will be discussed in detail later.

Let us denote by \(\mathcal {A}_{\mathbb {G}}\) the set of all admissible consumption and portfolio processes. We say that c and u are admissible and write \(c,u\in \mathcal {A}_{\mathbb {G}}\) if, in addition, u is self-financing and \(\mathbb {E}\Big [\int _{0}^{T}(u(t)^{2}+c(t)^{2})dt\Big ]<\infty , \) where \(\mathbb {E}\) denotes expectation with respect to P. Note that if cu are admissible, then the corresponding wealth process \(X(t)=X^{c,u}(t)\) is described by the equation

$$\begin{aligned} dX(t)&=X(t^{-})[[(1-u(t))r(t)+u(t)\mu (t)-c(t)]dt+u(t)\alpha (t)dL_{t}\nonumber \\&\quad +u(t)\sigma (t)dB(t)+u(t)\textstyle \int _{\mathbb {R}_{0}}\gamma (t,\zeta )\widetilde{N}(dt,d\zeta )]. \end{aligned}$$
(1.6)

For simplicity, we put the initial value \(X(0)=1.\)

The optimal consumption and portfolio problem we study is the following:

Problem 1.1

Let \(a>0,b>0\) be given constants. Find admissible \(c^{*},u^{*},\) such that

$$\begin{aligned} J(c^{*},u^{*})=\sup _{c,u}J(c,u), \end{aligned}$$
(1.7)

where

$$\begin{aligned} J(c,u)=\mathbb {E}\Big [\int _{0}^{T}a\ln (c(t)X(t))dt+b\ln (X(T))\Big ]. \end{aligned}$$
(1.8)

Our results are the following:

Using methods from white noise calculus we find explicit expressions for the optimal consumption rate \(c^{*}(t)\) and the optimal portfolio \(u^{*} (t)\). Then we show that the value is finite for all positive delays in the information flow. In particular, this shows that there is no arbitrage in that case. This result appears to be new.

We also show that, under additional assumptions, the value goes to infinity when the delay goes to 0. This shows in particular that also when there are jumps the value is infinite when there is no delay, in agreement with the arbitrage results of Karatzas and Shreve [11] and Jarrow and Protter [10] in the Brownian motion case.

Remark 1.2

In our problem we are using the logarithmic utility function, both for the consumption and for the terminal value. It is natural to ask if similar results can be obtained for other utility functions. The method used in this paper is quite specific for the logarithmic utility and will not work for other cases. This issue will be discussed in a broader context in a future research.

2 Preliminaries

As we have mentioned above, we will use white noise calculus to find explicit expressions for the optimal consumption and the optimal portfolio. Specifically, we will define the local time in the terms of the Donsker delta function which is an element of the Hida space of stochastic distributions \((\mathcal {S})^{*}\). A brief introduction to white noise calculus is given in the Appendix. For more information on the underlying white noise theory we refer to Hida et al. [9], Oliveira [14], Holden et al. [8] and Di Nunno et al. [7] and Agram and Øksendal [3].

2.1 The Donsker delta function

We now define the Donsker delta function and give some of its properties. It will play a crucial role in our computations.

Definition 2.1

Let \(Y:\Omega \rightarrow \mathbb {R}\) be a random variable which also belongs to the Hida space \((\mathcal {S})^{*}\) of stochastic distributions. Then a continuous functional

$$\begin{aligned} \delta _{Y}(\cdot ):\mathbb {R}\rightarrow (\mathcal {S})^{*} \end{aligned}$$
(2.1)

is called a Donsker delta function of Y if it has the property that

$$\begin{aligned} \int _{\mathbb {R}}g(y)\delta _{Y}(y)dy=g(Y),\quad \text {a.s.} \end{aligned}$$
(2.2)

for all (measurable) \(g:\mathbb {R}\rightarrow \mathbb {R},\) such that the integral converges.

Explicit formulas for the Donsker delta function are known in many cases. For the Gaussian case, see Section 3.2. For details and more general cases, see e.g. Aase et al. [1].

In particular, for our process Y described by the diffusion (1.3), it is well known (see e.g. [6, 7, 13]) that the Donsker delta functional exists in \((\mathcal {S})^{*}\) and is given by

$$\begin{aligned} \delta _{Y(t)}(y)&=\frac{1}{2\pi }{\int _{\mathbb {R}}}\exp ^{\diamond }\left[ {\int _{0}^{t}\int _{\mathbb {R}_{0}}}(e^{ix\psi (s,\zeta )}-1)\tilde{N}(ds,d\zeta )\right. \nonumber \\&\quad +\int _{0}^{t}ix\phi (s)dB(s)\nonumber \\&\quad +\int _{0}^{t}\left\{ \int _{\mathbb {R}_{0}}(e^{ix\psi (s,\zeta )}-1-ix\psi (s,\zeta ))\nu (d\zeta )\right. \nonumber \\&\quad \left. \left. -\frac{1}{2}x^{2}\phi ^{2}(s)\right\} ds-ixy\right] dx, \end{aligned}$$
(2.3)

where \(\exp ^{\diamond }\) denotes the Wick exponential.

