1. Introduction
The theory of backward stochastic differential equations (BSDEs), initiated by Bismut [Reference Bismut7] and Pardoux and Peng [Reference Pardoux and Peng32], has been extensively studied over the past three decades, particularly in relation to stochastic control, finance, and insurance (see e.g. [Reference Delong17, Reference Zhang39]). Important applications include dynamic risk measures [Reference Barrieu and El Karoui4] and g-expectations [Reference Coquet, Hu, Memin and Peng16, Reference Peng33], which generalize classical expectations and martingales to nonlinear settings. Recent applications to financial economics include [Reference Beissner, Lin and Riedel6], [Reference Bouchard, Fukasawa, Herdegen and Muhle-Karbe8], [Reference Gonon, Muhle-Karbe and Shi20], [Reference Herdegen, Muhle-Karbe and Possamaï24], [Reference Kardaras, Xing and Žitkovi28], and [Reference Muhle-Karbe, Nutz and Tan30].
While BSDEs are powerful theoretical tools, their solutions are typically implicit and require discretization for numerical implementation. As discrete analogs, backward stochastic difference equations (BS
$\Delta$
Es) have been widely studied, falling into two main categories. The first focuses on BS
$\Delta$
Es as weak approximations of BSDEs [Reference Briand, Delyon and Mémin9, Reference Briand, Delyon and Mémin10, Reference Briand, Geiss, Geiss and Labart11, Reference Cheridito and Stadje12, Reference Ma, Protter, Martín and Torres29, Reference Nakayama31, Reference Stadje35, Reference Tanaka37]. The second explores the structure of BS
$\Delta$
Es themselves. A general framework is provided in [Reference Cohen and Elliot15], while specific cases involving driving martingales with the predictable representation property are studied in [Reference Cohen and Elliott14] and [Reference Elliott, Siu and Cohen19].
This paper falls into the second category and studies a class of BS
$\Delta$
Es including the one introduced in [Reference Nakayama31] and [Reference Tanaka37], where a d-dimensional scaled random walk – whose increments take only
$d+1$
values – is used to approximate Brownian motion in BSDEs. Such a random walk is minimal among discrete-time processes that converge to d-dimensional Brownian motions. The weak convergence of this BS
$\Delta$
E to the BSDE was proved in [Reference Nakayama31], and the convergence rate in a Markovian setting was given in [Reference Tanaka37], generalizing the one-dimensional case in [Reference Briand, Geiss, Geiss and Labart11].
This minimal BS
$\Delta$
E is computationally efficient, as it involves only a
$(d + 1)$
-dimensional problem, in contrast to a
$2^d$
-dimensional problem required when using d-dimensional Bernoulli random walks [Reference Cheridito and Stadje12]. Although Cohen and Elliott [Reference Cohen and Elliot15] have developed a general BS
$\Delta$
E theory, we focus on the specific structure of this minimal BS
$\Delta$
E, which exhibits properties not covered in their general framework. In the one-dimensional case, it reduces to a BS
$\Delta$
E on a binomial tree, studied in [Reference Elliott, Siu and Cohen19] in the context of dynamic risk measures. We treat the general multi-dimensional setting here.
Our key contribution is to identify a gradient constraint on the BS
$\Delta$
E driver, which endows the solution with certain properties as a generalized conditional expectation. This allows us to link the driver to a measure change for the driving random walk, and apply these insights to market equilibrium analysis.
The g-expectation is part of the solution of a BSDE or BS
$\Delta$
E, and generalizes the expectation and the certainty equivalent of an expected utility. A subclass of them with concavity and translation invariance has been employed as the utility functional for market equilibrium analyses in [Reference Antar and Dempster3], [Reference Cheridito, Horst, Kupper and Pirvu13], [Reference Horst, Pirvu and Dos Reis25], and [Reference Kardaras, Xing and Žitkovi28]. In this paper, we also apply our BS
$\Delta$
E to a market equilibrium analysis. In contrast to the preceding studies, which place an emphasis on incomplete markets, we are interested in explicit computations in a dynamically complete market.
Anderson and Raimondo [Reference Anderson and Raimondo2] proved the existence of equilibrium in a continuous-time dynamically complete market by means of non-standard analysis, where an approximation to a Brownian motion by a minimal random walk played a key role. We consider a simpler dynamically complete market to derive explicit conditions for market equilibrium.
