Skip to main content
Erschienen in: Mathematics and Financial Economics 2/2023

Open Access 13.01.2023

A pricing formula for delayed claims: appreciating the past to value the future

verfasst von: Enrico Biffis, Beniamin Goldys, Cecilia Prosdocimi, Margherita Zanella

Erschienen in: Mathematics and Financial Economics | Ausgabe 2/2023

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

We consider the valuation of contingent claims with delayed dynamics in a Samuelson complete market model. We find a pricing formula that can be decomposed into terms reflecting the current market values of the past and the future, showing how the valuation of prospective cashflows cannot abstract away from the contribution of the past. As a practical application, we provide an explicit expression for the market value of human capital in a setting with wage rigidity. The formula we derive has successfully been used to explicitly solve the infinite dimensional stochastic control problems addressed in Biffis et al. (SIAM J Control Optim 58(4):1906–1938, 2020), Djeiche et al. (Stoch Process Appl 145:48–85, 2022) and Biagini et al. (SIAM J Financial Math 13(3):1004–1039, 2022).
Hinweise
The original online version of this article was revised: In the original publication of the article, Missing Open Access funding information has been added in the Funding Note.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

It is a standard result in asset pricing theory that the absence of arbitrage opportunities is “essentially” equivalent to the existence of an equivalent probability measure under which the price of any contingent claim is a local martingale after deflation by the money market account; see [17, 24, 25]. In this paper, we preserve the standard formulation of arbitrage pricing in a complete market model with security prices evolving as geometric Brownian motions (GBM), in the spirit of Samuelson’s contribution [36] (see [37] for an overview). The main novelty of our work is that we consider contingent claims that have dynamics described by a stochastic delay differential equation (SDDE).
It is perhaps surprising that using the no-arbitrage pricing machinery we are able to derive an explicit valuation formula for dynamics with memory, a case that is notoriously difficult to handle. However, we find a particularly appealing solution showing that the price can be decomposed into a term related to the ‘current market value of the past’ (in a sense to be made precise below), and the more traditional term reflecting the ‘current market value of the future’ (in the spirit of a discounted cashflow approach). In our setting the contribution of the past is represented by the portion of a contingent claim’s past trajectory that shapes its dynamics going forward.1 Our pricing formula demonstrates that the market consistent valuation of future cashflows cannot ignore the contribution of the past. This is important for a number of applications in which path-dependency is a key feature of the state variables, as we now discuss.
As a practical application of our results, we consider in detail contingent claims representing stochastic wages received by an agent over her lifetime (e.g., [7, 20]). It is well known that when labor income is spanned by the assets available for trade, the market value of human capital can be easily derived via risk-neutral valuation in a setting with labor income driven by a Stochastic Differential Equation (SDE); see [13] for an overview. However, the empirical literature on labor income dynamics widely relies on auto-regressive moving average (ARMA) processes (e.g., [1, 26, 31, 33]).
The contributions by [19, 29, 35] show how SDDEs can be understood as the weak limit of discrete time ARMA processes. Therefore specific classes of SDDEs can be used to model labor income, thereby providing the continuous time counterpart of ARMA models supported by the empirical literature. We therefore consider the introduction of delayed drift and volatility coefficients in a GBM labor income model to provide a tractable example of wage dynamics that adjusts slowly to financial market shocks. We obtain a closed form solution for human capital, which makes explicit the contributions of the market value of the past and the future. Our results demonstrate that SDDEs are valuable modeling tools that can address the findings of a large body of empirical literature on wage rigidity (e.g., [3, 15, 28, 30]). Moreover, the results open the way to finding explicit solutions to interesting classes of lifecycle portfolio choice problems with state costraints (see [6, 7, 16]), as discussed in Sect. 3.
Although in this paper we discuss the application to human capital more extensively, the results can be used for other applications of interest. For instance, we provide some references to the literature on counterparty risk and derivatives valuation, in which analogous dynamics arise in the context of collateralization procedures entailing a delay in the marking-to-market procedure of over-the-counter derivatives (e.g., [11, 12]).
It should be noted that no-arbitrage pricing in the case of delayed price dynamics has been recently studied by several authors (see [2, 32], for example). Their focus, however, is on proving completeness of a market with security prices with delayed dynamics and hence very different from ours. On the other hand, their work suggests the opportunity to explore extensions of our results to broader settings in which market completeness is preserved, including those in which tradable assets have delayed drift and volatility terms.
The paper is organized as follows. In the next section, we introduce the setup and state our main result. In Section 3 we discuss in detail an application to human capital valuation and briefly outline other areas of application. In Sect. 4, we prove the main result of the paper. Finally, Sect. 5 offers concluding remarks. We relegate to an “Appendix” the proofs of some more technical results.

2 Setup and statement of the main result

We consider a filtered probability space \((\Omega , {\mathcal {F}}, {\mathbb {F}}, {\mathbb {P}})\) and an \({\mathbb {F}}\)-adapted vector valued process \((S_0,S)\), representing the price evolution of a money market account, \(S_0\), and n risky assets, \(S=(S_1,\ldots ,S_n)^\top \), whose dynamics is given by
$$\begin{aligned} {\left\{ \begin{array}{ll} \text {d}S_0(t)= S_0(t) r \text {d}t,\\ \text {d}S(t) =\text {diag}(S(t)) \left( \mu \text {d}t + \sigma \text {d}Z(t)\right) ,\\ S_0(0)=1,\\ S(0)\in {{\mathbb {R}}}^n_{+}, \end{array}\right. } \end{aligned}$$
(1)
where \(\mu \in {\mathbb {R}}^n\), and the matrix \(\sigma \in {\mathbb {R}}^{n \times n}\) is assumed to be invertible. Here Z is an n-dimensional Brownian motion and we assume that \({\mathbb {F}}:=\{{\mathcal {F}}_t\}_{t\ge 0}\) is the filtration generated by Z, and enlarged with the \({\mathbb {P}}\)-null sets. In the following, for given vectors \(a, b \in \mathbb {R}^n\), we denote by \(a \cdot b\) the scalar product in \(\mathbb {R}^n\) and by \(\Vert \cdot \Vert \) the corresponding norm.
We are interested in the valuation of a payment stream represented by an \({\mathbb {F}}\)-adapted process \((y_t)_{t\ge 0}\). The payment stream can be thought of as capturing the mark-to-market process of a trading account, the flow of profits and losses generated from a trading strategy, the collateral flows arising from an over-the-counter derivative transaction, or the labor income received by an agent over time. The latter is the application we will consider in detail in the next section.
We assume that the payment stream y obeys the following SDDE with delay in both the drift and the diffusion terms:
$$\begin{aligned} {\left\{ \begin{array}{ll} \text {d}y(t) = &{}\left[ y(t) \mu _y+\int _{-d}^0 y(t+s) \phi (\text {d}s) \right] \text {d}t \\ &{}+\left[ y(t)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(t+s) \varphi _1(\text {d}s) \\ \vdots \\ \int _{-d}^0y(t+s) \varphi _n(\text {d}s) \end{pmatrix} \right] \cdot \text {d}Z(t),\\ y(0)=&{} x_0,\\ y(s)=&{} x_1(s ) \quad \hbox { for }s \in [-d,0), \end{array}\right. } \end{aligned}$$
(2)
where \(\mu _y \in {\mathbb {R}}\), \(\sigma _y \in {\mathbb {R}}^n\), the signed measures \(\phi ,\varphi _i\) (for \(i=1,\dots ,n\)) have bounded variation on \(\left[ -d,0\right] \) and \(x_0 \in {\mathbb {R}}\), \(x_1 \in L^2 \big ( [-d,0]; {\mathbb {R}}\big ) \), where \(L^2 \big ( [-d,0]; {\mathbb {R}}\big )\) denotes the space of Lebesgue square integrable (deterministic) functions on \([-d,0]\). In the following, we denote by \(|\phi |\) the total variation of a signed measure of bounded variation \(\phi \) on \([-d,0]\). Existence of a unique solution to SDDE (2) is ensured by the following result, which is proved in “Appendix A.1”.
Proposition 2.1
For any given initial datum \((x_0,x_1)\in \mathbb {R} \times L^2(-d,0;\mathbb {R})\) equation (2) admits a unique strong (in the probabilistic sense) solution \(y\in L^2(\Omega \times [0,T] )\), for all \(T>0\), with \(\mathbb {P}\)-a.s. continuous paths.
Assuming the process y to represent an agent’s income stream, we see that (2) provides a compelling example (which will be shown to be tractable) of wage dynamics adjusting slowly to financial market shocks. In particular, the measures \(\phi ,\varphi _1,\ldots \varphi _n\) can be thought as modulating the impact of past income’s realizations on wages going forward. As discussed in [7], this setting can also be related to the literature on ‘learning your income’ (e.g., [22]), whereby an agent learns about her earning potential based on past wages. Here, moving averages can be used as a tractable substitute for fully fledged Bayesian filters and can be justified on the grounds of bounded rationality.
As the market is complete and the stream process y is spanned by the stock S, we can find uniquely the current value of the future uncertain income stream by considering the expected discounted value under the unique stochastic discount factor \(\xi \) (see [18]):
$$\begin{aligned} H\left( t_0\right) :=\xi (t_0)^{-1} {\mathbb {E}}\left[ \int _{t_0}^{+ \infty } \xi (s) y(s) \text {d}s \Big \vert {\mathcal {F}}_{t_0}\right] , \end{aligned}$$
(3)
where the process \(\xi \) is \({\mathbb {F}}\)-adapted and evolves according to the following SDE,
$$\begin{aligned} \left\{ \begin{array}{ll} \text {d} \xi (t)&{} = - \xi (t) \left( r \text {d}t + \kappa \cdot \text {d}Z(t)\right) \\ \xi (0)&{}=1, \end{array}\right. \end{aligned}$$
(4)
with the constant \(\kappa \) representing the market price of risk, which is given by
$$\begin{aligned} \kappa = (\sigma ^\top )^{-1} (\mu - r {\textbf{1}}), \end{aligned}$$
(5)
with \({\textbf{1}} = (1,\dots , 1)^\top \) denoting the unitary vector in \({\mathbb {R}}^n\).
Our main goal now is to provide an explicit formula for the quantity \(H(t_0)\) given in (3). The challenging aspect of the problem lies in the fact that we consider a payment stream whose dynamics is path dependent, thus making the problem considerably different from and harder to prove than the cases discussed in the extant literature. Before stating the main result of the paper, we introduce the following conditions that will be assumed to hold throughout the paper.
Assumption 2.2
The following conditions are assumed to apply:
(i)
The measure \(\Phi \) on \([-d,0]\) defined as
$$\begin{aligned} \Phi (\cdot ):=\phi (\cdot )-(\varphi _1(\cdot ),\ldots ,\varphi _n(\cdot )) \cdot \kappa \end{aligned}$$
(6)
is a signed measure of bounded variation.
 
(ii)
The quantity
$$\begin{aligned} r-\mu _0+\sigma _y \cdot \kappa - \int _{-d}^0 e^{r\tau }|\Phi |(\textrm{d} \tau ) \end{aligned}$$
(7)
is assumed to be strictly positive.
 
Under assumption 2.2, which will be discussed in details in Sect. 4.5, the following result holds.
Theorem 2.3
Under Assumption 2.2, for any \(t_0\ge 0\), the quantity \(H(t_0)\) defined in (3) admits the following explicit expression:
$$\begin{aligned} H(t_0)= \frac{1}{K} \left( y(t_0)+ \int _{-d}^0 G(s) y(t_0+s)\, \textrm{d}s\right) ,\quad {\mathbb {P}}-a.s., \end{aligned}$$
(8)
where \(y(t_0)\) denotes the solution at time \(t_0\) of equation (2) and the constant K and function G are given by:
$$\begin{aligned} K:=r-\mu _0+\sigma _y \cdot \kappa - \int _{-d}^0 e^{r\tau }\Phi (\textrm{d} \tau ),\\ G(s) : = \int _{-d}^s e^{-r(s-\tau )} \Phi (\textrm{d}\tau ). \end{aligned}$$
In expression (8), we can recognize an annuity factor, \(K^{-1}\), multiplying a term representing the current market value of labor income, \(y(t_0)\), and a term representing the current market value of the past trajectory of y over the time window \((t_0-d,t_0)\). The ‘market value of the past’ trades off the returns on the payment stream against its exposure to financial risk, as can be seen from expression (6). When the delay terms in the drift and volatility coefficients vanish, the valuation of the payment stream reduces to \(K^{-1}y(t_0)\), in line with [20], among others.
Remark 2.4
The setup described above can be extended to the case of payments over a bounded horizon in some interesting situations. When the payment stream is received until an exogenous Poisson stopping time \(\tau _\delta \) (representing death or an irreversible unemployment shock when y represents labor income), expression (8) still applies, provided discounting is carried out at rate \(r+\delta \) instead of r, where \(\delta >0\) represents the Poisson parameter; see [6, 7, 16]. The case in which payments are received until a finite, deterministic time (capturing irreversible retirement when y represents labor income) can also be considered, at the price of additional technical work; see [5].