Moreover, if \(0\le s\le t,\) we can compute the conditional expectation

$$\begin{aligned}&\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {F}_{s}]\nonumber \\&\quad =\frac{1}{2\pi }{\int _{\mathbb {R}}}\exp \big [\int _{0}^{s} \int _{\mathbb {R}_{0}}ix\psi (r,\zeta )\tilde{N}(dr,d\zeta )+\int _{0}^{s} ix\phi (r)dB(r)\nonumber \\&\qquad +{\int _{s}^{t}\int _{\mathbb {R}_{0}}}(e^{ix\psi (r,\zeta )} -1-ix\psi (r,\zeta ))\nu (d\zeta )dr-\int _{s}^{t}\frac{1}{2}x^{2}\phi ^{2}(r)dr-ixy\bigg ]dx. \end{aligned}$$
(2.4)

Note that if we put \(s=0\) in (2.4), we get

$$\begin{aligned} \mathbb {E}[\delta _{Y(t)}(y)]&={\frac{1}{2\pi }\int _{\mathbb {R}} \exp \Big (-\frac{1}{2}x^{2}\int _{0}^{t}\phi ^{2}(r)dr}\\&\quad +{\int _{0}^{t}\int _{\mathbb {R}_{0}}(e^{ix\psi (r,\zeta )} -1-ix\psi (r,\zeta ))\nu (d\zeta )dr-ixy\Bigg )dx}<\infty . \end{aligned}$$

Putting \(\nu =0\) in (2.4), yields

$$\begin{aligned}&\frac{1}{2\pi }\int _{\mathbb {R}}\exp \Big [\int _{0}^{s} ix\phi (r)dB(r)-\textstyle {\int _{s}^{t}\frac{1}{2}}x^{2}\phi ^{2} (r)dr-ixy\Big ]dx\nonumber \\&\quad =\left( 2\pi \textstyle {\int _{s}^{t}}\phi ^{2}(r)dr\right) ^{-\frac{1}{2} }\exp \Bigg (-\frac{\left( \int _{0}^{s}\phi (r)dB(r)-y\right) ^{2}}{2 \int _{s} ^{t}\phi ^{2}(r)dr}\Bigg ), \end{aligned}$$
(2.5)

where we have used, in general, for \(a>0,b\in \mathbb {R}\), that

$$\begin{aligned} \int _{{\mathbb {R}}}e^{-ax^{2}-2bx}dx=\sqrt{\frac{\pi }{a}} e^{\frac{b^{2}}{a}}. \end{aligned}$$
(2.6)

In particular, applying the above to the random variable \(Y(t):=B(t)\) for some \(t\in (0,T]\) where B is Brownian motion starting at 0 , we get for all \(0\le s < t\),

$$\begin{aligned} \mathbb {E}[\delta _{B(t)}(y)|\mathcal {F}_s]=(2\pi (t-s))^{-\frac{1}{2}}\exp \Big [-\frac{(B(s)-y)^{2}}{2(t-s)}\Big ]. \end{aligned}$$
(2.7)

We will also need the following estimate:

Lemma 2.2

Assume that \(0\le s\le t\le T\). Then

$$\begin{aligned} \mathbb {E}[\delta _{Y(t)}(y)|\mathcal {F}_{s}]\le \Big (2\pi \textstyle {\int _{s}^{t} }\left\{ \phi ^{2}(r)+\int _{\mathbb {R}_{0}}\psi ^{2}(r,\zeta )\nu (d\zeta )\right\} dr \Big )^{-\frac{1}{2}}. \end{aligned}$$
(2.8)

Proof

From (2.4) we get, with \(i=\sqrt{-1}\),

$$\begin{aligned}&|\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {F}_{s}]| \le {\frac{1}{2\pi }\int _{\mathbb {R}}\exp }\left[ \int _{s} ^{t}\int _{\mathbb {R}_{0}}Re(e^{ix\psi (r,\zeta )}\right. \\&\qquad \left. -1-ix\psi (r,\zeta ))\nu (d\zeta )dr-\frac{1}{2}\int _{s}^{t}x^{2}\phi ^{2}(r)dr\right] {dx}\\&\quad \le {\frac{1}{2\pi }\int _{\mathbb {R}}\exp }\left[ {\int _{s} ^{t}\int _{\mathbb {R}_{0}}-\frac{1}{2}x^{2}\psi ^{2}(r,\zeta )\nu (d\zeta )dr-\frac{1}{2}\int _{s}^{t}x^{2}\phi ^{2}(r)dr}\right] {dx}\\&\quad =\textstyle {\frac{1}{2\pi }\int _{\mathbb {R}}\exp }\left[ {-\frac{1}{2} x^{2}\int _{s}^{t}}\left\{ {\phi ^{2}(r)+\int _{\mathbb {R}_{0}}\psi ^{2} (r,\zeta ))\nu (d\zeta )}\right\} {dr}\right] {dx}\\&\quad ={\Big (2\pi }\left( {\int _{s}^{t}}\left\{ {\phi ^{2} (r)+\int _{\mathbb {R}_{0}}\psi ^{2}(r,\zeta )\nu (d\zeta )}\right\} {dr}\right) {\Big )^{-\frac{1}{2}}}. \end{aligned}$$

\(\square \)

2.2 Local time in terms of the Donsker delta function

In this subsection we define the local time of \(Y(\cdot )\) at y and we give a representation of it in terms of the Donsker delta function.