Under a unique equivalent martingale measure, our asset price model is a multi-dimensional extension of the recombining binomial tree. In our approach, an asset price process is given as a stochastic process taking values on a lattice. We do not argue the existence of an equilibrium price but characterize the agents’ utilities under which the given discrete (in both time and space) price process is to be in general equilibrium. This feature is in contrast to the preceding studies [Reference Antar and Dempster3], [Reference Cheridito, Horst, Kupper and Pirvu13], and [Reference Horst, Pirvu and Dos Reis25] and similar to [Reference Bouchard, Fukasawa, Herdegen and Muhle-Karbe8], [Reference He and Leland23], [Reference Kardaras, Xing and Žitkovi28], and [Reference Sekine34] in continuous time.
Our framework includes heterogeneous agents with exponential utilities under heterogeneous beliefs. Their risk-aversion coefficients may be stochastic and time-varying. We observe in particular that under equilibrium with heterogeneous beliefs, agents trade with each other, even in the absence of random endowments to hedge, complementing earlier studies of heterogeneous beliefs [Reference Basak5, Reference Hara21, Reference Harris and Raviv22, Reference Jouini and Napp26, Reference Muhle-Karbe, Nutz and Tan30, Reference Varian38].
In Section 2.1 we describe a lattice in
$\mathbb{R}^d$
where a stochastic process
$\{X_n\}$
takes values, and give some elementary linear algebraic lemmas as a preliminary. In Section 2.2 we introduce the process
$\{X_n\}$
that is the source of randomness in this paper and generates a filtration. It is minimal in the sense that the increment
$\Delta X_n$
takes values in a set
$\{v_0,\ldots,v_d\}$
of
$d+1$
points in
$\mathbb{R}^d$
. Some elementary measure change formulas are also given as a preliminary.
In Section 2.3 our BS
$\Delta$
E
is formulated. Due to the minimality of
$\{X_n\}$
, there exists a unique solution
$\{(Y_n,Z_n)\}$
to the above equation, without orthogonal martingale terms needed in [Reference Briand, Delyon and Mémin10] and [Reference Cheridito and Stadje12]. The process
$\{X_n\}$
itself takes more than
$d+1$
points, so this BS
$\Delta$
E is different from the one studied in [Reference Cohen and Elliott14]. The g-expectation
$\mathcal{E}^g_n$
for
$g = \{g_n\}$
is defined by
$\mathcal{E}^g_n(Y) = Y_n$
. Proposition 2.1 concerns the case
$g_n(z) = f_n(X_{n-1},z)$
and
$Y_N = h(X_N)$
for deterministic functions
$f_n$
and h to provide a nonlinear Feynman–Kac-type formula, which is a computationally efficient recurrence equation on the lattice for a deterministic function
$u_n$
such that
$Y_n = u_n(X_n)$
.
Section 2.4 is about the aforementioned gradient constraint. First we observe that the g-expectation is a conditional expectation when
$g_n$
are linear with slope coefficients included in the convex hull
$\Theta$
of the set
$\{v_0,\ldots,v_d\}$
. The importance of this constraint on the slope is a special feature of our BS
$\Delta$
E, and to the best of our knowledge has not been recognized in the preceding studies of multi-dimensional BS
$\Delta$
Es. A balance condition introduced by Cohen and Elliot [Reference Cohen and Elliot15] for a comparison theorem to hold is translated in terms of
$\Theta$
for our BS
$\Delta$
E. We also prove a robust representation when
$g_n$
are concave, where the set
$\Theta$
again plays an important role. In Section 2.5 we show that a translation-invariant filtration-consistent nonlinear expectation is a g-expectation.
In Section 3 we regard
$\{X_n\}$
as a d-dimensional asset price process. In Section 3.1 we consider an optimal investment strategy which maximizes the g-expectation of terminal wealth. By the minimality, the market is complete, extending the well-known binomial tree model for a one-dimensional asset. Our asset price model can be seen as a discrete approximation of the multi-dimensional Bachelier model with constant covariance and general stochastic drift. An advantage of our use of the minimal process as an approximation is that the completeness of the Bachelier model is preserved. Further, the minimality property naturally arises in a variance swap pricing model as illustrated in Example 3.2. In Sections 3.2–3.4 we give a market equilibrium analysis. We consider agents whose utility functionals are g expectations and seek conditions on those g expectations under which
$\{X_n\}$
is an equilibrium price process.