3 Applications

3.1 Optimal portfolio problems with path dependent labor income

We now consider in detail the case in which the contingent claim y in (2) represents stochastic wages received by an agent over her lifetime. As discussed in the Introduction, SDDEs allow us to rely on continuous time labor income dynamics better matching some of the salient features documented in the empirical literature. The path-dependency of (2) captures the slow adjustment of labor income to financial market shocks and provides a counterpart to discrete time ARMA models used to model wage dynamics. See [7] for a comprehensive list of references and [5] for the complexities brought about by considering unspanned shocks in labor income dynamics.
In the context of lifecycle portfolio choice, obtaining an explicit expression for human capital is often crucial not only to reveal the structure of optimal solutions (e.g., [13]), but also to handle state constraints (e.g., [7, 20]). We will discuss these points by making reference to some applications in Sects. 3.1.13.1.3. Before proceeding, we note that assumption 2.2 is all we need to provide the explicit valuation result of Theorem 2.3, but the particular application to human capital requires labor income to be positive almost surely. A sufficient condition for this to be the case is as follows (see [7, Proposition 2.7] for a proof).
Remark 3.1
When \(\varphi _i=0\) for all \(i=1,ldots,n\) (i.e., when the delay term in the volatility coefficient of (2) vanishes), the variation of constants formula yields the following explicit representation:
$$\begin{aligned} y(t)= {\mathcal {E}}(t)\big ( x_0 + {\mathcal {I}}(t)\big ), \end{aligned}$$
(9)
where
$$\begin{aligned} {\mathcal {E}}(t):= e^{(\mu _y - \frac{1}{2}|\sigma _y|^2)t+ \sigma _y Z(t) }, \qquad {\mathcal {I}}(t) := \int _0^t {\mathcal {E}}^{-1}(u) \int _{-d}^0 \,y(s+u) \phi (\textrm{d}s)\, \textrm{d}u. \end{aligned}$$
One can then see that in this case \(y(t)>0\) \(\mathbb {P}\)-a.s. if \(x_0>0\), \(x_1 \ge 0\) a.s. and \(\phi \ge 0\) a.s..
The results of Theorem 2.3 and the solution approach followed in this paper show how tools from infinite-dimensional analysis can be successfully used to address non-Markovian valuation problems, which are beyond the reach of conventional approaches. In the following, we illustrate some of those applications to show how the findings of Theorem 2.3 (or suitable generalization of it) can be successfully used to identify explicit solutions in some interesting situations.

3.1.1 Infinite horizon lifecycle portfolio choice

In [7] the authors solve an infinite horizon portfolio problem with borrowing constraints in which an agent receives labor income affected by financial market shocks in a path dependent way. The framework is the complete market model described by (1). An agent is endowed with initial wealth \(w\ge 0\) and receives wages until death, which coincides with the first jump time \(\tau _{\delta }\) of a Poisson process with parameter \(\delta >0\). The financial wealth of the agent at time \(t\ge 0\), denoted by W(t), can be invested in the riskless and risky assets. The agent receives wages at rate y(t) and consumes at rate \( c(t)\ge 0\). The wealth amount allocated to the risky assets is \( \theta (t)\in {\mathbb {R}}^n\) at each time \(t\ge 0\). The agent sets a bequest target \(B(\tau _\delta )\ge 0\) at death. In line with [20], the agent can purchase life insurance at the (actuarially fair) premium rate \(\delta ( B(t)-W(t))\), for \(t<\tau _\delta \), to reach the bequest target and hence cover any shortfall \(B(\tau _{\delta })-W(\tau _{\delta })\) at death. The agent’s wealth (before death) is assumed to obey the standard dynamic budget constraint given by the SDE
$$\begin{aligned} {\left\{ \begin{array}{ll} dW(t) = \left[ W(t) r + \theta (t)\cdot (\mu -r{\textbf{1}}) + y(t) - c(t)\right. \\ \qquad \quad \left. -\delta \left( B(t)-W(t)\right) \right] \textrm{d}t + \theta (t)\cdot \sigma \textrm{d}Z(t),\\ W(0)=w . \end{array}\right. } \end{aligned}$$
(10)
In line with the empirical evidence on labour income dynamics recalled in the Introduction, labor income is modelled via the following SDDE:
$$\begin{aligned} {\left\{ \begin{array}{ll} \textrm{d}y(t) =\left[ \mu _y y(t)+\int _{-d}^0 \phi (s) y(t+s) \textrm{d}s \right] \textrm{d}t + y(t)\sigma _y \textrm{d}Z(t),\\ y(0)= x_0, \quad y(s) = x_1(s) \text{ for } s \in [-d,0), \end{array}\right. } \end{aligned}$$
(11)
where \(\mu _y \in {\mathbb {R}}\), \(\sigma _y \in {\mathbb {R}}^n\) and the functions \(\phi (\cdot ), x_1(\cdot )\) belong to \(L^2\left( -d,0; {\mathbb {R}}\right) \) (thus making (11) a particular case of (2)).
Denoting by \(k>0\) the intensity of preference for leaving a bequest, by \(\gamma \in (0,1) \cup (1, +\infty )\) the risk-aversion coefficient and by \(\rho >0\) the subjective discount rate, the objective is to maximize the expected utility from lifetime consumption and bequest, which on \(\{\tau ^\delta >t\}\) takes the form:
$$\begin{aligned} {\mathbb {E}} \left( \int _{0}^{+\infty } e^{-(\rho + \delta ) t } \left( \frac{c(t)^{1-\gamma }}{1-\gamma } + \delta \frac{\big (k B(t)\big )^{1-\gamma }}{1-\gamma }\right) \textrm{d}t \right) . \end{aligned}$$
(12)
Maximization is carried out over all triplets \(\left( c,\theta ,B\right) \in \Big \{{\mathbb {F}}-\text{ predictable } \ c(\cdot ), B(\cdot ), \theta (\cdot ) :c(\cdot ), B(\cdot ) \in L^1 (\Omega \times [0, +\infty );{\mathbb {R}}_{+}),\theta (\cdot ) \in L^2(\Omega \times {\mathbb {R}}; {\mathbb {R}}^n)\Big \} \) subject to the state constraint
$$\begin{aligned} W(t) + \xi ^{-1}(t){\mathbb {E}}\left( \int _t^{+\infty } \xi (u) y(u) \textrm{d}u \Bigg \vert {\mathcal {F}}_t\right) \ge 0, \end{aligned}$$
(13)
which is a ‘no-borrowing-without-repayment’ constraint [20]. As the second term appearing in (13) represents the agent’s human capital at time t, constraint (13) captures the situation in which human capital can be pledged as collateral and represents the agent’s maximum borrowinge capacity. In line with [7], we note that the triplets \(\left( c,\theta ,B\right) \) must be understood as representing the pre-death counterparts of the controls predictable relative to the reference filtration \({\mathbb {F}}\). Similarly, on the trace of \({\mathbb {F}}\) on \(\{\tau ^\delta >t\}\), the process \(\xi \) satisfies equation (4) with a drift of the form \(-\xi (t)(r+\delta )\), as explained in Remark 2.4.
In [7], the authors find an explicit solution to the optimization problem under power utility. The proof of the result relies on the resolution of an infinite-dimensional Hamilton-Jacobi-Bellman (HJB) equation, which can be considered as an infinite-dimensional version of the classical Merton problem. From a technical point of view, the key idea is to extend the state space so as to include the past trajectory of y over \([-d,0]\). In this way, the problem becomes infinite dimensional and Markovian (in the current wealth level and path of y over the time window \([-d,0]\)). In this infinite-dimensional reformulation of the problem, it becomes essential to rewrite the constraint (13) by exploiting the explicit expression given in Theorem 2.3. Importantly, the optimal risky asset allocation found in [7],
$$\begin{aligned} \theta ^{*}(t) := (\sigma \sigma ^\top )^{-1}(\mu -r {\textbf{1}}) \frac{W^*(t)+H(t)}{ \gamma } -\sigma ^{-1} \sigma _y \frac{1}{K}y(t), \end{aligned}$$
(14)
reveals that the decomposition of human capital into a past and future component is essential in understanding the agent’s hedging demand, as we now discuss. We first note that the solution follows the logic of Merton’s, in that the agent chooses constant fractions of total wealth (given by financial wealth and human capital). In line with [9, 20], the agent considers the capitalized value of future wages as if they were a traded asset. As the agent’s labor income is instantaneously perfectly correlated with the risky assets, a negative income hedging demand arises (the term \(\sigma ^{-1}\sigma _y K^{-1}y(t)\)), and the allocation to risky assets is reduced by a term proportional to the regression coefficient of labor income shocks on risky asset returns ([13]). The key difference with the benchmark model with no delay is that the hedging demand only depends on the ‘future component’ of human capital, and not on the capitalized market value of the labor income’s past trajectory. The intuition is that when wages respond to market shocks with a delay, then human capital is more predictable and only the portion of human capital exposed to future market shocks drives the negative income hedging demand.

3.1.2 Extensions: robustness and finite horizon

The authors in [6] study a robust version of the lifecycle optimal portfolio choice problem presented in Sect. 3.1.1. Again, the result in Theorem 2.3 is essential in obtaining an explicit solution of the problem considered.
The authors in [5] consider the same problem as in the previous section, but now with a finite time horizon, which can be interpreted as the agent’s fixed retirement date. The authors extend the results of Theorem 2.3 to the finite horizon case, obtaining again a decomposition of human capital into two components pertaining to the past and the future evolution of labor income. They then proceed to solve the lifecycle portfolio choice model, which entails time-dependent state constraints, as horizon effects are now material during the working life of the agent. This appears to be a novel type of problem in the infinite dimensional stochastic control literature, which is again successfully solved by exploiting the structure of the explicit solution for human capital.

3.1.3 Mean-field games

Another generalization of the problem addressed in [7] is considered in [16]. Similarly to [7], the authors consider a lifecycle optimal portfolio choice problem faced by an agent receiving labor income and allocating her wealth to risky securities and a riskless asset subject to a borrowing constraint. However, in addition to assuming the dynamics of labor income to adjust slowly to financial market shocks, they also assume it to be benchmarked to the wages of a population of agents with comparable tasks and/or ranks. Specifically, each agent i’s labor income \(y_i\) is benchmarked against the wages \(y^n:=(y_1,y_2,\ldots ,y_n)\) of a population of n agents. As n grows larger, the problem falls into the family of optimal control of infinite dimensional McKean-Vlasov dynamics type. By adding a suitable new variable, the authors simplify the problem and are able to find explicitly the solution of the associated infinite-dimensional HJB equation and the optimal feedback controls. A necessary step to solve the problem is to provide a suitable reformulation of the no-borrowing without repayment constraint (13), where now labor income obeys an SDDE with a drift containing not only a path-dependent term but also a mean reverting term. This is carried out in [16, Section 3], where the authors provide a generalization of formula (8) by carefully adapting the techniques used in Sect. 4 of this paper.

3.2 Counterparty risk and derivatives valuation

As a simple example of application of our setup to the context of over-the-counter derivatives, let us consider in equation (2) the case of \(n=1\), \(\mu _0=0\), \(\phi =0\), \(\sigma _0=0\), and \( \varphi (s) =\delta _{-d}(s)\), where \(\delta _{a}(s)\) indicates the delta-Dirac measure at a, so that equation (2) now reads
$$\begin{aligned} \text {d}X_0(t) = X_0(t-d) \text {d} Z(t). \end{aligned}$$
(15)
Then, for \(t \in [0,d )\) we have
$$\begin{aligned} X_0(t) = x_0+ \int _0^t X_0(s-d) \text {d} Z(s)= x_0+ \int _{-d}^{t-d} x_1(\tau ) \text {d}Z(\tau +d) . \end{aligned}$$
(16)
In this case \(X_0(t)\) is Gaussian, and dynamics (15) could be used to model, for example, the variation margin of an over-the-counter swap, when the collateralization procedure relies on a delayed mark-to-market value of the instrument (see [11], page 316, or [8, 12] for some examples).

4 Proof of the result

Within this section we consider the unique continuous \({\mathbb {F}}\)-adapted solution y of (2) given in Proposition 2.1. The proof of Theorem 2.3 can be divided in the following steps:
  • we incorporate the discount factor \(\xi \) in the equivalent risk-neural probability measure \(\tilde{\mathbb {P}}\) and rewrite the dynamics of y under \(\tilde{\mathbb {P}}\). Derive the deterministic delayed equation satisfies by the quantity \(\tilde{\mathbb {E}}[y(t)|\mathcal {F}_{t_0}]\). (Sect. 4.1).
  • We rewrite the delayed equation for \(\tilde{\mathbb {E}}[y(t)|\mathcal {F}_{t_0}]\) as a deterministic evolution equation, which takes values in a suitable Hilbert space incorporatingthe past and the present in its structure. We will appeal to the so-called product-space framework for path-dependent equations (Sect. 4.2).
  • We exploit suitable spectral properties of the operator that appears in the above mentioned infinite-dimensional formulation of the problem in order to obtain expression (8) for \(H(t_0)\) (Sects. 4.3 and 4.4).
  • We clarify the relationship between the spectral properties used and our Assumption 2.2 (Sect. 4.5).
The above first three steps will lead to Proposition 4.8, whereas the last step will be formalized in Lemmas 4.10 and 4.11. Theorem 2.3 will then follow as an immediate consequence of these results. For readability, we relegate to an “Appendix” the proofs of some technical lemmas.