Definition 2.3

The local time \(L_{t}(y)\) of \(Y(\cdot )\) at the point y and at time t is defined by

$$\begin{aligned} L_{t}(y)=\lim _{\epsilon \rightarrow 0}\frac{1}{2\epsilon }\lambda (\{s\in [0,t];Y(s)\in (y-\epsilon ,y+\epsilon )\}), \end{aligned}$$

where \(\lambda \) denotes Lebesgue measure on \(\mathbb {R}\) and the limit is in \(L^{2}(\lambda \times P)\).

Remark 2.4

Note that this definition differs from the definition in Protter [15] Corollary 3, page 230, in two ways:

  1. (i)

    We are using Lebesgue measure \(d\lambda (s)=ds\) as integrator, not \(d[Y,Y]_s\).

  2. (ii)

    Protter [15] is defining left-sided and right-side local times. Our local time corresponds to the average of the two.

If the process Y is Brownian motion both definitions coincide with the standard one. We choose our definition because it is convenient for our purpose.

There is a close connection between local time and the Donsker delta function of Y(t), given by the following result.

Theorem 2.5

The local time \(L_{t}(y)\) of Y at the point y and the time t is given by the following \(\mathcal {S}^{*}\)-valued integral

$$\begin{aligned} L_{t}(y)=\int _{0}^{t}\delta _{Y(s)}(y)ds, \end{aligned}$$
(2.9)

where the integral converges in \((\mathcal {S})^{*}\).

Proof

In the following we let \(\chi _F\) denote the indicator function of the Borel set F, i.e.

$$\begin{aligned} \chi _F(x)&= {\left\{ \begin{array}{ll} 1 \text { if } x \in F,\\ 0 \text { if } x \notin F. \end{array}\right. } \end{aligned}$$
(2.10)

By definition of the local time and the Donsker delta function, we have

$$\begin{aligned} L_{t}(y)&=\lim _{\epsilon \rightarrow 0} \int _{0}^{t} \textstyle {\frac{1}{2\epsilon }}\chi _{(y-\epsilon ,y+\epsilon )} (Y(s))ds \\&=\lim _{\epsilon \rightarrow 0} \int _{0}^{t} \Big (\int _{\mathbb {R}}\frac{1}{2\epsilon } \chi _{(y-\epsilon ,y+\epsilon )} (x) \delta _{Y(s)}(x) dx\Big )ds\\&=\lim _{\epsilon \rightarrow 0} \int _{\mathbb {R}}\textstyle {\frac{1}{2\epsilon }} \chi _{(y-\epsilon ,y+\epsilon )} (x) \Big (\int _{0}^{t} \delta _{Y(s)}(x) ds\Big )dx =\int _{0}^{t} \delta _{Y(s)}(y) ds, \end{aligned}$$

because the function \(y \mapsto \delta _{Y(s)}(y)\) is continuous in \((\mathcal {S})^{*}\). \(\square \)

3 Optimal consumption and portfolio in a market with a local time drift term under partial information

We now return to the model in the Introduction. Thus we consider the optimal portfolio and consumption problem (1.7)–(1.8) of an agent in the financial market (1.1) and (1.2). The agent has access to a partial information flow \(\mathbb {G}=\{\mathcal {G}_{t}\}_{t\ge 0}\) where \(\mathcal {G}_{t}\subseteq \mathcal {F}_{t}\) for all t. It is known that if \(\mathbb {G}=\mathbb {F}\), i.e. \(\mathcal {G}_{t}=\mathcal {F}_{t}\) for all t, and if there are no jumps (\(N=\nu =0\)), then the market is complete and it allows an arbitrage. See Karatzas and Shreve [11] and Jarrow and Protter [10]. It is clear that our market with jumps is not complete, even if \(\mathbb {G}=\mathbb {F}\). However, we will show that if \(\mathcal {G}_{t}=\mathcal {F}_{t-\theta }\) for some delay \(\theta >0\), then the market is viable (i.e. the optimal consumption and portfolio problem has a finite value) and it has no arbitrage. Moreover, we will find explicitly the optimal consumption and portfolio rates. If the delay goes to 0, we show that the value goes to infinity, in agreement with the existence of arbitrage in the no-delay case.

First we need the following auxiliary result.