Throughout our financial application, we have short maturity problems in mind, and so, for brevity, assume interest rates, dividend rates, and consumption rates to be zero as in [Reference Antar and Dempster3], [Reference Cheridito, Horst, Kupper and Pirvu13], [Reference Horst, Pirvu and Dos Reis25], and [Reference Kardaras, Xing and Žitkovi28].
We use the convention that
for any sequence
$\{a_i\}$
if
$m \gt n$
.
2. BS
$\Delta$
E on a lattice
2.1. Lattice
We start by describing a lattice. Let
$\{v_1,\ldots, v_d\}$
be a basis of
$\mathbb{R}^d$
. The subset
\begin{equation*} L = \Biggl\{\sum_{i=1}^d z_i v_i, \ ; \ z_i \in \mathbb{Z}, i=1,\ldots, d\Biggr\}\end{equation*}
of
$\mathbb{R}^d$
is a d-dimensional lattice generated by the basis. Notice that L admits an alternative expression
\begin{equation} L = \Biggl\{\sum_{i=0}^d n_i v_i, \ ; \ n_i \in \mathbb{N}, i=0,1,\ldots, d\Biggr\},\end{equation}
where
$v_0 = -v_1 - \cdots - v_d$
and
$\mathbb{N}$
is the set of the non-negative integers. Let
be the
$d \times (d+1)$
matrix with
$v_i$
,
$i=0,\ldots, d$
as its column vectors. Put
The following lemmas will be of repeated use in this paper.
Lemma 2.1. The
$(d+1)\times (d+1)$
matrix
$(\mathbf{1},\mathbf{v}^\top)$
is invertible.
Proof. We show that the row vectors of
$(\mathbf{1},\mathbf{v}^\top)$
are linearly independent. Suppose
or equivalently
\begin{equation*} \sum_{j=0}^d \alpha_j = 0,\quad \sum_{j=0}^d \alpha_j v_j = 0\end{equation*}
for scalars
$\alpha_j$
. By the second equation and the definition of
$v_0$
, we have
\begin{equation*} \sum_{j=1}^d (\alpha_j -\alpha_0) v_j = 0,\end{equation*}
from which we can conclude
$\alpha_j = \alpha_0 $
for all j because
$\{v_1,\ldots,v_d\}$
is a basis. Together with the first equation, we then conclude
$\alpha_j = 0$
for all j.
Lemma 2.2. Let
$y \in \mathbb{R}^{d+1}$
. The unique solution to the equation
is given by
Proof. By Lemma 2.1, there exists a unique
$z \in \mathbb{R}^{d+1}$
such that (2.2) holds. Since
$\mathbf{v}\mathbf{1} = 0$
, multiplying both sides of (2.2) by
$\mathbf{1}^\top$
, we obtain the first equation of (2.3). Also, multiplying both sides of (2.2) by
$\mathbf{v}$
and again using
$\mathbf{v}\mathbf{1} = 0$
, we have
$\mathbf{v}y = \mathbf{v}\mathbf{v}^\top (z_1,\ldots,z_d)^\top$
. Since
$\{v_1,\ldots,v_d\}$
is a basis,
$\mathbf{v}$
has rank d. Therefore
$\mathbf{v}^\top$
has rank d and so, for any
$x \in \mathbb{R}^d\setminus\{0\}$
,
$x^\top \mathbf{v}\mathbf{v}^\top x = | \mathbf{v}^\top x|^2 \neq 0$
. This implies that the
$d\times d$
matrix
$\mathbf{v}\mathbf{v}^\top$
is invertible and, in turn, the second equation of (2.3) is valid.