4.1 Equivalent probability measure

We find it more convenient to change perspective from a valuation formula relying on the use of the stochastic discount factor \(\xi \) under the measure \({\mathbb {P}}\) to the one discounting at the risk-free rate r the equivalent martingale measure \(\tilde{\mathbb {P}}\), which is therefore called ‘risk-neutral measure’. It is well know that the two approaches are equivalent under no arbitrage in our setting.
We start by considering the equivalent probability measure \({\tilde{{\mathbb {P}}}}\) on \(\mathcal {F}_s\) such that2
$$\begin{aligned} \frac{\text {d}\tilde{{\mathbb {P}}}}{\text {d} {\mathbb {P}}}\bigg \vert _{{\mathcal {F}}_s}=\exp \left( -\frac{1}{2} |\kappa |^2 s - \kappa \cdot Z(s)\right) = e^{rs} \xi (s). \end{aligned}$$
(17)
By [27, Lemma 3.5.3] we can write
$$\begin{aligned} {\mathbb {E}} \left[ \xi (s) y(s) \mid {\mathcal {F}}_{t_0} \right] = \xi (t_0) e^{-r(s-t_0)} \tilde{{\mathbb {E}}}_s \left[ y(s) \mid {\mathcal {F}}_{t_0} \right] , \end{aligned}$$
and thus3
$$\begin{aligned} \begin{aligned} \int _{t_0}^{+\infty } {\mathbb {E}} \left[ \xi (s) y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s= \xi (t_0) e^{rt_0} \int _{t_0}^{+\infty } e^{-rs} \tilde{{\mathbb {E}}}_s \left[ y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s. \end{aligned} \end{aligned}$$
(18)
The idea is to now understand what kind of differential equation the quantity \(\tilde{{\mathbb {E}}}\left[ y(s)\mid {\mathcal {F}}_{t_0} \right] ={\tilde{{\mathbb {E}}}}_s\left[ y(s)\mid {\mathcal {F}}_{t_0} \right] \) satisfies. Let \(\tilde{{\mathbb {P}}}\) the measure such that \(\left. \tilde{{\mathbb {P}}}\right| _{{\mathcal {F}}_s}=\tilde{{\mathbb {P}}}(s)\) for all \(s\ge 0\). By the Girsanov Theorem the process \({\tilde{Z}}(t) = Z(t) + \kappa t\) is an n-dimensional Brownian motion under \( \tilde{{\mathbb {P}}}\). The dynamics of y under \( \tilde{{\mathbb {P}}}\) is then
$$\begin{aligned} dy(s)= & {} \big [ (\mu _y - \sigma _y \cdot \kappa ) y(s) + \int _{-d}^0 y(s+\tau ) \,\Phi (\text {d}\tau ) \big ] \text {d}s\nonumber \\{} & {} \quad \quad +\left[ y(t)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(s+\tau ) \varphi _1(\text {d}\tau ) \\ \vdots \\ \int _{-d}^0 y(s+\tau ) \varphi _n(\text {d}\tau ) \end{pmatrix} \right] \cdot \text {d} {\tilde{Z}}(s), \end{aligned}$$
(19)
where \(\Phi \) is defined in (6). Integrating between \(t_0\) and t we obtain
$$\begin{aligned} \begin{aligned} y(t)&= y(t_0) + \int _{t_0}^t (\mu _y - \sigma _y \cdot \kappa ) y(s) \text {d}s + \int _{t_0}^t \int _{-d}^0 y(s+\tau )\Phi (\text {d}\tau ) \text {d}s\\&\quad + \int _{t_0}^t \left[ y(s)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(s+\tau ) \varphi _1(\text {d}\tau ) \\ \vdots \\ \int _{-d}^0 y(s+\tau ) \varphi _n(\text {d}\tau ) \end{pmatrix} \right] \cdot \text {d}{\tilde{Z}}(s)\,, \end{aligned} \end{aligned}$$
(20)
and therefore, by taking the conditional expected value on both sides, we get
$$\begin{aligned} \begin{aligned} \tilde{{\mathbb {E}}} \left[ y(t)\mid {\mathcal {F}}_{t_0} \right] =&y(t_0) + (\mu _y - \sigma _y\cdot \kappa ) \tilde{{\mathbb {E}}} \left[ \int _{t_0}^t y(s) \text {d}s \mid {\mathcal {F}}_{t_0} \right] \\&+ \tilde{{\mathbb {E}}} \left[ \int _{t_0}^t \int _{-d}^0 y(s+ \tau ) \Phi (\text {d} \tau ) \text {d}s \mid {\mathcal {F}}_{t_0} \right] \\&+ \tilde{{\mathbb {E}}} \left[ \int _{t_0}^t \left[ y(s)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(s+\tau )\varphi _1(\text {d}\tau ) \\ \vdots \\ \int _{-d}^0 y(s+\tau ) \varphi _n(\text {d}\tau ) \end{pmatrix} \right] \cdot \text {d}{\tilde{Z}}(s) \mid {\mathcal {F}}_{t_0} \right] . \end{aligned} \end{aligned}$$
(21)
The following Lemma guarantees that the stochastic integral with respect to \({\tilde{Z}} \) is a martingale, and has zero mean. The proof is provided in “Appendix A.2”.
Lemma 4.1
It holds that
$$\begin{aligned} \tilde{\mathbb {E}}\left[ \int _{t_0}^t \left\| y(s)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(s+\tau )\varphi _1(\textrm{d}\tau )\\ \vdots \\ \int _{-d}^0 y(s+\tau ) \varphi _n(\textrm{d}\tau ) \end{pmatrix} \right\| ^2\, \textrm{d} s \right] < +\infty \,. \end{aligned}$$
We thus obtain that
$$\begin{aligned} \tilde{{\mathbb {E}}} \left[ \int _{t_0}^t \left[ y(s)\sigma _y + \begin{pmatrix} \int _{-d}^0 y(s+\tau )\varphi _1(\text {d}\tau ) \\ \vdots \\ \int _{-d}^0 y(s+\tau ) \varphi _n(\text {d}\tau ) \end{pmatrix} \right] \cdot \text {d}{\tilde{Z}}(s) \mid {\mathcal {F}}_{t_0} \right] =0, \end{aligned}$$
and, by definition of conditional mean and by Fubini’s Theorem, the expression in (21) reduces to
$$\begin{aligned} \begin{aligned} \tilde{{\mathbb {E}}} \left[ y(t)\mid {\mathcal {F}}_{t_0} \right] = \ {}&y(t_0) + (\mu _y - \sigma _y\cdot \kappa ) \int _{t_0}^t \tilde{{\mathbb {E}}} \left[ y(s) \mid {\mathcal {F}}_{t_0} \right] \, \textrm{d}s\\&+ \int _{t_0}^t \int _{-d}^0 \tilde{{\mathbb {E}}} \left[ y(s+ \tau ) \mid {\mathcal {F}}_{t_0} \right] \, \Phi (\text {d} \tau ) \text {d}s. \end{aligned} \end{aligned}$$
(22)
Therefore, defining
$$\begin{aligned} M_{t_0}(t):=\tilde{\mathbb {E}}\left[ y(t)|\mathcal {F}_{t_0}\right] , \end{aligned}$$
(23)
we have that \(M_{t_0}\) satisfies for \(t\ge t_0\) the equation (with random initial conditions)
$$\begin{aligned} {\left\{ \begin{array}{ll} \textrm{d}M_{t_0}= [(\mu _y - \sigma _y \cdot \kappa ) M_{t_0}(t)+\int _{-d}^0 M_{t_0}(t+s)\, \Phi (\text {d} s)]\, \textrm{d}t,\\ M_{t_0}(t_0)=y(t_0),&{}\\ M_{t_0}(t_0+s)=y(t_0+s), \ \quad \qquad s \in [-d,0). \end{array}\right. } \end{aligned}$$
(24)
Existence of a unique solution of the above system is guaranteed by the following generalization of [4, Part II, Chapter 4, Theorem3.2] to random initial conditions.
Lemma 4.2
Given any fixed \(\mathcal {F}_{t_0}\)-measurable \(\mathbb {R} \times L^2([-d,0];\mathbb {R})\)-valued random variable \(m=(m_0,m_1)\), the Cauchy problem
$$\begin{aligned} {\left\{ \begin{array}{ll} \textrm{d} m(t_0;t)= [(\mu _y - \sigma _y \cdot \kappa ) m(t_0;t)+\int _{-d}^0 m(t_0;t+s)\, \Phi (\textrm{d} s)]\, \textrm{d}t,\\ m(t_0;t_0)=m_0,&{}\\ m(t_0;t_0+s)=m_1(s), \ \quad \qquad s \in [-d,0). \end{array}\right. } \end{aligned}$$
(25)
admits a unique absolutely continuous solution. Moreover, system (25) is equivalent to (24) when we choose \((m_0,m_1)=(y(t_0),y(t_0+\cdot ))\).

4.2 Reformulation of the problem in an infinite-dimensional framework

We now reformulate the differential equation with delay (24) as an evolution equation with values in the so called Delfour-Mitter Hilbert space, defined as
$$\begin{aligned} \mathcal {H}:={\mathbb {R}}\times L^2(-d,0;{\mathbb {R}}), \end{aligned}$$
whose elements are denoted as \( x=(x_0,x_1)\). \(\mathcal {H}\) is a Hilbert space when endowed with the inner product \(\langle (x_0,x_1),(y_0,y_1)\rangle _{\mathcal {H}}=x_0y_0+\langle x_1,y_1\rangle \), the latter being the usual inner product of \(L^2(-d,0;{\mathbb {R}})\).
We define the operator \(A:\mathcal {D}(A) \subset \mathcal {H} \rightarrow \mathcal {H}\) as
$$\begin{aligned} {\mathcal {D}}\left( A\right) :=\left\{ (x_0,x_1) \in {\mathcal {H}}: x_1(\cdot ) \in W^{1,2}\big ( [-d, 0]; {\mathbb {R}} \big ),\, x_0 = x_1(0)\right\} , \end{aligned}$$
$$\begin{aligned} A(x_0,x_1):=\Big ( (\mu _0 - \sigma _0\cdot \kappa ) x_0+\int _{-d}^0 x_1(s)\Phi (\text {d}s), \frac{\text {d}}{\text {d}s} x_1 \Big ), \end{aligned}$$
(26)
with \(\Phi \) defined in (6). Here by \(W^{1,2}\big ( [-d, 0]; {\mathbb {R}} \big )\) we denote the Sobolev space of weakly differentiable square-integrable functions.
We can then reformulate equation (24) as an evolution equation in \(\mathcal {H}\).
Consider, again for any fixed \(\mathcal {F}_{t_0}\)-measurable \(\mathcal {H}\)-valued random variable \({\textbf{m}}=\left( m_0,m_1\right) \), the \(\mathcal {H}\)-valued process \({\textbf{M}}(t_0;\cdot )\) that is the solution on \([t_0,+\infty )\) of
$$\begin{aligned} {\left\{ \begin{array}{ll} \textrm{d} {\textbf{M}}(t_0;t) = A {\textbf{M}}(t_0;t) \textrm{d}t,\\ {\textbf{M}}(t_0;t_0) ={\textbf{m}}. \end{array}\right. } \end{aligned}$$
(27)
We collect in the following Proposition some useful results about the above equation (for more details see e.g. [14, AppendixA]).
Proposition 4.3
(i)
The operator A generates a strongly continuous semigroup \(\left\{ S(t)\right\} _{t\ge 0}\) in \(\mathcal {H}\).
 
(ii)
S(t) is a compact operator for every \(t\ge d\).
 
(iii)
For every \(\mathcal {F}_{t_0}\)-measurable \(\mathcal {H}\)-valued random variable m the process
$$\begin{aligned} S(t-t_0){\textbf{m}} ; \end{aligned}$$
(28)
is the unique weak (in distributional sense) solution of (27); in particular
$$\begin{aligned} {\textbf{M}}(t_0;t)={\textbf{M}}(0;t-t_0)\ . \end{aligned}$$
(29)
 
(iv)
The Cauchy problem (27) is equivalent to (25).
 
Proof
See “Appendix A.3”. \(\square \)
As an immediate consequence of the above result we obtain the desired equivalence between equations (27) and (24).
Corollary 4.4
Let y be a solution of (2) on \([0,t_0]\); when choosing \({\textbf{m}}\) as \((m_0,m_1)= (y(t_0), y(t_0+\cdot ))\), (27) is equivalent to (24) and in this case we have
$$\begin{aligned} {\textbf{M}}(t_0;t)=S(t-t_0){\textbf{m}} =\left( m(t_0;t),m(t_0;t+\cdot )\right) =\left( M_{t_0}(t),\left\{ M_{t_0}(t+s)\right\} _{s\in [-d,0]}\right) . \end{aligned}$$
From now on we thus will work with formulation (27). The spectral properties of the operator A, that appears in this infinite-dimensional formulation, will be crucial to prove our result. We devote the next section to the analysis of these properties.