Lemma 3.1

Suppose that \(\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_t] \in L^{2}(P)\) and that

$$\begin{aligned} \mu (t)-r(t)+\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_{t}]>0. \end{aligned}$$

Then there exists a unique solution \(u(t)=u^{*}(t)>0\) of the equation

$$\begin{aligned}&(a+b)\sigma ^{2}(t)u^{*}(t)+[a(T-t)+b]\int _{\mathbb {R}_{0} }\dfrac{u^{*}(t)\gamma ^{2}(t,\zeta )}{1+u^{*}(t)\gamma (t,\zeta )} \nu (d\zeta )\\&\quad =(a(T-t)+b)[\mu (t)-r(t)+\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_{t}]]. \end{aligned}$$

Proof

Define

$$\begin{aligned} F(u)=a_{1}u+a_{2}\int _{\mathbb {R}_{0}}\frac{u\gamma ^{2}(t,\zeta )}{1+u\gamma (t,\zeta )}\nu (d\zeta ),\quad u\ge 0, \end{aligned}$$

where \(a_{1}=(a+b)\sigma ^{2}(t),a_{2}=a(T-t)+b.\) Then

$$\begin{aligned} F^{\prime }(u)=a_{1}+a_{2}\int _{\mathbb {R}_{0}}\frac{\gamma ^{2}(t,\zeta )}{(1+\gamma (t,\zeta ))^{2}}\nu (d\zeta )>0, \end{aligned}$$

and

$$\begin{aligned} F(0)=0,\quad \lim _{u\rightarrow \infty }F(u)=\infty . \end{aligned}$$

Therefore, for all \(a>0\) there exists a unique \(u>0\) such that \(F(u)=a\). \(\square \)

We can now proceed to our first main result:

Theorem 3.2

(Optimal consumption and portfolio) Assume that \(\alpha \) and \(\gamma >0\) are \(\mathbb {G}\)-adapted and that

$$\begin{aligned}&\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_{t}]\in L^{2}(\lambda \times P)\text { and } \mathbb {E}[ \mu (t)-r(t)|\mathcal {G}_t] +\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_{t}]>0,\text { for all }t\in [0,T]. \end{aligned}$$

Then the optimal consumption rate is

$$\begin{aligned} c^{*}(t)=c^{*}(t)=\frac{a}{b+a(T-t)}, \end{aligned}$$

and the optimal portfolio is given as the unique solution \(u^{*}(t)>0\) of the equation

$$\begin{aligned}&(a+b)\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t] u^{*}(t)+(a(T-t)+b)\int _{\mathbb {R}_{0} }\frac{u^{*}(t)\gamma ^{2}(t,\zeta )}{1+u^{*}(t)\gamma (t,\zeta )}\nu (d\zeta )\\&\quad =(a(T-t)+b)\Big (\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]+\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G} _{t}]\Big ). \end{aligned}$$

In particular, if there are no jumps (\(N=\nu =0\)), the optimal portfolio will be

$$\begin{aligned} u^{*}(t)=\frac{(a(T-t)+b)\Big (\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]+\alpha (t)\mathbb {E}[\delta _{Y(t)} (y)|\mathcal {G}_{t}]\Big )}{(a+b)\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t]}. \end{aligned}$$

Proof

By the Itô formula for semimartingales, see e.g. Protter [15], we get that the solution of (4.2) is

$$\begin{aligned} X(t)&=\exp \Big (\int _{0}^{t}u(s)\sigma (s)dB(s)\\&\quad +\int _{0}^{t}\left\{ r(s)+[\mu (s)-r(s)]u(s)-c(s)-\frac{1}{2}\sigma ^{2}(s)u^{2}(s)\right\} ds\\&\quad +\int _{0}^{t}u(s)\alpha (s)dL_{s} +\int _{0}^{t}\int _{\mathbb {R}_{0}}\{\ln (1+u(s)\gamma (s,\zeta ))-u(s)\gamma (s,\zeta )\}\nu (d\zeta )ds\\&\quad +\int _{0}^{t}\int _{\mathbb {R}_{0}}\ln \left\{ 1+u(s)\gamma (s,\zeta )\right\} \widetilde{N}(ds,d\zeta )\Big ). \end{aligned}$$

Since \(\sigma \) and \(\gamma \) are bounded and \(u \in \mathcal {A}_{\mathbb {G}}\) the stochastic integrals in the exponent have expectation 0. Therefore we get

$$\begin{aligned} \mathbb {E}|\ln (X(t))]&=\mathbb {E}\Bigg [\int _{0}^{t}\{r(s)+[\mu (s)-r(s)]u(s)-c(s)-\frac{1}{2}\sigma ^{2}(s)u^{2}(s)\}ds\nonumber \\&\quad +\int _{0}^{t}u(s)\alpha (s)dL_{s}+\int _{0}^{t}\int _{\mathbb {R}_{0}} \{\ln (1+u(s)\gamma (s,\zeta ))-u(s)\gamma (s,\zeta )\}\nu (d\zeta )ds\Bigg ]. \end{aligned}$$
(3.1)

Formulas (4.2) and (1.8) and the Itô formula, lead to

$$\begin{aligned} J(c,u)&=\mathbb {E}\Bigg [\int _{0}^{T}a\ln (c(t)X(t))dt+b\ln (X(T))\Bigg ]\\&=\mathbb {E}\Big [\int _{0}^{T}\Bigg \{a\ln (c(t))+a\ln (X(t))\\&\quad +b\Bigg (r(t)+[\mu (t)-r(t)]u(t)-c(t)-\frac{1}{2}\sigma ^{2}(t)u^{2} (t)\Bigg )\Bigg \}dt\\&\quad +b\int _{0}^{T}u(t)\alpha (t)dL_{t}\\&\quad +b\int _{0}^{T}\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )dt\Bigg ]. \end{aligned}$$