2.2. Probability space
Let
$(\Omega,\mathscr{F},\mathbb{P})$
be a probability space. For a stochastic process
$\{X_n\}_{n \in \mathbb{N}}$
, we put
$\Delta X_n = X_n - X_{n-1}$
. Let
$\{X_n\}$
be a d-dimensional stochastic process with
$\Delta X_n$
taking values in
$\{v_0,v_1,\ldots,v_d\}$
for all
$n \geq 1$
and
$X_0 = 0$
. By (2.1),
$X_n$
takes values in L for all
$n \in \mathbb{N}$
. Let
$\mathscr{F}_n = \sigma(X_0,\ldots,X_n)$
,
$ n\in \mathbb{N}$
, be the natural filtration generated by
$\{X_n\}$
, and put
$P_n = (P_{n,0},\ldots, P_{n,d})^\top$
for
$n\geq 1$
, where
Note that
$\{P_n\}$
is a
$\Delta_d$
-valued predictable process, where
We assume
$P_{n,j}$
is positive for all
$n \geq 1$
and
$j = 0,\ldots, d$
. Let
$N \in \mathbb{N}\setminus\{0\}$
. The following lemma will be of repeated use in this paper.
Lemma 2.3. For any
$\Delta_d$
-valued predictable process
$\{\hat{P}_n\}$
, there exists a probability measure
$\hat{\mathbb{P}}$
on
$(\Omega,\mathscr{F}_N)$
such that
$\hat{P}_n = (\hat{P}_{n,0},\ldots, \hat{P}_{n,d})^\top$
, where
Proof. Define
$\hat{\mathbb{P}}$
by
\begin{equation} \hat{\mathbb{P}}(A) = \mathbb{E}[L_N1_A],\quad L_n = \prod_{k=1}^n \Biggl(\sum_{j=0}^d\frac{\hat{P}_{k,j}}{P_{k,j}}1_{\{\Delta X_k = v_j \}} \Biggr) \end{equation}
for
$A \in \mathscr{F}_N$
. The measure
$\hat{\mathbb{P}}$
is a probability measure because
\begin{equation*} \mathbb{E}\Biggl[ \sum_{j=0}^d\frac{\hat{P}_{n,j}}{P_{n,j}}1_{\{\Delta X_n = v_j \}} \bigg| \mathscr{F}_{n-1}\Biggr] = \sum_{j=0}^d\frac{\hat{P}_{n,j}}{P_{n,j}} P_{n,j} = 1\end{equation*}
and so
$L_n$
is a martingale with
$L_0 = 1$
. Using Bayes’ formula, we derive
by the martingale property of
$L_n$
.
Define a measure
$\mathbb{Q}$
on
$\mathscr{F}_N$
by
\begin{equation*} \mathbb{Q}(A) = \mathbb{E}[L_N1_A],\quad L_N = \prod_{n=1}^N \Biggl( \frac{1}{d+1} \sum_{j=0}^d\frac{1}{P_{n,j}}1_{\{\Delta X_n = v_j \}} \Biggr). \end{equation*}
Let
$\mathbb{E}_\mathbb{Q}$
denote the integration under
$\mathbb{Q}$
.
Lemma 2.4. The measure
$\mathbb{Q}$
is the unique probability measure on
$\mathscr{F}_N$
under which
$\{X_n\}$
is a martingale. Under
$\mathbb{Q}$
,
$\{\Delta X_n\}$
is i.i.d. with
for all
$n =1,\ldots, N$
and
$j=0,\ldots, d$
. We also have
Proof. By Lemma 2.3,
$\mathbb{Q}$
is a probability measure with
$ \mathbb{Q}(\Delta X_n =v_j \mid \mathscr{F}_{n-1}) = 1/(d+1)$
, which implies
Therefore
$\{X_n\}$
is a martingale with (2.6). There is no other such measure because
\begin{equation*} \sum_{j=0}^d \alpha_j = 1, \quad \sum_{j=0}^d \alpha_j v_j = 0\end{equation*}
implies
$\alpha_j = 1/(d+1)$
as in the proof of Lemma 2.1. Since the conditional law of
$\Delta X_n$
given
$\mathscr{F}_{n-1}$
is deterministic for every n,
$\{\Delta X_n\}$
is i.i.d.