4.3 Spectral properties of A

In the present section we collect some technical results concerning the spectral properties of the operator A. Proof of Theorem 2.3 is based on the Lemmas presented here. The technical proofs are postponed to the “Appendix”.
Lemma 4.5
The spectrum of the operator A is given by
$$\begin{aligned} \sigma (A)=\{\lambda \in \mathbb {C}:K(\lambda )=0\}, \end{aligned}$$
where
$$\begin{aligned} K(\lambda ):=\lambda - (\mu _y- \sigma _y \cdot \kappa ) - \int _{-d}^0 e^{\lambda \tau } \Phi (\textrm{d}\tau )\,,\quad \lambda \in {\mathbb {C}}\,. \end{aligned}$$
(30)
The spectrum \(\sigma \left( A\right) \) is a countable set and every \(\lambda \in \sigma \left( A\right) \) is an isolated eigenvalue of finite multiplicity.
The spectral bound of A is
$$\begin{aligned} \lambda _0=\sup \left\{ \textrm{Re}\,\lambda :\, K(\lambda )=0\right\} . \end{aligned}$$
(31)
Proof
See [23, Chapter 7, Lemma 2.1 and Theorem4.2]\(\square \)
We ca explicitly compute the resolvent operator of A.
Lemma 4.6
Let \(\rho (A)\) denote the resolvent set of A and let \(\lambda \in {\mathbb {R}}\cap \rho \left( A\right) \). The resolvent operator of A at \(\lambda \), denoted by \(R(\lambda ,A)\) is given by
$$\begin{aligned} R(\lambda , A) \left( m_0, m_1\right) = \left( u_0, u_1\right) \end{aligned}$$
(32)
with
$$\begin{aligned} \begin{aligned} u_0&=\frac{1}{K(\lambda )}\left[ m_0+ \int _{-d}^0\int _{-d}^{s}e^{-\lambda (s-\tau )}\Phi (\textrm{d}\tau ) \, m_1(s) \textrm{d} s\right] ,\\ u_1(s)&= u_0e^{\lambda s} +\int _s^0e^{-\lambda (\tau -s)}m_1(\tau )\, \textrm{d}\tau . \end{aligned} \end{aligned}$$
(33)
Proof
See “Appendix A.4”. \(\square \)
Lemma 4.7
For every real \(\lambda \) such that \(\lambda >\lambda _0\) and every \({\textbf{m}}=(m_0,m_1)\in \mathcal {H}\) we have
$$\begin{aligned} \int _0^{+\infty }e^{-\lambda t}S(t) {\textbf{m}}\, \textrm{d}t=R(\lambda ,A){\textbf{m}}. \end{aligned}$$
(34)
Proof
Identity (34) is well known to hold for all real \(\lambda \) larger than the type of S(t). Since S(t) is compact for every \(t\ge d\), its type is actually equal to its spectral radius \(\lambda _0\). For a reference see e.g. [4, Part II, Chapter 1, Corollary 2.5]. \(\square \)

4.4 Deriving the explicit formula for H

In this section we exploit the results derived in the above sections to prove the following.
Proposition 4.8
Assume \(r>\lambda _0\), then for any \(t_0\ge 0\), the quantity \(H(t_0)\) defined in (3) has the following explicit form
$$\begin{aligned} H(t_0)= \frac{1}{K} \left( y(t_0)+ \int _{-d}^0 G(s) y(t_0+s)\, \text {d}s\right) ,\quad {\mathbb {P}}-a.s., \end{aligned}$$
where \(y(t_0)\) denotes the solution at time \(t_0\) of equation (2),
$$\begin{aligned} K:=r-\mu _0+\sigma _y \cdot \kappa - \int _{-d}^0 e^{r\tau }\Phi (\textrm{d} \tau ), \end{aligned}$$
(35)
and G is given by
$$\begin{aligned} G(s) : = \int _{-d}^s e^{-r(s-\tau )} \Phi (\textrm{d}\tau ). \end{aligned}$$
(36)
Remark 4.9
Notice that the statement of the above result is the same of Theorem 2.3, but the assumptions here are different: we assume \(r>\lambda _0\) instead of Assumption 2.2. An explanation of why we do actually consider Assumption 2.2 in Theorem 2.3, will be provided in the next section.
Proof
Let \({\textbf{m}}=(m_0,m_1)= (y(t_0), y(t_0+\cdot ))\). We denote here by \(\Pi \) the projection on the first (finite-dimensional) component of \(\mathcal {H}\), i.e. \( \Pi [ {\textbf{m}}] =\Pi [(m_0,m_1)]=m_0\).
Starting from (18), we have
$$\begin{aligned}&\frac{1}{\xi (t_0)}\int _{t_0}^{+\infty } {\mathbb {E}} \left[ \xi (s) y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s =e^{rt_0}\int _{t_0}^{\infty }e^{-rs}\tilde{\mathbb {E}}\left[ y(s)|\mathcal {F}_{t_0}\right] \, \textrm{d}s \quad \quad \text {(by} (18)) \nonumber \\&=e^{rt_0}\int _{t_0}^{\infty }e^{-rs}M_{t_0}(s)\, \textrm{d}s \quad \quad \quad \quad \quad \text {(by} 23) \nonumber \\&=e^{rt_0}\int _{t_0}^{\infty }e^{-rs}\Pi \left[ {\textbf{M}}(t_0;s)\right] \, \textrm{d}s\quad \quad \quad \quad \text {(by Corollary}~4.4) \nonumber \\&=e^{rt_0}\int _{0}^{\infty }e^{-rt_0}e^{-rs}\Pi \left[ {\textbf{M}}(0;s)\right] \, \textrm{d}s\quad \quad \quad \quad \text {(by} (29)) \nonumber \\&= \int _0^{\infty }e^{-rs}\Pi \left[ S(s){\textbf{m}}\right] \,\textrm{d}s\quad \quad \quad \quad \text {(by (28))} \nonumber \\&=\Pi \left[ R(r,A){\textbf{m}}\right] \quad \quad \quad \quad \quad \text {(by Lemma}~4.7, \text {since}\,\, r>\lambda _0) \nonumber \\&= \frac{1}{K(r)}\left[ y(t_0)+\int _{-d}^0\int _{-d}^{s} e^{-r(s- \tau )} \Phi (\text {d}\tau ) \, y(t_0+s) \text {d} s\right] \quad \quad \quad \quad \quad \text {(by Lemma}~4.6). \end{aligned}$$
(37)
From the above equalities we infer, in particular, the \(\mathbb {P}\)-integrability of \( \int _{t_0}^{+\infty } {\mathbb {E}} [ \xi (s) y(s) \mid {\mathcal {F}}_{t_0} ]\,\textrm{d}s\), which justifies the equality
$$\begin{aligned} {\mathbb {E}} \left[ \int _{t_0}^{+\infty } \xi (s) y(s) \text {d}s \mid {\mathcal {F}}_{t_0} \right] =\int _{t_0}^{+\infty } {\mathbb {E}} \left[ \xi (s) y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s. \end{aligned}$$
(38)
In fact, by the characteristic property of the conditional mean, and Fubini’s Theorem we have that, for any \(F \in {\mathcal {F}}_{t_0}\)
$$\begin{aligned} \int _F \int _{t_0}^{+\infty } {\mathbb {E}} [ \xi (s) y(s) \mid {\mathcal {F}}_{t_0} ]\,\textrm{d}s \, \textrm{d}{\mathbb {P}}&= \int _{t_0}^{+\infty } \int _F {\mathbb {E}} [ \xi (s) y(s) \mid {\mathcal {F}}_{t_0} ] \textrm{d}{\mathbb {P}} \, \textrm{d}s\\&= \int _{t_0}^{+\infty } \int _F \xi (s) y(s) \, \textrm{d} {\mathbb {P}} \, \textrm{d}s = \int _F \int _{t_0}^{+\infty } \xi (s) y(s) \, \textrm{d}s \, \textrm{d} {\mathbb {P}} \\&=\int _F {\mathbb {E}} \left[ \int _{t_0}^{+\infty } \xi (s) y(s)\, \textrm{d}s \mid {\mathcal {F}}_{t_0} \right] \, \textrm{d} {\mathbb {P}}. \end{aligned}$$
Defining now \(K:=K(r)\) and recalling (3), (35) and (36), by (37) and (38), the result immediately follows. \(\square \)

4.5 Motivations for Assumption 2.2

In Proposition 4.8 we proved our main result under the Assumption \(r>\lambda _0\). This requirement is difficult to check in practice, since it requires an explicit computation of the spectral bound \(\lambda _0\). In the present section we therefore look for some sufficient conditions easier to check.
Set for \(\lambda \in \mathbb {C}\)
$$\begin{aligned} {\widetilde{K}}(\lambda ):=\lambda - (\mu _y- \sigma _y \cdot \kappa ) - \int _{-d}^0 e^{\lambda \tau } |\Phi |(\text {d}\tau ), \end{aligned}$$
(39)
where by \(|\Phi |\) we denote the total variation measure of \(\Phi \). Set
$$\begin{aligned} \widetilde{\lambda _0}=\sup \left\{ \textrm{Re}\,\lambda :\, {\widetilde{K}}(\lambda )=0\right\} . \end{aligned}$$
(40)
We note that \(\widetilde{\lambda _0}\) is the spectral radius of the operator \(\widetilde{A}: \mathcal {D}(\widetilde{A})\subset \mathcal {H} \rightarrow \mathcal {H}\) defined as follows:
$$\begin{aligned} \begin{aligned} \mathcal {D}(\widetilde{A}):= \left\{ \left( x_0,x_1\right) \in \mathcal {H}: x_1\in W^{1,2}(-d,0;\mathbb {R}), \ x_1(0)=x_0\right\} ,\\ \widetilde{A}\left( x_0,x_1\right) := \left( (\mu _y-\sigma _y\cdot \kappa )x_0+ \int _{-d}^0 x_1(s)|\Phi |(\textrm{d}s), \frac{\textrm{d}}{\textrm{d}s} x_1\right) . \end{aligned} \end{aligned}$$
Lemma 4.10
The function \({\widetilde{K}}\), restricted to the real numbers, is strictly increasing and the spectral bound \(\widetilde{\lambda _0}\) is the only real root of the equation \({\widetilde{K}}(\xi )=0\). In particular,
$$\begin{aligned} {\widetilde{K}}(r)>0 \ \iff r>\widetilde{\lambda _0}. \end{aligned}$$
(41)
Proof
See “Appendix A.5”. \(\square \)
Recall the definition of K given in (30) and the definition of the spectral bound of A, \(\lambda _0\), given (31).
Lemma 4.11
It holds
$$\begin{aligned} \widetilde{\lambda _0}\ge \lambda _0. \end{aligned}$$
Proof
See “Appendix A.6”. \(\square \)
Thanks to the above two Lemmas it becomes now clear why we work under Assumption 2.2 in Theorem 2.3. It provides a sufficient condition for the condition \(r>\lambda _0\), imposed in Proposition 4.8, to hold. In fact, assume \({\widetilde{K}}>0\) as in Assumption 2.2, then by Lemmas 4.10 and 4.11 we immediately get \(r>\lambda _0\).
Remark 4.12
Notice that, if \(\Phi \) is a positive measure, then \({\widetilde{K}} \equiv K\), \(\lambda _0 \equiv \widetilde{\lambda _0}\) and the condition \(K>0\) becomes also necessary, that is \(K>0 \iff r>\lambda _0\).

5 Conclusion

In this paper we have considered a complete market model in the spirit of Samuelson’s original contribution [36], in which security prices evolve as geometric Brownian motions. Despite working within a classical setup, we have obtained a novel explicit pricing formula for stream of payments with delayed dynamics, by using techniques from infinite-dimensional stochastic analysis. Our valuation formula results in an explicit expression demonstrating the importance of appreciating the past to quantify the current market value of the future. The approach followed in this paper highlights how tools from infinite-dimensional analysis can be successfully used to address valuation problems that are non-Markovian, and hence beyond the reach of conventional approaches. As highlighted in the applications discussed in Sect. 3, it is apparent how our results and the techniques developed here can be used successfully to explicitly solve interesting classes of infinite dimensional stochastic optimal control problems with nontrivial state constraints.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Anhänge

A Further technical details

A.1 Proof of Proposition 2.1

The existence and uniqueness result for (2) is not covered by the extant literature. When the initial datum \(x = (x_0,x_1)\), seen as a function on \([-d,0]\), is continuous, existence and uniqueness of the strong solution to the SDDE for y is proved by [34, Theorem I.2]. When the initial datum \(x\in {\mathbb {R}}\times L^2([-d,0),\textrm{d}t ;{\mathbb {R}})\), with the additional requirement that \(\phi \) and \((\varphi _1,\ldots ,\varphi _n)\) are absolutely continuous w.r.t the Lebesgue measure, that is \(\textrm{d}\phi = \varphi \textrm{d}t\), \(\textrm{d}(\varphi _1,\ldots ,\varphi _n)= (\phi _1,\ldots ,\phi _n)\,\textrm{d}t\) and the Radon-Nikodym densities \(\varphi , \phi _1,\ldots ,\phi _n \in L^2([-d,0), \textrm{d}t)\), the existence and uniqueness result follows by [34, Remark I.3(iv)]. We need to extend this latter result to the case in which \(\phi \) and \((\varphi _1,\ldots ,\varphi _n)\) are signed measures of bounded variation on \(\left[ -d,0\right] \). We will prove the result by means of the same procedure employed for the proof of [6, Proposition B.2]. There the authors prove the existence and uniqueness of the solution for an equation similar to (2), under more general assumptions on the measure \(\phi \), but with no delay in the diffusion term.
Let us start by introducing the standard notation for the past path at t of a (deterministic) function \(h:[-d,T] \rightarrow \mathbb {R}\), for \(0\le t\le T\), that is the function \(h_t\)
$$\begin{aligned} h_t(s) := h(t+s) \, \, \, \text { for } -d \le s \le 0. \end{aligned}$$
The past path of y at t for the realization \(\omega \) is thus \(y_t(s,\omega ): =y(t+s,\omega ) \, \, s \in [-d,0]\). As usual, we hide the dependence of the process on \(\omega \) and write the delay terms in the drift and in the diffusion as follows:
$$\begin{aligned} \int _{-d}^0y(t+s)\phi (\textrm{d}s) =\int _{-d}^0y_t(s)\phi (\textrm{d}s) \end{aligned}$$
(42)
and
$$\begin{aligned} \begin{pmatrix} \int _{-d}^0 y(t+s) \varphi _1(\text {d}s) \\ \vdots \\ \int _{-d}^0y(t+s) \varphi _n(\text {d}s) \end{pmatrix} = \begin{pmatrix} \int _{-d}^0 y_t(s) \varphi _1(\text {d}s) \\ \vdots \\ \int _{-d}^0y_t(s) \varphi _n(\text {d}s). \end{pmatrix} \end{aligned}$$
(43)
The delay parts in (2), given by (42) and (43) can be then expressed in terms of (an extension of) the following linear operators ok kernel type:
$$\begin{aligned} L:C([-d,0];\mathbb {R}) \rightarrow \mathbb {R}, \qquad Lf:=\int _{-d}^0f(s)\, \phi (\textrm{d}s), \end{aligned}$$
(44)
and
$$\begin{aligned} G:C([-d,0];\mathbb {R}) \rightarrow \mathbb {R}^n, \qquad Gf:= \begin{pmatrix} \int _{-d}^0 f(s) \varphi _1(\text {d}s) \\ \vdots \\ \int _{-d}^0 f(s) \varphi _n(\text {d}s) \end{pmatrix}. \end{aligned}$$
(45)
Since the operators L and G are well-defined only on the space of continuous functions \(C([-d,0]; \mathbb {R})\), when the initial datum does not belongs to \(C([-d,0]; \mathbb {R})\) but just to \(L^2([-d,0); \mathbb {R})\) problems may arise. In fact, consider the initial datum \((x_0, x_1) \in \mathbb {R} \times L^2(-d,0;\mathbb {R})\) and proceed by assuming that the solution to (2) exists. Denote the past path on the window \([t-d,t]\) by \(y_t:[-d,0] \rightarrow \mathbb {R}\), \(y_t(s):=y(t+s)\) a.e. \(t \ge 0\), \(s \in [-d,0]\). Then, for \(0 \le t <d\), the past path is
$$\begin{aligned} y_t(s)= {\left\{ \begin{array}{ll} y(t+s) &{}\text {if} \ -t \le s<0\\ x_1(s) &{}\text {if} -d \le s<-t. \end{array}\right. } \end{aligned}$$
which, in general, is not a continuous function, but only square integrable. Therefore, the operators L and G introduced in (44) and (45) cannot be applied to \(y_t\) since the integrals in (42) and (43) may not be well defined. The first issue is thus to show that L and G admit continuous extensions to the square integrable functions on \([-d,0]\), as made precise in the following lemma.
Lemma 5.1
Let \(L:C([-d,0];\mathbb {R}) \rightarrow \mathbb {R}\) and \(G: C([-d,0];\mathbb {R}) \rightarrow \mathbb {R}^n\) be the continuous and linear maps given in (44) and (45) respectively. Fix \(T>0\) and define on \(C([-d,T];\mathbb {R})\) the delay operators
$$\begin{aligned} \mathcal {L}(y)(t):=Ly_t, \qquad 0 \le t\le T,\\ \mathcal {G}(y)(t):=Gy_t, \qquad 0 \le t\le T. \end{aligned}$$
Then,
1.
the maps \(\mathcal {L}:C([-d,T];\mathbb {R}) \rightarrow L^2([0,T];\mathbb {R})\) and \(\mathcal {G}:C([-d,T];\mathbb {R}) \rightarrow L^2([0,T];\mathbb {R}^n)\) satisfy, respectively, the inequalities
$$\begin{aligned} \Vert \mathcal {L}(y)\Vert _{L^2([0,T];\mathbb {R})}\le |\phi |([-d,0]) \Vert y\Vert _{L^2([-d,T];\mathbb {R})}, \quad \forall y \in C([-d,T];\mathbb {R}). \nonumber \\ \end{aligned}$$
(46)
$$\begin{aligned} \Vert \mathcal {G}(y)\Vert _{L^2([0,T];\mathbb {R}^n)}\le \left( \sum _{i=1}^n [|\varphi _i|([-d,0])]^2\right) ^{\frac{1}{2}} \Vert y\Vert _{L^2([-d,T];\mathbb {R})}, \quad \forall y \in C([-d,T];\mathbb {R}).\nonumber \\ \end{aligned}$$
(47)
 