Substituting (3.1) in the above, gives

$$\begin{aligned}&J(c,u)=\mathbb {E}\Big [\int _{0}^{T}\Big \{a\ln (c(t))\\&\quad +a\Big (\int _{0}^{t}\{r(s)+[\mu (s)-r(s)]u(s)-c(s)-\frac{1}{2}\sigma ^{2}(s)u^{2}(s)\}ds\\&\quad +\int _{0}^{t}u(s)\alpha (s)dL_{s}\\&\quad +\int _{0}^{t}\int _{\mathbb {R}_{0}}\{\ln (1+u(s)\gamma (s,\zeta ))-\pi (s)\gamma (s,\zeta )\}\nu (d\zeta )ds\Big )\\&\quad +b\Big (\int _{0}^{T}\Big \{r(t)+[\mu (t)-r(t)]u(t)-c(t)-\frac{1}{2}\sigma ^{2}(t)u^{2}(t)\Big )\Big \}dt\\&\quad +\int _{0}^{T}u(t)\alpha (t)dL_{t}\\&\quad +\int _{0}^{T}\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )dt\Big )\Big ]. \end{aligned}$$

Note that in general, we have, by the Fubini theorem,

$$\begin{aligned} \int _{0}^{T}\Big (\int _{0}^{t}h(s)ds\Big )dt&=\int _{0}^{T}\Big (\int _{s} ^{T}h(s)dt\Big )ds\\&=\int _{0}^{T}(T-s)h(s)ds=\int _{0}^{T}(T-t)h(t)dt, \end{aligned}$$

and

$$\begin{aligned} \int _{0}^{T}\Big (\int _{0}^{t}h(s)dL_{s}\Big )dt&=\int _{0}^{T}\Big (\int _{s}^{T}h(s)dt\Big )dL_{s}\\&=\int _{0}^{T}(T-s)h(s)dL_{s}=\int _{0}^{T}(T-t)h(t)dL_{t}. \end{aligned}$$

Therefore, using that

$$\begin{aligned} dL_{t}=dL_{t}(y)=\delta _{Y(t)}(y)dt, \end{aligned}$$

we get from the above that

$$\begin{aligned} J(c,u)&=\mathbb {E}\Big [\int _{0}^{T}E\Big [\Big \{a\Big (\ln (c(t))+(T-t)\{r(t)+[\mu (t)-r(t)]u(t) -c(t)-\frac{1}{2}\sigma ^{2}(t)u^{2}(t)\nonumber \\&\quad +(T-t)u(t)\alpha (t)\delta _{Y(t)}(y)\nonumber \\&\quad +(T-t)\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )\Big )dt\nonumber \\&\quad +b\Big (r(t)+[\mu (t)-r(t)]u(t)-c(t)-\frac{1}{2}\sigma ^{2}(t)u^{2} (t)\}+u(t)\alpha (t)\delta _{Y(t)}(y)\nonumber \\&\quad +\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )\Big )\Big \}\Big |\mathcal {G}_{t}\Big ]dt\Big ]. \end{aligned}$$
(3.2)

Using that \(c,u,\alpha \) and \(\gamma \) are \(\mathbb {G}\)-adapted, we obtain

$$\begin{aligned} J(c,u)&=\mathbb {E}\Bigg [\int _{0}^{T}\Big \{a\Big (\ln (c(t))+(T-t)\{\mathbb {E}[r(t)|\mathcal {G}_t]\nonumber \\&\quad +\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]u(t)\nonumber \\&\quad -c(t)-\frac{1}{2}\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t]u^{2}(t)\}\nonumber \\&\quad +(T-t)u(t)\alpha (t)\mathbb {E}[\delta {Y(t)}(y)|\mathcal {G}_{t}]\nonumber \\&\quad +(T-t)\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )\Bigg )\nonumber \\&\quad +b\Bigg (\mathbb {E}[r(t)|\mathcal {G}_t]+\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]u(t)-c(t)-\frac{1}{2}\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t]u^{2} (t)\nonumber \\&\quad +u(t)\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G}_{t}]\nonumber \\&\quad +\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (t,\zeta ))-u(t)\gamma (t,\zeta )\}\nu (d\zeta )\Bigg )\Bigg \}dt\Bigg ]. \end{aligned}$$
(3.3)

This we can maximise pointwise over all possible values \(c,u\in \mathcal {A} _{\mathbb {G}}\) by maximising for each t the integrand. Then we get the optimal consumption rate

$$\begin{aligned} c^{*}(t)=\frac{a}{b+a(T-t)}, \end{aligned}$$

and the optimal portfolio is given as the unique solution \(u^{*}(t)>0\) of the equation

$$\begin{aligned}&(a+b)\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t]u^{*}(t)\\&\qquad +[a(T-t)+b]\textstyle \int _{\mathbb {R}_{0} }\frac{u^{*}(t)\gamma ^{2}(t,\zeta )}{1+u^{*}(t)\gamma (t,\zeta )}\nu (d\zeta )\\&\quad =(a(T-t)+b)\Big [\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]\\&\qquad +\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {G} _{t}]\Big ]. \end{aligned}$$