Remark 2.1. For any positive definite
$d\times d$
matrix
$\Sigma$
, we can construct such a lattice L that
$\mathbf{v}\mathbf{v}^\top = \Sigma$
. Indeed, starting with an arbitrary basis, say,
$\bar{v}_j = e_j$
(the standard basis of
$\mathbb{R}^d)$
with
$\bar{v}_0 = -\bar{v}_1 - \cdots - \bar{v}_d$
and
$\bar{\mathbf{v}} = [\bar{v}_0, \ldots, \bar{v}_d]$
, using the Cholesky decomposition
$\Sigma = CC^\top$
and
$\bar{\mathbf{v}}\bar{\mathbf{v}}^\top= \bar{C}\bar{C}^\top$
, if we take
$v_j = C\bar{C}^{-1}\bar{v}_j$
,
$j=0,\ldots, d$
, then
$\mathbf{v} = C\bar{C}^{-1}\bar{\mathbf{v}}$
and so we get
$\mathbf{vv}^\top = C\bar{C}^{-1}\bar{\mathbf{v}}\bar{\mathbf{v}}^\top (\bar{C}^\top)^{-1}C^\top = \Sigma$
. In particular, we can construct such
$v_j$
that
$\mathbf{vv}^\top$
is the identity matrix. In this case a scaling limit of
$\{X_n\}$
under
$\mathbb{Q}$
is the d-dimensional standard Brownian motion. Such a set of vectors played an essential role in proving the existence of continuous-time market equilibrium in Anderson and Raimondo [Reference Anderson and Raimondo2] by means of non-standard analysis, where the existence of the vectors was proved in a recursive manner. It is also the building block of a d-dimensional diamond in topological crystallography [Reference Sunada36].
2.3. Existence, uniqueness and representation
Here we introduce our BS
$\Delta$
E. Let
$\mathcal{A}$
denote the set of the sequences
$g = \{g_n\}_{n=1}^N$
of
$\mathscr{F}_{n-1}\otimes \mathscr{B}(\mathbb{R}^d)$
measurable functions
$g_n\colon \Omega \times \mathbb{R}^d \to \mathbb{R}$
. Now we state an elementary but fundamental result.
Theorem 2.1. Let
$Y_N$
be an
$\mathscr{F}_N$
-measurable random variable, and let
$g = \{g_n\} \in \mathcal{A}$
. Then there exist uniquely an adapted process
$\{Y_n\}_{n=0,\ldots,N-1}$
and an
$\mathbb{R}^d$
-valued predictable process
$\{Z_n\}_{n=1,\ldots,N}$
such that
Further, they admit the following representation:
where
$\bar{Y}_n = (\bar{Y}_{n,0}, \ldots, \bar{Y}_{n,d})^\top$
and
Proof. Since
$Y_N$
is
$\mathscr{F}_N$
-measurable, there exists a function
$f\colon L^N \to \mathbb{R}$
such that
$Y_N = f(X_1,\ldots,X_N)$
. Since
$\mathbb{P}_{N,j}$
are positive by the assumption, (2.7) for
$n=N$
is equivalent to the system of equations for
$\mathscr{F}_{N-1}$
-measurable random variables
\begin{equation*}Y\,:\!=\,\begin{bmatrix}f(X_1,\ldots,X_{N-1},X_{N-1} + v_0) \\ \vdots\\f(X_1,\ldots,X_{N-1},X_{N-1} + v_d)\end{bmatrix}= (\mathbf{1},\mathbf{v}^\top)\begin{bmatrix} Y_{N-1} - g_N(Z_N) \\ Z_N\end{bmatrix}\!.\end{equation*}
Applying Lemma 2.2, we obtain (2.7) for
$n=N$
with
where a(y) and b(y) are defined by (2.3). It is clear that both
$Z_N$
and
$Y_{N-1}$
are
$\mathscr{F}_{N-1}$
-measurable. By backward induction, we obtain
$\{Y_n\}$
and
$\{Z_n\}$
. The representation follows from (2.7) and (2.6) by taking the conditional expectation under
$\mathbb{Q}$
.
For
$g = \{g_n\} \in \mathcal{A}$
fixed, the
$\mathscr{F}_n$
-measurable random variable
$Y_n$
given by Theorem 2.1 is uniquely determined by the
$\mathscr{F}_N$
-measurable random variable
$Y_N$
. We write this mapping as
$Y_n = \mathcal{E}^g_n(Y_N)$
and call it the g-expectation of
$Y_N$
(with respect to
$\mathscr{F}_n$
). The stochastic process
$\{(Y_n,Z_n)\}$
given by Theorem 2.1 is called the solution of the BS
$\Delta$
E (2.7).