2.
\(\mathcal {L}\) and \(\mathcal {G}\) have \(L^2\)-norm continuous, linear extensions (still denoted by \(\mathcal {G}\) and \(\mathcal {L}\), respectively) to \(L^2([-d,T];\mathbb {R})\).
 
Proof
The proof follows the lines of [6, Lemma B.1] (see also [4, Part II, Chapter 4, Theorem3.3], but for the sake of completeness, we prove the result for the operator \(\mathcal {G}\). For the operator \(\mathcal {L}\) one follows the same reasoning.
1.
$$\begin{aligned} \Vert \mathcal {G}(y)\Vert _{L^2([0,T];\mathbb {R}^n)}&=\Vert Gy_{\cdot }\Vert _{L^2([0,T];\mathbb {R}^n)} = \sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1}\int _0^TGy_r \cdot h(r)\, \textrm{d}r\\&=\sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1}\int _0^T \begin{pmatrix} \int _{-d}^0 y_r(s) \varphi _1(\text {d}s) \\ \vdots \\ \int _{-d}^0y_r(s) \varphi _n(\text {d}s) \end{pmatrix} \cdot h(r)\, \textrm{d}r\\&= \sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1}\sum _{i=1}^{n} \int _0^T h_i(r)\int _{-d}^0 y_r(s)\varphi _i(\textrm{d}s)\, \textrm{d}r. \end{aligned}$$
We estimate the i-th component (\(i=1,\ldots ,n\)) of the above expression exploiting the Fubini Theorem and the Hölder inequality.
$$\begin{aligned} \int _0^Th_i(r) \int _{-d}^0y_r(s)\, \varphi _i(\textrm{d}s)\, \textrm{d}r&\le \int _0^T|h_i(r)| \int _{-d}^0|y(r+s)|\, |\varphi _i|(\textrm{d}s)\, \textrm{d} r\\&= \int _{-d}^0\int _0^T|h_i(r)||y(r+s)|\, \textrm{d}r\, |\varphi _i|(\textrm{d}s) \\&\le \int _{-d}^0\Vert h_i\Vert _{L^2([0,T];\mathbb {R})}\Vert y\Vert _{L^2([s,s+T];\mathbb {R})}\, |\varphi _i|(\textrm{d}s)\\&\le |\varphi _i|([-d,0])\Vert h_i\Vert _{L^2([0,T];\mathbb {R})}\Vert y\Vert _{L^2([-d,T];\mathbb {R})}, \end{aligned}$$
where for the last inequality we exploit the inclusion \([s, s+T] \subseteq [-d,T]\). Therefore, by means of the Hölder inequality we obtain
$$\begin{aligned}&\Vert \mathcal {G}(y)\Vert _{L^2([0,T];\mathbb {R}^n)} \le \sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1}\\&\quad \sum _{i=1}^{n}|\varphi _i|([-d,0])\Vert h_i\Vert _{L^2([0,T];\mathbb {R})}\Vert y\Vert _{L^2([-d,T];\mathbb {R})}\\&\quad \le \Vert y\Vert _{L^2([-d,T];\mathbb {R})}\sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1} \\&\quad \times \left( \sum _{i=1}^n [|\varphi _i|([-d,0])]^2\right) ^{\frac{1}{2}}\left( \sum _{i=1}^n \Vert h_i\Vert ^2_{L^2([0,T];\mathbb {R})}\right) ^{\frac{1}{2}}\\&\quad =\Vert y\Vert _{L^2([-d,T];\mathbb {R})}\sup _{h\in L^2([0,T];\mathbb {R}^n), \Vert h\Vert _{L^2}\le 1} \left( \sum _{i=1}^n [|\varphi _i|([-d,0])]^2\right) ^{\frac{1}{2}}\Vert h\Vert ^2_{L^2([0,T];\mathbb {R}^n)}\\&\quad \le \left( \sum _{i=1}^n [|\varphi _i|([-d,0])]^2\right) ^{\frac{1}{2}}\Vert y\Vert _{L^2([-d,T];\mathbb {R})}. \end{aligned}$$
 
2.
The existence of the bounded linear extension of \(\mathcal {L}\) and \(\mathcal {G}\) to \(L^2([-d,t];\mathbb {R})\) is a consequence of inequalities (46) and (47) and the fact that \(C([-d,T];\mathbb {R})\) is dense in \(L^2([-d,T];\mathbb {R})\).
 