In particular, if there are no jumps (\(N=\nu =0\)), we get

$$\begin{aligned} u^{*}(t)=\frac{(a(T-t)+b)\Big [\mathbb {E}[\mu (t)-r(t)|\mathcal {G}_t]+\alpha (t)\mathbb {E}[\delta _{Y(t)} (y)|\mathcal {G}_{t}]\Big ]}{(a+b)\mathbb {E}[\sigma ^{2}(t)|\mathcal {G}_t]}. \end{aligned}$$

\(\square \)

3.1 The case when \(\mathcal {G}_{t}=\mathcal {F}_{t-\theta }, \quad t \ge 0\)

From now on we restrict ourselves to the subfiltration \(\mathcal {G} _{t}=\mathcal {F}_{t-\theta },t\ge 0\) for some constant delay \(\theta >0\), where we put \(\mathcal {F}_{t-\theta }=\mathcal {F}_{0}\) for \(t\le \theta \). In this case we can compute the optimal portfolio and the optimal consumption explicitly. By (2.4) we have the following result:

Lemma 3.3

Assume that \(\alpha \) and \(\gamma > 0\) are \(\mathbb {G}_{\theta }\)-adapted, where \(\mathbb {G}_{\theta } =\{ \mathcal {F}_{t-\theta }\}_{t\ge 0}\). For \(t \ge \theta \) we have

$$\begin{aligned} \mathbb {E}\left[ \delta _{Y(t)}(y)|\mathcal {F}_{t-\theta }\right]&={\frac{1}{2\pi } \int _{\mathbb {R}}}\nonumber \\&\quad \exp \left[ \int _{0}^{t-\theta }\int _{\mathbb {R}_{0}} ix\psi (r,\zeta )\widetilde{N}(dr,d\zeta )+\int _{0}^{t-\theta }ix\phi (r)dB(r)\right. \nonumber \\&\quad +{\int _{t-\theta }^{t}}\int _{\mathbb {R}_{0}}(e^{ix\psi (r,\zeta )}-1-ix\psi (r,\zeta ))\nu (d\zeta )dr\nonumber \\&\left. \quad -\int _{t-\theta }^{t}\frac{1}{2}x^{2}\phi ^{2}(r)dr-ixy\right] dx. \end{aligned}$$
(3.4)

In particular, if \(\psi =0\) and \(\phi =1\), we get \(Y=B\) and (see also (2.7))

$$\begin{aligned} \mathbb {E}[\delta _{B(t)}(y)|\mathcal {F}_{t-\theta }]=(2\pi \theta )^{-\frac{1}{2}} \exp \Big [-\frac{(B(t-\theta )-y)^{2}}{2\theta }\Big ]. \end{aligned}$$
(3.5)

Then by Theorem 3.2, we get

Theorem 3.4

Suppose \(\mathcal {G}_{t}=\mathcal {F}_{t-\theta }\) with \(\theta >0.\) Then the optimal consumption rate is given by

$$\begin{aligned} c^{*}(t)=\frac{a}{b+a(T-t)}, \end{aligned}$$

and the optimal portfolio is given as the unique solution \(u^{*}(t)>0\) of the equation

$$\begin{aligned}&(a+b)\mathbb {E}[\sigma ^{2}(t)|\mathcal {F}_{t-\theta }]u^{*}(t)+(a(T-t)+b)\int _{\mathbb {R}_{0} }\frac{u^{*}(t)\gamma ^{2}(t,\zeta )}{1+u^{*}(t)\gamma (t,\zeta )}\nu (d\zeta )\\&\quad =(a(T-t)+b)\big (\mathbb {E}[\mu (t)-r(t)|\mathcal {F}_{t-\theta }]+\alpha (t)\mathbb {E}[\delta _{Y(t)}(y)|\mathcal {F} _{t-\theta }]\big ). \end{aligned}$$

In particular,

$$\begin{aligned} \sup _{c,u}J(c,u)=J(c^{*},u^{*})<\infty , \end{aligned}$$

and there is no arbitrage in the market.

4 The limiting case when the delay goes to 0

In this section, we concentrate on the delay case and with optimal portfolio only, i.e. without consumption. Thus we are only considering utility from terminal wealth, and we put \(a=0\) and \(b=1\) in Theorem 3.2. Moreover, we assume that \(\phi =1\) and \(\psi =0\), i.e. that

$$\begin{aligned} Y(t)=B(t); \quad t \in [0,T]. \end{aligned}$$
(4.1)

Also, to simplify the calculations we assume that \(r=0\) and \(\mu (t)=\mu>0,\alpha (t)=\alpha>0,\sigma (t)=\sigma >0\) are constants, and \(\gamma (t,\zeta )=\gamma (\zeta )\) is deterministic and does not depend on t. Then the wealth equation will take the form

$$\begin{aligned} dX(t)&=X(t)[u(t)\mu dt+u(t)\alpha dL_{t}\nonumber \\&\quad +u(t)\sigma dB(t)+u(t)\textstyle \int _{\mathbb {R}_{0}}\gamma (\zeta )\widetilde{N}(dt,d\zeta )];\quad t \in [0,T],\quad X(0)=1, \end{aligned}$$
(4.2)

where the singular term \(L_{t}=L_{t}(y)\) is represented as the local time at a point \(y\in \mathbb {R}\) of \(B(\cdot )\). The performance functional becomes