Remark 2.2. In the literature, say, in [Reference Cohen and Elliot15], BS
$\Delta$
E is formulated by decomposing
$\Delta Y_n$
into a predictable part and a martingale difference part. In our formulation (2.7),
$\Delta X_n$
is not necessarily a martingale difference. It is a minor reparametrization because (2.7) can be rewritten as
with
$\hat{g}_n(z) = g_n(z) - z^\top A_n$
,
$A_n = \mathbb{E}[\Delta X_n \mid \mathscr{F}_{n-1}]$
.
Example 2.1. Let
$\gamma \gt 0$
,
$\{(\hat{P}_{n,0},\ldots,\hat{P}_{n,d})^\top\}$
be a
$\Delta_d$
-valued predictable process, and
\begin{equation} g_n(z) = -\frac{1}{\gamma} \log \Biggl(\sum_{j=0}^{d} \,{\mathrm{e}}^{-\gamma z^\top v_j} {\hat{P}}_{n,j}\Biggr).\end{equation}
Then
for any
$\mathscr{F}_N$
-measurable random variable Y, where
$\hat{\mathbb{E}}$
is the expectation under the measure
$\hat{\mathbb{P}}$
on
$\mathscr{F}_N$
defined by (2.5). To see this, note that by Lemma 2.3,
Substituting
$Y_n= Y_{n-1}- g_n (Z_n) + Z_n ^\top \Delta X_n $
, we have
which implies (2.9) for
$n=N-1$
. The general case follows by backward induction.
Next we give a discrete analog of the nonlinear Feynman–Kac formula, which is computationally efficient when dealing with large N.
Proposition 2.1. Let
$f_n \colon L \times \mathbb{R}^d \to \mathbb{R}$
,
$n=1,\ldots, N$
and
$h \colon L \to \mathbb{R}$
. Define
$u_n \colon L \to \mathbb{R}$
,
$n=0,1,\ldots, N$
backward inductively by
with
$u_N = h$
, where
\begin{align*} \mathcal{L} u_n(x) & = \frac{1}{d+1} \sum_{j=0}^d (u_n(x+v_j)-u_n(x)), \\* \mathcal{N} u_n(x) & = (u_n(x+v_0)-u_n(x), \ldots, u_n(x+v_d)-u_n(x))^\top.\end{align*}
Then the unique solution to (2.7) with
$g_n = f_n(X_{n-1},\cdot)$
and
$Y_N = h(X_N)$
is given by
Proof. By definition,
$Y_N = h(X_N) = u_N(X_N)$
. Suppose
$Y_n = u_n(X_n)$
. Then, by Theorem 2.1,
$Z_n = (\mathbf{v}\mathbf{v}^\top)^{-1}\mathbf{v}\bar{Y}_n$
, where
Using
$\mathbf{v1} = 0$
, we conclude
$Z_n = (\mathbf{vv}^\top)^{-1}\mathbf{v}\mathcal{N} u_n(X_{n-1})$
. Further, again by Theorem 2.1,
which concludes the proof.
2.4. A gradient constraint
Let
$\Theta$
be the closed convex hull spanned by
$\{v_0,v_1,\ldots,v_d\}$
, or equivalently
In this section we study BS
$\Delta$
Es with the gradient of g being constrained in
$\Theta$
.
Example 2.2. The triangular lattice of
$\mathbb{R}^2$
is generated by
In this case,
$\mathbf{v}\mathbf{v}^\top = I$
and
$\Theta$
is an equilateral triangle.
Proposition 2.2. Let
$g_n(z) = A_n^\top z + B_n$
for a
$\Theta$
-valued predictable process
$\{A_n\}$
and a predictable process
$\{B_n\}$
,
$n=1,\ldots,N$
. Then
\begin{equation*} \mathcal{E}^g_n (Y) = \hat{\mathbb{E}}\Biggl[Y +\sum_{i=n+1}^N B_i \bigg|\mathscr{F}_n\Biggr], \quad n=0,1,\ldots,N, \end{equation*}
for any
$\mathcal{F}_N$
-measurable random variable Y, where
$\hat{\mathbb{E}}$
is the expectation under the measure
$\hat{\mathbb{P}}$
on
$\mathscr{F}_N$
defined by (2.5) with
$\hat{P}_n = (\hat{P}_{n,0},\ldots, \hat{P}_{n,d})^\top$
such that
$A_n = \mathbf{v}\hat{P}_n$
.