\(\square \)
We are now ready to prove Proposition 2.1. We will use the following notation: if functions \(a, b \ge 0\) satisfy the inequality \(a \le C(A)b\) with a constant \(C(A)>0\) depending on the expression A, we write \(a \lesssim _A b\).
Proof
(Proof of Proposition 2.1) The proof of the result relies on a contraction type argument. The same argument has been used in the proof of [6, Proposition B.2]. There the authors consider a SDDE of type (2) with no delay in the diffusion term. On the other hand they work in a more general setting considering a measure valued process \(\phi \) in the delay integral of the drift term.
We provide here a sketch of the proof referring to [6] for more details. We will give just the details of the estimates concerning the delay integral in the diffusion term that is missing in [6].
Let us fix the initial condition \((x_0, x_1) \in \mathbb {R}^2 \times L^2([-d,0];\mathbb {R})\). Let \(T>0\), we introduce the space
$$\begin{aligned} S_T:=\{y \in C([0,T];\mathbb {R}):y(0)=x_0\}, \end{aligned}$$
endowed with the sup norm
$$\begin{aligned} \Vert y\Vert _{S_T} = \sup _{t \in [0,T]}|y(t)|. \end{aligned}$$
We consider the space \(L^p(\Omega ;S_T)\), \(p \ge 2\), endowed with the norm
$$\begin{aligned} \Vert y\Vert _{L^p(\Omega ;S_T)} = \left( \mathbb {E}\left[ \Vert y\Vert ^p_{S_T}\right] \right) ^{\frac{1}{p}}= \left( \mathbb {E}\left[ \sup _{t \in [0,T]}|y(t)|^p\right] \right) ^{\frac{1}{p}}. \end{aligned}$$
In the sequel we will denote by \(p':=\frac{p}{p-1}\) the conjugate exponent to p and by \(p^*:=\frac{p}{p-2}\) the conjugate exponent to \(\frac{p}{2}\). Given \(y \in L^p(\Omega ;S_T)\), let
$$\begin{aligned} F(y)(t)&:= x_0+ \mu _y \int _0^t y(r)\,\textrm{d}r + \int _0^t \mathcal {L}({\bar{y}}^{x_1})\, \textrm{d}r+ \int _0^t y(r)\sigma _y \cdot \textrm{d}Z(r)\nonumber \\&\quad \quad + \int _0^t\mathcal {G}({\bar{y}}^{x_1})\cdot \textrm{d}Z(r), \qquad 0 \le t \le T. \end{aligned}$$
(48)
Here \(\mathcal {L}\) and \(\mathcal {G}\) are the continuous linear operators introduced in Lemma 5.1 and \({\bar{y}}^{x_1} \in L^p(\Omega ;L^2([-d,T];\mathbb {R}) )\) is defined as follows:
$$\begin{aligned} \bar{y}^{x_1}(t) = {\left\{ \begin{array}{ll} x_1(t), &{} \text {if } -d\le t< 0; \\ y(t), &{} \text {if } \ \ 0\le t\le T. \end{array}\right. } \end{aligned}$$
(49)
We aim at proving that F maps \(L^p(\Omega ;S_T)\) into itself for any \(p\ge 2\) and that it is a contraction on the same space when \(p>4\).
Let us start by proving that F maps \(L^p(\Omega , S_T)\), \(p\ge 2\), into itself. We write
$$\begin{aligned}&\Vert F(y)\Vert _{L^p(\Omega ;S_T)} \le |x_0|+ |\mu _y|\left\| \int _0^{\cdot }y(r)\, \textrm{d}r\right\| _{L^p(\Omega ;S_T)} \nonumber \\&\quad + \left\| \int _0^{\cdot } \mathcal {L}({\bar{y}}^{x_1})\, \textrm{d}r \right\| _{L^p(\Omega ;S_T)} +\left\| \int _0^{\cdot } y(r)\sigma _y \cdot \, \textrm{d}Z(r)\right\| _{L^p(\Omega ;S_T)}\nonumber \\&\quad + \left\| \int _0^{\cdot } \mathcal {G}({\bar{y}}^{x_1}) \cdot \, \textrm{d}Z(r)\right\| _{L^p(\Omega ;S_T)}. \end{aligned}$$
(50)
The boundedness of the terms that appears in the r.h.s. of (50), except the last one, can be proved following the lines of [6, Proposition B.2]. We estimate the last term in the r.h.s. of (50) by means of the Burkholder-Davies-Gundy inequality
$$\begin{aligned} \left\| \int _0^{\cdot } \mathcal {G}({\bar{y}}^{x_1}) \cdot \, \textrm{d}Z(r)\right\| _{L^p(\Omega ;S_T)}^p&= \mathbb {E} \left[ \sup _{t \in [0,T]}\left| \int _0^t \mathcal {G}({\bar{y}}^{x_1})\cdot \textrm{d}Z(r)\right| ^p \right] \\&\lesssim \mathbb {E} \left[ \left| \int _0^T\Vert \mathcal {G}({\bar{y}}^{x_1})\Vert ^2\, \textrm{d}r\right| ^{\frac{p}{2}} \right] \\&= \mathbb {E} \left[ \Vert \mathcal {G}({\bar{y}}^{x_1})\Vert ^p_{L^2([0,T];\mathbb {R}^n)}\right] \lesssim \mathbb {E} \left[ \Vert {\bar{y}}^{x_1}\Vert ^p_{L^2([-d,T];\mathbb {R})}\right] \\&=\Vert {\bar{y}}^{x_1}\Vert ^p_{L^p(\Omega ;L^2([-d,T];\mathbb {R}))} < \infty , \end{aligned}$$
where in the last inequality we exploited (47) of Lemma 5.1.
Let us now prove that, for \(p>4\), F defines a contraction in \(L^p(\Omega , S_T)\). We endow this space by the equivalent norm
$$\begin{aligned} \Vert y\Vert _{\alpha } := \left( \mathbb {E} \left[ \sup _{t \in [0,T]}\left( e^{-\alpha t}|y(t)|\right) ^p\right] \right) ^{\frac{1}{p}}, \end{aligned}$$
(51)
where \(\alpha >0\) will be chosen later on. Once we proved that F defines a contraction, by the Banach fixed point Theorem, we can infer the existence of a unique \(y \in L^p(\Omega ;S_T)\) such that \(y=F(y)\), i.e.
$$\begin{aligned} y(t)&= x_0+ \mu _y \int _0^t y(r)\,\textrm{d}r + \int _0^t \mathcal {L}({\bar{y}}^{x_1})\, \textrm{d}r+ \int _0^t y(r)\sigma _y \cdot \textrm{d}Z(r)\\&\quad + \int _0^t\mathcal {G}({\bar{y}}^{x_1})\cdot \textrm{d}Z(r), \qquad 0 \le t \le T, \qquad \mathbb {P}-a.s., \end{aligned}$$
and this will conclude the proof.
Given \(y, z \in L^p(\Omega ;S_T)\), from (48) and (51), we have
$$\begin{aligned} \Vert F(z)-F(y)\Vert _{\alpha }^p \lesssim _p&\mathbb {E} \left[ \sup _{t \in [0,T]} e^{-p\alpha t}\left( |\mu _y|\left| \int _0^{t}(z(r)-y(r))\textrm{d}r\right| ^p+ \left| \int _0^{t} \mathcal {L}({\bar{z}}^{x_1}-{\bar{y}}^{x_1})\, \textrm{d}r \right| ^p \right) \right] \nonumber \\&+\mathbb {E}\left[ \sup _{t \in [0,T]} e^{-p\alpha t}\left| \int _0^{t} (z(r)-y(r))\sigma _y \cdot \, \textrm{d}Z(r)\right| ^p\right] \nonumber \\&+\mathbb {E}\left[ \sup _{t \in [0,T]} e^{-p\alpha t}\left| \int _0^{t} \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1})\cdot \, \textrm{d}Z(r)\right| ^p\right] . \end{aligned}$$
(52)
We can estimate the first three terms in the r.h.s. of (52) proceeding as in [6, Proposition B.2]4. For the first term we obtain
$$\begin{aligned}&\mathbb {E}\left[ \sup _{t \in [0,T]} e^{-p\alpha t} |\mu _y|\left| \int _0^{t}(z(r)-y(r))\textrm{d}r\right| ^p\right] \nonumber \\&\quad \le |\mu _y|T\left( \frac{1}{\alpha p'}\right) ^{\frac{p}{p'}}\Vert z-y\Vert ^p_{\alpha } \lesssim _{\mu _y,T, p} C_1(\alpha )\Vert z-y\Vert ^p_{\alpha }. \end{aligned}$$
(53)
For the second term we get
$$\begin{aligned}{} & {} \mathbb {E} \left[ \sup _{t \in [0,T]} e^{-p\alpha t} \left| \int _0^{t} \mathcal {L}({\bar{z}}^{x_1}-{\bar{y}}^{x_1})\, \textrm{d}r \right| ^p \right] \nonumber \\{} & {} \quad \quad \le \left( \frac{|\phi |([-d,0])}{\alpha p'} \right) ^{\frac{p}{p'}}T |\phi |([-d,0])\Vert z-y\Vert ^p_{\alpha } \lesssim _{|\phi |, p, T}C_2(\alpha )\Vert z-y\Vert ^p_{\alpha }. \end{aligned}$$
(54)
For the third term, by means of the so called factorization method (see e.g. [?, Section 5.3]), for a given \(\delta \in \left( \frac{1}{p}, \frac{1}{2}\right) \)5, we have
$$\begin{aligned} \mathbb {E}&\left[ \sup _{t \in [0,T]} e^{-p\alpha t}\left| \int _0^{t} (z(r)-y(r))\sigma _y \cdot \, \textrm{d}Z(r)\right| ^p\right] \nonumber \\&\lesssim _{p, \delta } \left( \int _0^T u^{p'(\delta -1)}e^{-p'\alpha u}\, \textrm{d}u \right) ^{\frac{p}{p'}}T \Vert \sigma _y\Vert ^p\left( \sup _{u \in [0,T] }\int _0^u(u-r)^{-2\delta }e^{-2\alpha (u-r)}\, \textrm{d}r\right) ^{\frac{p}{2}}\nonumber \\&\quad \Vert z-y\Vert _{\alpha }^p \lesssim _{p,\delta , T, \Vert \sigma _y\Vert } C_3(\alpha )\Vert z-y\Vert ^p_{\alpha }. \end{aligned}$$
(55)
Let us now come to the estimate of the fourth term in (52). Exploiting the factorization method, for \(\eta \in \left( \frac{1}{p}, \frac{p-2}{2p}\right) \)6 we rewrite that stochastic integral as follows
$$\begin{aligned} \int _0^t \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1}) \cdot \,{ \mathrm d}Z(r)=c_{\eta }\int _0^t (t-u)^{\eta -1} Y(u)\, \textrm{d}u, \end{aligned}$$
with
$$\begin{aligned} \frac{1}{c_{\eta }}:= \int _r^t (t-u)^{\eta -1}(u-r)^{-\eta }\, \textrm{d}u = \frac{\pi }{\sin (\pi \eta )}, \end{aligned}$$
and
$$\begin{aligned} Y(u)=\int _0^u (u-r)^{-\eta }\mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1}) \cdot \textrm{d}Z(r). \end{aligned}$$
Thanks to the Hölder inequality we estimate
$$\begin{aligned} e^{-\alpha t}\left| \int _0^t \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1})\cdot \textrm{d}Z(r)\right|&= c_{\eta } e^{-\alpha t}\left| \int _0^t (t-u)^{\eta -1}Y(u)\, \textrm{d}u \right| \\&= c_{\eta } \left| \int _0^t e^{-\alpha (t-u)}(t-u)^{\eta -1}e^{-\alpha u}Y(u)\, \textrm{d}u \right| \\&\le c_{\eta }\left( \int _0^te^{-\alpha p'(t-u)}(t-u)^{p'(\eta -1)}\, \textrm{d}u\right) ^{\frac{1}{p'}} \\&\quad \left( \int _0^te^{-\alpha p u}|Y(u)|^p\, \textrm{d}u\right) ^{\frac{1}{p}}. \end{aligned}$$
Therefore we obtain
$$\begin{aligned} \mathbb {E}{} & {} \left[ \sup _{t \in [0,T]}e^{-\alpha p t}\left| \int _0^t \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1}) \cdot \textrm{d}Z(r) \right| ^p \right] \\{} & {} \le c_{\eta }^p \mathbb {E}\left[ \sup _{t \in [0,T]} \left( \int _0^te^{-\alpha p'(t-u)}(t-u)^{p'(\eta -1)}\, \textrm{d}u\right) ^{\frac{p}{p'}} \left( \int _0^te^{-\alpha p u}|Y(u)|^p\, \textrm{d}u\right) \right] \\{} & {} \le c^p_{\eta } \left( \int _0^Te^{-\alpha p'u}u^{p'(\eta -1)}\, \textrm{d}u\right) ^{\frac{p}{p'}} \int _0^T e^{-\alpha pu} \mathbb {E} \left[ |Y(u)|^p\right] \, \textrm{d}u. \end{aligned}$$
Now, recalling the definition of \(\mathcal {G}\) and that, when \(r<d\), \({\bar{z}}^{x_1}_r(s)-{\bar{y}}^{x_1}_r(s)=0\) for \(s \in [-d,-r)\) (see (49)), by means of the Burkholder-Davis-Gundy (BDG) and the Hölder (H) inequalities, we obtain for all \(u \in [0,T]\),
$$\begin{aligned} e^{-\alpha pu}{} & {} \mathbb {E}\left[ |Y(u)|^p\right] = e^{-\alpha pu} \mathbb {E}\left[ \left| \int _0^u (u-r)^{-\eta }\mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1}) \cdot \textrm{d}Z(r)\right| ^p \right] \\{} & {} \quad \overset{BDG}{\lesssim _p}\ e^{-\alpha up} \mathbb {E}\left[ \left| \int _0^u (u-r)^{-2\eta }\Vert \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1})\Vert ^2\, \textrm{d}r\right| ^{\frac{p}{2}}\right] \\{} & {} \quad \quad =e^{-\alpha up}\mathbb {E}\left[ \left| \int _0^u (u-r)^{-2\eta } \sum _{i=1}^n \left| \int _{-d}^0({\bar{z}}_r^{x_1}-{\bar{y}}_r^{x_1})(s)\, \varphi _i(\textrm{d}s) \right| ^2 \, \textrm{d}r\right| ^{\frac{p}{2}}\right] \\{} & {} \quad \quad =e^{-\alpha up}\mathbb {E}\left[ \left| \int _0^u (u-r)^{-2\eta } \sum _{i=1}^n \left| \int _{-d\vee -r}^0\left( (z(r+s)-y(r+s)\right) \, \varphi _i(\textrm{d}s) \right| ^2 \, \textrm{d}r\right| ^{\frac{p}{2}}\right] \\{} & {} \quad \quad =\mathbb {E}\left[ \left| \int _0^u (u-r)^{-2\eta }e^{-2\alpha (u-r-s)}e^{-2\alpha (r+s)}\sum _{i=1}^n \right. \right. \\{} & {} \qquad \left. \left. \left| \int _{-d\vee -r}^0\left( z(r+s)-y(r+s)\right) \, \varphi _i(\textrm{d}s) \right| ^2 \, \textrm{d}r\right| ^{\frac{p}{2}}\right] \\{} & {} \quad \quad \overset{H}{\le }\mathbb {E}\left[ \left| \int _0^u (u-r)^{-2\eta }e^{-2\alpha (u-r-s)}e^{-2\alpha (r+s)}\sum _{i=1}^n |\varphi _i|([-d,0]) \right. \right. \\{} & {} \quad \left. \left. \int _{-d\vee -r}^0\left| (z(r+s)-y(r+s)\right| ^2\, \varphi _i(\textrm{d}s) \, \textrm{d}r\right| ^{\frac{p}{2}}\right] \\{} & {} \quad \overset{H}{\le }\left( \sum _{i=1}^n \left( |\varphi _i|([-d,0]) \right) ^{p^*}\int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\{} & {} \qquad \quad \ \mathbb {E} \left[ \sum _{i=1}^n\int _0^u \int _{-d \vee -r}^0 e^{-\alpha p(r+s)}|z(r+s)-y(r+s)|^p\, \varphi _i(\textrm{d}s)\, \textrm{d}r \right] \\{} & {} \quad \quad \le \left( \sum _{i=1}^n \left( |\varphi _i|([-d,0]) \right) ^{p^*}\int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\{} & {} \qquad \sum _{i=1}^n\int _0^u \int _{-d \vee -r}^0\mathbb {E} \left[ \sup _{(r+s) \in [0,u]}\left( e^{-\alpha p(r+s)}|z(r+s)-y(r+s)|^p\right) \right] \, \varphi _i(\textrm{d}s)\, \textrm{d}r \\{} & {} \quad \quad \le \left( \sum _{i=1}^n \left( |\varphi _i|([-d,0]) \right) ^{p^*}\int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\{} & {} \qquad \quad \ u \sum _{i=1}^n |\varphi _i|([-d,0]) \mathbb {E} \left[ \sup _{(r+s) \in [0,u]}\left( e^{-\alpha p(r+s)}|z(r+s)-y(r+s)|^p\right) \right] . \end{aligned}$$
Therefore,
$$\begin{aligned} \int _0^T{} & {} e^{-\alpha pu}\mathbb {E}\left[ |Y(u)|^p\right] \, \textrm{d}u \\{} & {} \lesssim _{|\varphi _i|,p} \int _0^Tu \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\{} & {} \qquad \quad \mathbb {E} \left[ \sup _{(r+s) \in [0,u]}\left( e^{-\alpha p(r+s)}|z(r+s)-y(r+s)|^p\right) \right] \, \textrm{d}u \\{} & {} \lesssim _{|\varphi _i|,p} \int _0^Tu \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\{} & {} \qquad \quad \ \mathbb {E} \left[ \sup _{(r+s) \in [0,T]}\left( e^{-\alpha p(r+s)}|z(r+s)-y(r+s)|^p\right) \right] \, \textrm{d}u \\{} & {} =\left( \int _0^Tu \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \, \textrm{d}u \right) \ \Vert z-y\Vert _{\alpha }^p \\{} & {} \lesssim _{|\varphi _i|,T,p} \left( \int _0^Tr^{-2p^*\eta }e^{-2\alpha p^*r} \, \textrm{d}r\right) ^{\frac{p}{2p^*}} \Vert z-y\Vert _{\alpha }^p \end{aligned}$$
where the last inequality is obtained as follows:
$$\begin{aligned} \int _0^T&u \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r-s)}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \, \textrm{d}u \\&\le T \int _0^T \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r)} e^{2\alpha p^*s}\, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \, \textrm{d}u \\&\le T \int _0^T \left( \sum _{i=1}^n \int _0^u\int _{-d \vee -r}^0(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r)} \, \varphi _i(\textrm{d}s)\, \textrm{d}r\right) ^{\frac{p}{2p^*}} \, \textrm{d}u \\&\le T \sup _{u \in [0,T]} \left( \sum _{i=1}^n |\varphi _i|([-d,0])\int _0^u(u-r)^{-2p^*\eta }e^{-2\alpha p^*(u-r)} \, \textrm{d}r\right) ^{\frac{p}{2p^*}} \\&\lesssim _{|\varphi _i|, T}\left( \int _0^Tr^{-2p^*\eta }e^{-2\alpha p^*r} \, \textrm{d}r\right) ^{\frac{p}{2p^*}}. \end{aligned}$$
Putting together the above estimates we obtain
$$\begin{aligned} \mathbb {E}&\left[ \sup _{t \in [0,T]}e^{-\alpha p t}\left| \int _0^t \mathcal {G}({\bar{z}}^{x_1}-{\bar{y}}^{x_1}) \cdot \textrm{d}Z(r) \right| ^p \right] \nonumber \\&\lesssim _{T, |\varphi _i|,\eta ,p} \left( \int _0^Te^{-\alpha p'u}u^{p'(\eta -1)}\, \textrm{d}u\right) ^{\frac{p}{p'}} \left( \int _0^Tr^{-2p^*\eta }e^{-2\alpha p^*r} \, \textrm{d}r\right) ^{\frac{p}{2p^*}} \Vert z-y\Vert _{\alpha }^p \nonumber \\&\lesssim _{T, |\varphi _i|, \eta ,p} C_4(\alpha ) \Vert z-y\Vert _{\alpha }^p, \end{aligned}$$
(56)
Finally, from (53), (54), (55) and (56) we infer
$$\begin{aligned} \Vert F(z)-F(y)\Vert ^p_{\alpha } \lesssim _{\mu _y,T,p,|\phi |,|\varphi _i|, \Vert \sigma _y\Vert , \delta , \eta } \sum _{i=1}^4 C_i(\alpha ) \Vert z-y\Vert _{\alpha }^p, \end{aligned}$$
where \(C_i(\alpha ) \rightarrow 0\) as \(\alpha \rightarrow \infty \), for \(i=1,\ldots ,4\). So, by taking \(\alpha >0\) sufficiently large, this proves that F is a contraction, thus there exists a unique fixed point of it. In this way we prove the existence and uniqueness of the solution in the space \(L^p(\Omega ,S_T)\) for \(p >4\). Since, for such p, \(L^p(\Omega ,S_T)\subset L^2(\Omega ,S_T)\), such solution also belongs to \(L^2(\Omega ,S_T)\). To get uniqueness in the space \(L^2(\Omega ,C([0,T];\mathbb {R}))\) one can take two solutions y and \({{\tilde{y}}}\) in this space and take their difference. Using the fact that both are fixed points of F, by means of the Gronwall Lemma one gets \(\sup _{t\in [0,T]} \mathbb {E}\left[ |y(t)-{{\tilde{y}}}(t)|^2\right] =0\) and this concludes the proof.\(\square \)