$$\begin{aligned} J_{0}(u)=\mathbb {E}[\ln X^{(u)}(T)];\quad u\in \mathcal {A}_{\theta }, \end{aligned}$$

where \(\mathcal {A}_{\theta }\) denotes the set of all \(\mathcal {F}_{t-\theta }\) -predictable control processes. This now gets the form

$$\begin{aligned} J_{0}(u)&=\mathbb {E}\Big [\int _{0}^{T}\Big \{\mu u(t)-\frac{1}{2}\sigma ^{2} u^{2}(t)+u(t)\alpha \mathbb {E}[\delta _{Y(t)}(y)|\mathcal {F}_{t-\theta }]\nonumber \\&\quad +\int _{\mathbb {R}_{0}}\{\ln (1+u(t)\gamma (\zeta ))-u(t)\gamma (\zeta )\}\nu (d\zeta )\Big \}dt\Big ]. \end{aligned}$$
(4.3)

Our second main result is the following:

Theorem 4.1

Suppose in addition to the above that

$$\begin{aligned} \int _{\mathbb {R}_{0}}\gamma ^{2}(\zeta )\nu (d\zeta )<\sigma ^{2}. \end{aligned}$$

Then

$$\begin{aligned} \lim _{\theta \rightarrow 0^{+}}\sup _{u\in \mathcal {A}_{\theta }}J_{0}(u)=\infty . \end{aligned}$$

In particular, if there is no delay (\(\theta =0\)) the value of the optimal portfolio problem is infinite.

Proof

For given \(\theta >0\) choose

$$\begin{aligned} u_{\theta }(t)=\frac{\mu +\alpha R}{\sigma ^{2}}, \end{aligned}$$

where we for simplicity put

$$\begin{aligned} R=R_{\theta }=\mathbb {E}[\delta _{B(t)}(y)|\mathcal {F}_{t-\theta }]. \end{aligned}$$

Then we see that

$$\begin{aligned} J_{0}(u_{\theta })\ge \frac{1}{2}\mathbb {E}[(\mu +\alpha R)^{2}]\Big (1-{\frac{\int _{\mathbb {R}_{0}}\gamma ^{2}(\zeta )\nu (d\zeta )}{\sigma ^{2}}}\Big )=:C_{1} \mathbb {E}[(\mu +\alpha R)^{2}]\ge C_{2}+C_{3}\mathbb {E}[R^{2}], \end{aligned}$$

since, by (2.8), \(\mathbb {E}[R]<\infty \). Here \(C_{1},C_{2},C_{3}\) are positive constants.

It remains to prove that

$$\begin{aligned} \mathbb {E}[R_{\theta }^{2}]\rightarrow \infty \text { when }\theta \rightarrow 0^{+}. \end{aligned}$$

To this end, note that by (2.5) we have

$$\begin{aligned} \mathbb {E}[R_{\theta }^{2}]=\mathbb {E}[(\delta _{B(t)}(y)|\mathcal {F}_{t-\theta })^{2}]=(2\pi \theta )^{-1}\mathbb {E}\Big [\exp \Big (-\frac{2(B(t-\theta )-y)^{2}}{2\theta }\Big )\Big ]. \end{aligned}$$
(4.4)

By formula 1.9.3(1) p.168 in [4] we have, with \(\kappa >0\) constant,

$$\begin{aligned} \mathbb {E}[\exp (-\kappa (B(t-\theta )-y)^{2})]=\frac{1}{1+2\kappa (t-\theta )} \exp \Big (-\frac{\kappa y^{2}(t-\theta )}{1+2\kappa (t-\theta )}\Big ). \end{aligned}$$
(4.5)

Applying this to \(\kappa =\frac{1}{\theta }\) we get

$$\begin{aligned} \mathbb {E}[R_{\theta } ^2]= \frac{1}{2\pi \sqrt{\theta } \sqrt{2t-\theta } } \exp \Big ( -\frac{y^2}{2t-\theta } \Big )\rightarrow \infty , \end{aligned}$$
(4.6)

when \(\theta \rightarrow 0\). \(\square \)

5 The Brownian motion case

In the case when \(Y(t)=B(t)\) the computations above can be made more explicit. We now illustrate this, assuming for simplicity that \(y=0\). Then by Theorem 3.4 the optimal portfolio \(\widehat{u}(t)\) is given by

$$\begin{aligned} \widehat{u}(t)=\frac{\mu +\alpha \Lambda (t)}{\sigma ^{2}}, \end{aligned}$$

where

$$\begin{aligned} \Lambda (t)&=\mathbb {E}[\delta _{B(t)}(0)|\mathcal {F}_{t-\theta }]=(2\pi \theta )^{-\frac{1}{2}}\exp \Big [-\frac{B(t-\theta )^{2}}{2\theta }\Big ];\quad t\ge \theta ,\nonumber \\ \Lambda (t)&=\frac{1}{\sqrt{2\pi \theta }};\quad 0\le t\le \theta . \end{aligned}$$
(5.1)