Proof. By Lemma 2.3,
\begin{equation*} \hat{\mathbb{E}}[\Delta X_n \mid \mathcal{F}_{n-1}] = \frac{\mathbb{E}[L_N \Delta X_n \mid \mathcal{F}_{n-1}]}{\mathbb{E}[L_N \mid \mathcal{F}_{n-1}]}= \sum_{j=0}^d v_j \hat{P}_{n,j}= \mathbf{v}\hat{P}_n = A_n\end{equation*}
for all n. On the other hand, from (2.7), we have
\begin{equation*} Y= Y_N = Y_n + \sum_{i=n+1}^N \big({-}g_i(Z_i) + Z_i^\top \Delta X_i\big) = Y_n + \sum_{i=n+1}^N \big({-}B_i + Z_i^{\top}(\Delta X_i - A_i)\big).\end{equation*}
Taking the conditional expectation under
$\hat{\mathbb{P}}$
, we get the conclusion.
Example 2.3. Let
$N=1$
,
$d=1$
,
$\Omega = \{+,-\}$
,
$v_0 = -1, v_1 = 1$
, and
$\Delta X_1(\pm) = \pm 1$
. Then
$L = \mathbb{Z}$
,
$\mathbf{v} = ({-}1,1)$
,
$\Theta = [-1,1]$
and
$\mathbf{vv}^\top = 2$
. Following the proof of Theorem 2.1, the solution of the linear BS
$\Delta$
E
$\Delta Y_1 = -AZ_1 + Z_1\Delta X_1$
can be constructed as
for any
$A \in \mathbb{R}$
. The expression given by Proposition 2.2 is
$Y_0 = \hat{E}[Y_1]$
, which can be directly seen with
$\hat{P}_1 = ((1-A)/2, (1+A)/2)^\top$
for
$A \in \Theta = [-1,1]$
. For
$A \notin \Theta$
, we observe that
$Y_0$
is not increasing in either
$Y_1(+)$
or
$Y_1({-})$
. In particular,
$Y_0$
cannot be represented as an expectation in this case.
The set
$\Theta$
plays a key role also for a comparison theorem. Let
$\mathcal{B}$
denote the set of the sequence
$g = \{g_n\}_{n=1}^N \in \mathcal{A}$
with
for all
$z_1, z_2 \in \mathbb{R}^d$
.
Proposition 2.3. (Comparison theorem.) For
$i=1,2$
, let
$Y^{(i)}$
be
$\mathscr{F}_N$
-measurable random variables with
$Y^{(1)} \geq Y^{(2)}$
, and
$g^{(i)} = \{g^{(i)}_n\} \in \mathcal{A}$
with
$g^{(1)}_n \geq g^{(2)}_n$
. Let
$\mathcal{E}^{(i)}_n(Y^{(i)})$
denote
$\mathcal{E}^g_n(Y^{(i)})$
for
$g = g^{(i)}$
,
$i=1,2$
respectively. Assume also
$g^{(i)} \in \mathcal{B}$
for either
$i=1$
or
$i=2$
. Then
Proof. Note first that
for any
$z \in \mathbb{R}^d$
and n. Therefore, under (2.10) for
$g= g^{(i)}$
,
$g^{(i)}$
is balanced in the terminology of [Reference Cohen and Elliot15], so the result follows from Theorem 3.2 of [Reference Cohen and Elliot15]. Here we repeat essentially the same proof for the readers’ convenience. Let
$\bigl\{\bigl(Y^{(i)}_n,Z^{(i)}_n\bigr)\bigr\}$
be the solution of (2.7) with
$g = g^{(i)}$
and
$Y_N = Y^{(i)}$
. We have
$Y^{(1)}_N \geq Y^{(2)}_N$
by assumption. Suppose
$Y^{(1)}_k \geq Y^{(2)}_k$
for some k. Then
and so
Since the right-hand side is
$\mathscr{F}_{k-1}$
-measurable, this implies further
by (2.11). Therefore