A.2 Proof of Lemma 4.1

Proof
Let us denote with \(\sigma _y^i\) the i-th component of \(\sigma _y\), and let us show that
$$\begin{aligned} {\mathbb {E}} \left[ \int _{t_0}^t \left| y(s)\sigma _y^i + \int _{-d}^0 y(s+\tau ) \varphi _i(\text {d} \tau )\right| ^2 \text {d}s \right] < +\infty . \end{aligned}$$
By the trivial inequality \((a+b)^2 \le 2( a^2 + b^2)\), it is sufficient to show that
$$\begin{aligned} {\mathbb {E}} \left[ \int _{t_0}^t |y(s)\sigma _y^i|^2 \text {d}s \right] < +\infty , \end{aligned}$$
(57)
and
$$\begin{aligned} {\mathbb {E}} \left[ \int _{t_0}^t \left| \int _{-d}^0 y(s+\tau ) \varphi _i(\text {d}\tau ) \right| ^2 \text {d}s \right] < +\infty . \end{aligned}$$
(58)
We immediately see that (57) holds true thanks to Proposition 2.1.
To show (58), we use the Hölder inequality and the Fubini Theorem to estimate
$$\begin{aligned} \mathbb {E} \left[ \int _{t_0}^t\left| \int _{-d}^0 y(\tau +s)\, \varphi _i(\textrm{d}\tau )\right| ^2\,\textrm{d}s \right]&\le |\varphi _i| ([-d,0]) \mathbb {E} \left[ \int _{t_0}^t \int _{-d}^0 |y(\tau +s)|^2\, |\varphi _i|(\textrm{d}\tau )\,\textrm{d}s \right] \\&=|\varphi _i| ([-d,0]) \mathbb {E} \left[ \int _{-d}^0 \int _{t_0}^t |y(\tau +s)|^2\, \textrm{d}s\, |\varphi _i|(\textrm{d}\tau )\right] \\&=|\varphi _i| ([-d,0]) \mathbb {E} \left[ \int _{-d}^0 \int _{t_0+\tau }^{t+\tau } |y(r)|^2\, \textrm{d}r\, |\varphi _i|(\textrm{d}\tau )\right] \\&\le |\varphi _i| ([-d,0]) \mathbb {E} \left[ \int _{-d}^0 \int _{t_0-d}^{t} |y(r)|^2\, \textrm{d}r\, |\varphi _i|(\textrm{d}\tau )\right] \\&=\left( |\varphi _i| ([-d,0])\right) ^2 \mathbb {E} \left[ \int _{t_0-d}^{t} |y(r)|^2\, \textrm{d}r)\right] , \end{aligned}$$
which is finite, thanks to Proposition 2.1. \(\square \)

A.3 Proof of Proposition 4.3

Proof
(i)
The operator A can be written in the form
$$\begin{aligned} A\left( x_0,x_1\right) =\left( \int _{-d}^0 x_1(\theta )a(\text {d}\theta ), \frac{\text {d}}{\text {d}s}x_1\right) \,, \end{aligned}$$
(59)
where
$$\begin{aligned}a(\text {d}\theta )=\mu _y\delta _0(\text {d}\theta )+\Phi (\text {d}\theta )\,,\end{aligned}$$
and \(\delta _0\) is the delta-Dirac measure at zero. The measure a defines a finite measure on \(\left[ -d,0\right] \). The result is thus an immediate consequence of [14, PropositionA.27].
 
(ii)
See e.g. [23, Chapter 7, Lemma 1.2].
 
(iii)
Existence and uniqueness of a weak solution given by (28) for deterministic \({\textbf{m}}\) is a classical result (see [14, Proposition A.5]). One can then easily generalize the result to random \({\textbf{m}}\). Property (29) follows from uniqueness of the solution.
 
(iv)
If \(m(t_0;\cdot )\) is the unique solution to (25) then the \(\mathcal {H}\)-valued process \(\left( m(t_0;t),m(t_0;t+\cdot )\right) _{t\ge t_0}\) solves (27) by [4, Part II,Chapter 4, Theorem 4.3]. Since also the latter has a unique solution, its first component must be the solution to (27).
 
\(\square \)

A.4 Proof of Lemma 4.6

Proof
If \(\lambda \in \mathbb {R} \cap \rho (A)\) then \(K(\lambda ) \ne 0\) by Lemma 4.5. To compute \(R(\lambda , A)\), we will consider for a fixed \({\textbf{m}}=\left( m_0,m_1\right) \in \mathcal {H}\) the equation
$$\begin{aligned} (\lambda -A)\left( u_0,u_1\right) =\left( m_0,m_1\right) , \end{aligned}$$
(60)
in the unknown \((u_0,u_1)\in \mathcal {D}(A)\), that by definition of A is equivalent to
$$\begin{aligned} \left\{ \begin{aligned} (\lambda - (\mu _y - \sigma _y \cdot \kappa ) ) u_0- \int _{-d}^0u_1(\tau )\Phi (\text {d}\tau )&=m_0\\ \lambda u_1-\frac{\text {d}u_1}{\text {d}s}&=m_1. \end{aligned}\right. \end{aligned}$$
Then
$$\begin{aligned} u_1(s)=e^{\lambda s} u_0+\int _s^0e^{-\lambda (\tau -s)}m_1(\tau )\, \text {d}\tau ,\quad s\in [-d,0], \end{aligned}$$
and \(u_0\) is determined by the equation
$$\begin{aligned} \big (\lambda - (\mu _y - \sigma _y \cdot \kappa )\big )u_0=\left[ m_0+\int _{-d}^0\left( e^{\lambda \tau } u_0 +\int _{\tau }^0e^{-\lambda (s-\tau )}m_1(s)\, \text {d}s \right) \Phi (\text {d}\tau )\right] , \end{aligned}$$
yielding
$$\begin{aligned} K(\lambda )u_0 =m_0+ \int _{-d}^0\int _{-d}^{s}e^{-\lambda (s-\tau )}\Phi (\text {d}\tau ) \, m_1(s) \text {d} s. \end{aligned}$$
Then the result immediately follows. \(\square \)

A.5 Proof of Lemma 4.10

Proof
It is immediate to check that the function \({\widetilde{K}}:{\mathbb {R}}\rightarrow {\mathbb {R}}\) is continuous and differentiable with
$$\begin{aligned} {\widetilde{K}}^\prime (\xi )=1+\int _{-d}^0e^{\xi \tau }|\tau |\,|\Phi |(\text {d}\tau )>0, \end{aligned}$$
and that
$$\begin{aligned} \lim _{\xi \rightarrow \pm \infty }{\widetilde{K}}(\xi )=\pm \infty \,. \end{aligned}$$
Equation \({\widetilde{K}}(\xi )=0\) has thus exactly one real solution \(\bar{\xi }\). Let us now show that \(\bar{\xi }=\widetilde{\lambda _0}\). By the definition of \(\widetilde{\lambda _0}\), clearly we have that \(\bar{\xi }\le \widetilde{\lambda _0}\). To show the opposite inequality, \(\bar{\xi }\ge \widetilde{\lambda _0}\), we consider an arbitrary \(\lambda =a+i b\in \mathbb {C}\) such that \({\widetilde{K}}(\lambda )=0\). Then
$$\begin{aligned} \begin{aligned} 0&=\text {Re}(\widetilde{K}(\lambda ))=a-\mu _y+\sigma _y\cdot \kappa -\int _{-d}^0e^{a\tau }\cos (b\tau )\,|\Phi |(\text {d}\tau )\\&\ge a-\mu _y+\sigma _y \cdot \kappa -\int _{-d}^0e^{a\tau }\,|\Phi |(\text {d}\tau ) =:{\widetilde{K}}(a)\,. \end{aligned} \end{aligned}$$
Since \(\widetilde{K}\) is an increasing function, we can infer Re\(\lambda \le \bar{\xi }\) and taking the supremum in the definition of \(\widetilde{\lambda _0}\) we obtain \(\widetilde{\lambda _0} \le \bar{\xi }\). By the same argument, the relation \({\widetilde{K}}(r)> 0 \iff r > \widetilde{\lambda _0}\) immediately follows. \(\square \)

A.6 Proof of Lemma 4.11

Proof
Since by Lemma 4.10 we know that \(\widetilde{K}\) is an increasing function and \(\widetilde{K}(\widetilde{\lambda _0})=0\), we just need to prove that \(\widetilde{K}(\lambda _0) \le 0\). For every \(\lambda =a+i b \in \mathbb {C}\) we have
$$\begin{aligned} \text {Re}(K(\lambda ))=a-(\mu _y+\sigma _y\cdot \kappa )-\int _{-d}^0 e^{a\tau }\cos (b\tau )\,\Phi ( \textrm{d}\tau ). \end{aligned}$$
Recalling the definition of \(\lambda _0\) it is enough to show that, for every \(\lambda =a+i b \in \mathbb {C}\) such that \(K(\lambda )=0\), it holds \(\widetilde{K}(a)\le 0\). We have that
$$\begin{aligned} {{\tilde{K}}} (a)&=a-(\mu _y+\sigma _y\cdot \kappa )-\int _{-d}^0 e^{a\tau }|\Phi |(\textrm{d}\tau ) \\&=a-(\mu _y+\sigma _y\cdot \kappa )-\int _{-d}^0 e^{a\tau }\cos (b\tau )\,\Phi (\textrm{d}\tau ) -\int _{-d}^0 e^{a\tau }|\Phi |(\textrm{d}\tau ) \\&\quad + \int _{-d}^0 e^{a\tau }\cos (b\tau )\,\Phi (\textrm{d}\tau ) \\&\le \text {Re}(K(\lambda ))+\int _{-d}^0e^{a\tau } \, [\Phi -|\Phi |](\textrm{d}\tau ) \le 0. \end{aligned}$$
This concludes the proof. \(\square \)
Fußnoten
1
The importance of the past in understanding the qualitative feature of a model with delay was also emphasized in Fabbri and Gozzi [21], although in a deterministic setting, when solving the endogenous growth model with vintage capital of Boucekkine et al. [10].
 
2
Recall that the state price density \(\xi \) characterizes the Radon-Nikodym derivative that defines the change of probability measure from the objective probability measure \(\mathbb {P}\) to the risk-neutral measure \(\tilde{\mathbb {P}}\) via the relationship \(\xi (s)=e^{-rs}\rho (s)= e^{-rs}\frac{\textrm{d}\tilde{\mathbb {P}}}{\textrm{d}\mathbb {P}}(s)\).
 
3
Recall (see (3)) that our aim is to evaluate the expectation \({\mathbb {E}} \left[ \int _{t_0}^{+\infty } \xi (s) y(s) \text {d}s \mid {\mathcal {F}}_{t_0} \right] \). We will prove that \(\int _{t_0}^{+\infty } {\mathbb {E}} \left[ \xi (s) y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s\) is equal to the r.h.s. of (8) and then justify the equality \({\mathbb {E}} \left[ \int _{t_0}^{+\infty } \xi (s) y(s) \text {d}s \mid {\mathcal {F}}_{t_0} \right] =\int _{t_0}^{+\infty } {\mathbb {E}} \left[ \xi (s) y(s)\mid {\mathcal {F}}_{t_0} \right] \text {d}s\).
 
4
For more details on the estimates, the interested reader can consult that paper.
 
5
Notice that this condition require to work with \(p>2\).
 
6
This condition is made in order to guarantee the convergence of the integrals that will appear in what follows. Notice that this condition require to work with \(p>4\).
 