By (4.3) and (5.1) we see, after some algebraic operations, that the corresponding performance \(\widehat{J}_{\theta }=J(0,\widehat{\pi })\) is

$$\begin{aligned} \widehat{J}_{\theta }&=\mathbb {E}\Big [\int _{0}^{T}\Big (\mu \widehat{\pi }(t)-\frac{1}{2}\sigma ^{2}\widehat{\pi }^{2}(t)+\widehat{\pi }(t)\alpha \Lambda (t)\Big )dt\Big ]=\mathbb {E}\Big [\int _{0}^{T}\frac{(\mu +\alpha \Lambda (t))^{2}}{2\sigma ^{2}}dt\Big ]\\&=A_{1}+A_{2}+A_{3}, \end{aligned}$$

where

$$\begin{aligned} A_{1}=\frac{\mu ^{2}}{2\sigma ^{2}}T,\quad A_{2}=\frac{\mu \alpha }{\sigma ^{2} }\mathbb {E}\Big [\int _{0}^{T}\Lambda (t)dt\Big ],\quad A_{3}=\frac{\alpha ^{2}}{2\sigma ^{2}}\mathbb {E}\Big [\int _{0}^{T}\Lambda ^{2}(t)dt\Big ]. \end{aligned}$$

Using the density of B(s), we get

$$\begin{aligned} \mathbb {E}\Big [\exp \Big (-\frac{B^{2}(s)}{2\theta }\Big )\Big ]&=\int _{\mathbb {R}} \exp \left( -\frac{y^{2}}{2\theta }\right) \frac{1}{\sqrt{2\pi s}}\exp \left( -\frac{y^{2}}{2s}\right) dy\nonumber \\&=\frac{1}{\sqrt{2\pi s}}\int _{\mathbb {R}}\exp \left( -\frac{1}{2} y^{2}(\frac{1}{\theta }+\frac{1}{s})\right) dy. \end{aligned}$$
(5.2)

In general we have, for \(a>0\),

$$\begin{aligned} \int _{\mathbb {R}}\exp (-ay^{2})dy=\sqrt{\frac{\pi }{a}}, \end{aligned}$$

we conclude, by putting \(s=t-\theta \) in (5.2), that

$$\begin{aligned} A_{2}=\frac{\theta }{\sqrt{2\pi \theta }}+\int _{\theta }^{T}\frac{\mu \alpha }{\sigma ^{2}\sqrt{2\pi \theta }}\sqrt{\frac{\theta }{t}}dt=\sqrt{\frac{\theta }{2\pi }}+\frac{2\mu \alpha (\sqrt{T}-\sqrt{\theta })}{\sigma ^{2}\sqrt{2\pi }}. \end{aligned}$$

Finally we use similar calculations to compute

$$\begin{aligned} A_{3}=\frac{\alpha ^{2}}{2\sigma ^{2}}(2\pi \theta )^{-1}\left( \theta +\int _{\theta }^{T}\Psi (t)dt\right) , \end{aligned}$$

where, putting \(t-\theta =s\),

$$\begin{aligned} \psi (t)&=\mathbb {E}\left[ \exp \left( -\frac{B(s)^{2}}{\theta }\right) \right] =\int _{\mathbb {R}}e^{-\frac{y^{2}}{\theta }}\frac{1}{\sqrt{2\pi s}} e^{-\frac{y^{2}}{2s}}dy\\&=\frac{1}{\sqrt{2\pi s}}\int _{\mathbb {R}}\exp \left( -y^{2}(\frac{1}{\theta }+\frac{1}{2s})\right) dy=\frac{1}{\sqrt{2\pi s}}\sqrt{\frac{\pi }{\frac{1}{\theta }+\frac{1}{2s}}}=\frac{1}{\sqrt{\frac{2s}{\theta }+1}}. \end{aligned}$$

This gives

$$\begin{aligned} A_{3}=\frac{\alpha ^{2}}{2\sigma ^{2}}\left( \frac{1}{2\pi }+\int _{0}^{T-\theta }\frac{1}{2\pi \sqrt{\theta }\sqrt{2s+\theta }}ds\right) =\frac{\alpha ^{2}}{4\pi \sigma ^{2}}\left( 1+\frac{\sqrt{2T-\theta }-\sqrt{\theta }}{\sqrt{\theta } }\right) . \end{aligned}$$

We have proved the following:

Theorem 5.1

The optimal performance with a given delay \(\theta > 0\) is given by

\(\widehat{J}_{\theta }=\frac{\mu ^{2}}{2\sigma ^{2}}T+\sqrt{\frac{\theta }{2\pi } }+\frac{2\mu \alpha (\sqrt{T}-\sqrt{\theta })}{\sigma ^{2}\sqrt{2\pi }} +\frac{\alpha ^{2}}{4\pi \sigma ^{2}}(1+\frac{\sqrt{2T-\theta }-\sqrt{\theta } }{\sqrt{\theta }}).\)

In particular, \(\widehat{J}_{\theta }\rightarrow \infty \) when \(\theta \rightarrow 0.\)

Corollary 5.2

  1. (i)

    For all information delays \(\theta >0\) the value of the optimal portfolio problem is finite.

  2. (ii)

    When there is no information delay, i.e. when \(\theta =0\), the value is infinite.