Literatur
1.
Zurück zum Zitat Abowd, J.M., Card, D.: On the covariance structure of earnings and hours changes. Econometrica 57(2), 411–445 (1989)CrossRef Abowd, J.M., Card, D.: On the covariance structure of earnings and hours changes. Econometrica 57(2), 411–445 (1989)CrossRef
2.
Zurück zum Zitat Arriojas, M., Hu, Y., Mohammed, S.A., Pap, G.: A delayed Black and Scholes formula. Stoch. Anal. Appl. 25(2), 471–492 (2007)MathSciNetCrossRefMATH Arriojas, M., Hu, Y., Mohammed, S.A., Pap, G.: A delayed Black and Scholes formula. Stoch. Anal. Appl. 25(2), 471–492 (2007)MathSciNetCrossRefMATH
3.
Zurück zum Zitat Barattieri, A., Basu, S., Gottschalk P.: Some evidence on the importance of sticky wages, NBER Working Paper No. 16130 (2010) Barattieri, A., Basu, S., Gottschalk P.: Some evidence on the importance of sticky wages, NBER Working Paper No. 16130 (2010)
4.
Zurück zum Zitat Bensoussan, A., Da Prato, G., Delfour, M.C., Mitter, S.K.: Representation and control of infinite dimensional systems, Second Edition, Birkhauser (2007) Bensoussan, A., Da Prato, G., Delfour, M.C., Mitter, S.K.: Representation and control of infinite dimensional systems, Second Edition, Birkhauser (2007)
5.
Zurück zum Zitat Biagini, S., Biffis, E., Gozzi, F., Zanella, M.: Wage rigidity and retirement in optimal portfolio choice, Working Paper: arXiv:2101.09732 (2022) Biagini, S., Biffis, E., Gozzi, F., Zanella, M.: Wage rigidity and retirement in optimal portfolio choice, Working Paper: arXiv:​2101.​09732 (2022)
6.
Zurück zum Zitat Biagini, S., Gozzi, F., Zanella, M.: Robust portfolio choice with sticky wages. SIAM J. Financial Math. 13(3), 1004–1039 (2022)MathSciNetCrossRefMATH Biagini, S., Gozzi, F., Zanella, M.: Robust portfolio choice with sticky wages. SIAM J. Financial Math. 13(3), 1004–1039 (2022)MathSciNetCrossRefMATH
7.
Zurück zum Zitat Biffis, E., Gozzi, F., Prosdocimi, C.: Optimal portfolio choice with path dependent labor income: the infinite horizon case. SIAM J. Control Optim. 58(4), 1906–1938 (2020)MathSciNetCrossRefMATH Biffis, E., Gozzi, F., Prosdocimi, C.: Optimal portfolio choice with path dependent labor income: the infinite horizon case. SIAM J. Control Optim. 58(4), 1906–1938 (2020)MathSciNetCrossRefMATH
8.
Zurück zum Zitat Biffis, E., Blake, D., Pitotti, L., Sun, A.: The cost of counterparty risk and collateralization in longevity swaps. J. Risk Insur. 83(2), 387–419 (2016)CrossRef Biffis, E., Blake, D., Pitotti, L., Sun, A.: The cost of counterparty risk and collateralization in longevity swaps. J. Risk Insur. 83(2), 387–419 (2016)CrossRef
9.
Zurück zum Zitat Bodie, Z., Merton, R.C., Samuelson, W.F.: Labor supply exibility and portfolio choice in a life cycle model. J. Econ. Dyn. Control 16(3), 427–449 (1992)CrossRef Bodie, Z., Merton, R.C., Samuelson, W.F.: Labor supply exibility and portfolio choice in a life cycle model. J. Econ. Dyn. Control 16(3), 427–449 (1992)CrossRef
10.
Zurück zum Zitat Boucekkine, R., Licandro, O., Puch, L., Del Rio, F.: Vintage capital and the dynamics of the AK model. J. Econ. Theory 120(1), 39–72 (2005)MathSciNetCrossRefMATH Boucekkine, R., Licandro, O., Puch, L., Del Rio, F.: Vintage capital and the dynamics of the AK model. J. Econ. Theory 120(1), 39–72 (2005)MathSciNetCrossRefMATH
11.
Zurück zum Zitat Brigo, D., Morini, M., Pallavicini, A.: Counterparty Credit Risk, Collateral and Funding: With Pricing Cases for all Asset Classes. Wiley, Collateral and Funding (2013)CrossRefMATH Brigo, D., Morini, M., Pallavicini, A.: Counterparty Credit Risk, Collateral and Funding: With Pricing Cases for all Asset Classes. Wiley, Collateral and Funding (2013)CrossRefMATH
12.
Zurück zum Zitat Brigo, D., Pallavicini, A.: Nonlinear consistent valuation of CCP cleared or CSA bilateral trades with initial margins under credit, funding and wrong-way risks. J. Financial Eng. 1(1), 1–60 (2014)MathSciNetCrossRef Brigo, D., Pallavicini, A.: Nonlinear consistent valuation of CCP cleared or CSA bilateral trades with initial margins under credit, funding and wrong-way risks. J. Financial Eng. 1(1), 1–60 (2014)MathSciNetCrossRef
13.
Zurück zum Zitat Campbell, J.Y., Viceira, L.M.: Strategic Asset Allocation: Portfolio Choice for Long-Term Investors. Oxford University Press, Oxford (2001) Campbell, J.Y., Viceira, L.M.: Strategic Asset Allocation: Portfolio Choice for Long-Term Investors. Oxford University Press, Oxford (2001)
14.
Zurück zum Zitat Da Prato, G., Zabczyk, J.: Stochastic Equations in Infinite Dimensions. Cambridge University Press, Cambridge (1992)CrossRefMATH Da Prato, G., Zabczyk, J.: Stochastic Equations in Infinite Dimensions. Cambridge University Press, Cambridge (1992)CrossRefMATH
15.
Zurück zum Zitat Dickens, W.T., Goette, L., Groshen, E.L., Holden, S., Messina, J., Schweitzer, M.E., Turunen, J., Ward, M.E.: How wages change: micro evidence from the international wage flexibility project. J. Econ. Perspect. 21(2), 195–214 (2007)CrossRef Dickens, W.T., Goette, L., Groshen, E.L., Holden, S., Messina, J., Schweitzer, M.E., Turunen, J., Ward, M.E.: How wages change: micro evidence from the international wage flexibility project. J. Econ. Perspect. 21(2), 195–214 (2007)CrossRef
16.
Zurück zum Zitat Djeiche, B., Gozzi, F., Zanco, G., Zanella, M.: Optimal portfolio choice with path dependent benchmarked labor income: a mean field model. Stoch. Process. Appl. 145, 48–85 (2022)MathSciNetCrossRefMATH Djeiche, B., Gozzi, F., Zanco, G., Zanella, M.: Optimal portfolio choice with path dependent benchmarked labor income: a mean field model. Stoch. Process. Appl. 145, 48–85 (2022)MathSciNetCrossRefMATH
17.
Zurück zum Zitat Delbaen, F., Schachermayer, W.: A general version of the fundamental theorem of asset pricing. Math. Ann. 300(1), 463–520 (1994)MathSciNetCrossRefMATH Delbaen, F., Schachermayer, W.: A general version of the fundamental theorem of asset pricing. Math. Ann. 300(1), 463–520 (1994)MathSciNetCrossRefMATH
18.
Zurück zum Zitat Duffie, D.: Dynamic Asset Pricing Theory. Princeton University Press, Princeton (2001)MATH Duffie, D.: Dynamic Asset Pricing Theory. Princeton University Press, Princeton (2001)MATH
19.
Zurück zum Zitat Dunsmuir, W., Goldys, B., Tran, C.V.: Stochastic delay differential equations as weak limits of autoregressive moving average time series, Working Paper, University of New South Wales and University of Sydney (2015) Dunsmuir, W., Goldys, B., Tran, C.V.: Stochastic delay differential equations as weak limits of autoregressive moving average time series, Working Paper, University of New South Wales and University of Sydney (2015)
20.
Zurück zum Zitat Dybvig, P., Liu, H.: Lifetime consumption and investment: retirement and constrained borrowing. J. Econ. Theory 145(3), 885–907 (2010)MathSciNetCrossRefMATH Dybvig, P., Liu, H.: Lifetime consumption and investment: retirement and constrained borrowing. J. Econ. Theory 145(3), 885–907 (2010)MathSciNetCrossRefMATH
21.
Zurück zum Zitat Fabbri, G., Gozzi, F.: Solving optimal growth models with vintage capital: the dynamic programming approach. J. Econ. Theory 143(1), 331–373 (2008)MathSciNetCrossRefMATH Fabbri, G., Gozzi, F.: Solving optimal growth models with vintage capital: the dynamic programming approach. J. Econ. Theory 143(1), 331–373 (2008)MathSciNetCrossRefMATH
22.
Zurück zum Zitat Guvenen, F.: Learning your earning: Are labor income shocks really very persistent? Am. Econ. Rev. 97(3), 687–712 (2007)CrossRef Guvenen, F.: Learning your earning: Are labor income shocks really very persistent? Am. Econ. Rev. 97(3), 687–712 (2007)CrossRef
23.
Zurück zum Zitat Hale, J.K., Verduyn Lunel, S.M.: Introduction to Functional Differential Equations. Springer, Berlin (1993)CrossRefMATH Hale, J.K., Verduyn Lunel, S.M.: Introduction to Functional Differential Equations. Springer, Berlin (1993)CrossRefMATH
24.
Zurück zum Zitat Harrison, J.M., Kreps, D.M.: Martingales and arbitrage in multiperiod securities markets. J. Econ. Theory 20(3), 381–408 (1979)MathSciNetCrossRefMATH Harrison, J.M., Kreps, D.M.: Martingales and arbitrage in multiperiod securities markets. J. Econ. Theory 20(3), 381–408 (1979)MathSciNetCrossRefMATH
25.
Zurück zum Zitat Harrison, J.M., Pliska, S.R.: Martingales and stochastic integrals in the theory of continuous trading. Stoch. Process. Appl. 11(3), 215–260 (1981)MathSciNetCrossRefMATH Harrison, J.M., Pliska, S.R.: Martingales and stochastic integrals in the theory of continuous trading. Stoch. Process. Appl. 11(3), 215–260 (1981)MathSciNetCrossRefMATH
26.
Zurück zum Zitat Hubbard, R.G., Skinner, J., Zeldes, S.P.: Precautionary saving and social insurance. J. Polit. Econ. 103(2), 360–399 (1995)CrossRef Hubbard, R.G., Skinner, J., Zeldes, S.P.: Precautionary saving and social insurance. J. Polit. Econ. 103(2), 360–399 (1995)CrossRef
27.
Zurück zum Zitat Karatzsas, I., Shreve, S.E.: Brownian Motion and Stochastic Calculus. Springer, Berlin (1991) Karatzsas, I., Shreve, S.E.: Brownian Motion and Stochastic Calculus. Springer, Berlin (1991)
28.
Zurück zum Zitat Khan, S.: Evidence of nominal wage stickiness from microdata. Am. Econ. Rev. 87(5), 993–1008 (1997) Khan, S.: Evidence of nominal wage stickiness from microdata. Am. Econ. Rev. 87(5), 993–1008 (1997)
29.
Zurück zum Zitat Lorenz, R.: Weak approximation of stochastic delay differential equations with bounded memory by discrete time series, Ph.D. Thesis, Humboldt University, Berlin (2005) Lorenz, R.: Weak approximation of stochastic delay differential equations with bounded memory by discrete time series, Ph.D. Thesis, Humboldt University, Berlin (2005)
30.
Zurück zum Zitat Le Bihan, H., Montornès, J., Heckel, T.: Sticky wages: evidence from quarterly microeconomic data. Am. Econ. J. Macroecon. 4(3), 1–32 (2012)CrossRef Le Bihan, H., Montornès, J., Heckel, T.: Sticky wages: evidence from quarterly microeconomic data. Am. Econ. J. Macroecon. 4(3), 1–32 (2012)CrossRef
31.
Zurück zum Zitat MaCurdy, T.E.: The use of time series processes to model the error structure of earnings in a longitudinal data analysis. J. Econ. 18(1), 83–114 (1982)CrossRef MaCurdy, T.E.: The use of time series processes to model the error structure of earnings in a longitudinal data analysis. J. Econ. 18(1), 83–114 (1982)CrossRef
32.
33.
Zurück zum Zitat Meghir, C., Pistaferri, L.: Income variance dynamics and heterogeneity. Econometrica 72(1), 1–32 (2004)CrossRefMATH Meghir, C., Pistaferri, L.: Income variance dynamics and heterogeneity. Econometrica 72(1), 1–32 (2004)CrossRefMATH
34.
Zurück zum Zitat Mohammed, S.A.: Stochastic Differential Systems with Memory: Theory. Southern Illinois University Carbondale, Examples and Applications (1998) Mohammed, S.A.: Stochastic Differential Systems with Memory: Theory. Southern Illinois University Carbondale, Examples and Applications (1998)
35.
Zurück zum Zitat Reiss, M.: Nonparametric estimation for stochastic delay differential equations, Ph.D. Thesis, Humboldt University, Berlin (2002) Reiss, M.: Nonparametric estimation for stochastic delay differential equations, Ph.D. Thesis, Humboldt University, Berlin (2002)
36.
Zurück zum Zitat Samuelson, P.A.: Rational theory of warrant pricing. Ind. Manag. Rev. 6, 13–32 (1965) Samuelson, P.A.: Rational theory of warrant pricing. Ind. Manag. Rev. 6, 13–32 (1965)
Metadaten
Titel
A pricing formula for delayed claims: appreciating the past to value the future
verfasst von
Enrico Biffis
Beniamin Goldys
Cecilia Prosdocimi
Margherita Zanella
Publikationsdatum
13.01.2023
Verlag
Springer Berlin Heidelberg
Erschienen in
Mathematics and Financial Economics / Ausgabe 2/2023
Print ISSN: 1862-9679
Elektronische ISSN: 1862-9660
DOI
https://doi.org/10.1007/s11579-022-00331-7

Weitere Artikel der Ausgabe 2/2023

Mathematics and Financial Economics 2/2023 Zur Ausgabe