\begin{align*} Y^{(1)}_{k-1} - Y^{(2)}_{k-1} & \geq g^{(1)}_k\bigl(Z^{(1)}_k\bigr) - g^{(2)}_k\bigl(Z^{(2)}_k\bigr) - \min_{\theta \in \Theta} \theta^\top \bigl(Z^{(1)}_k- Z^{(2)}_k\bigr) \\* & = g^{(1)}_k\bigl(Z^{(2)}_k\bigr) - g^{(2)}_k\bigl(Z^{(2)}_k\bigr) +g^{(1)}_k\bigl(Z^{(1)}_k\bigr) - g^{(1)}_k\bigl(Z^{(2)}_k\bigr) - \min_{\theta \in \Theta} \theta^\top \bigl(Z^{(1)}_k- Z^{(2)}_k\bigr) \\* & \geq 0 \end{align*}
under (2.10) for
$g_n = g^{(1)}_n$
, and also
\begin{align*} Y^{(1)}_{k-1} - Y^{(2)}_{k-1} & \geq g^{(1)}_k\bigl(Z^{(1)}_k\bigr) - g^{(2)}_k\bigl(Z^{(2)}_k\bigr) - \min_{\theta \in \Theta} \theta^\top \bigl(Z^{(1)}_k- Z^{(2)}_k\bigr) \\* & = g^{(1)}_k\bigl(Z^{(1)}_k\bigr) - g^{(2)}_k\bigl(Z^{(1)}_k\bigr) +g^{(2)}_k\bigl(Z^{(1)}_k\bigr) - g^{(2)}_k\bigl(Z^{(2)}_k\bigr) - \min_{\theta \in \Theta} \theta^\top \bigl(Z^{(1)}_k- Z^{(2)}_k\bigr) \\* & \geq 0 \end{align*}
under (2.10) for
$g_n = g^{(2)}_n$
. The result then follows by induction.
Remark 2.3. A sufficient condition for
$g_n$
to meet (2.10) is that
$g_n(z)$
is continuously differentiable in z with
$\nabla g_n(z)$
taking values in
$\Theta$
. Indeed, by Taylor’s theorem,
and then notice that
$A_n$
is
$\Theta$
-valued because
$\Theta$
is a convex set.
Example 2.4.
(
Locally entropic monetary utility.) Let
$\{(\hat{P}_{n,0},\ldots,\hat{P}_{n,d})^\top\}$
be a
$\Delta_d$
-valued predictable process, let
$\{B_n\}$
and
$\{\Gamma_n\}$
be positive predictable processes, and
\begin{equation} g_n(z) = -\frac{1}{\Gamma_n} \log \Biggl(\sum_{j=0}^{d} \,{\mathrm{e}}^{-\Gamma_n z^\top v_j} \hat{P}_{n,j}\Biggr) - \frac{1}{\Gamma_n}\log B_n.\end{equation}
Using a similar calculation to Example 2.1, we deduce the relation
In particular, when
$B_n=1$
,
${\mathcal E}_{n-1}^g$
is locally the minus of the entropic risk measure with risk-aversion parameter
$\Gamma_n$
extending (2.9). In contrast to the dynamic entropic risk measure studied in [Reference Acciaio and Penner1], we have the time-consistency property
$\mathcal{E}_m^g(\mathcal{E}_n^g(Y))=\mathcal{E}_m^g(Y)$
for any
$m \leq n$
when
$B_n = 1$
for all n even if the process
$\{\Gamma_n\}$
is not constant. We allow
$B_n \neq 1$
in order to include an example in Section 3. We call
$\mathcal{E}^g_n$
a locally entropic monetary utility. A brief numerical study for this utility is provided in Appendix B. We have
where
$\hat{P}_n(z) = (\hat{P}_{n,0}(z),\ldots,\hat{P}_{n,d}(z))^\top$
and
Since
$\hat{P}_n(z)$
is continuous in z and
$\Delta_d$
-valued for all n, by Remark 2.3, the assumptions of Proposition 2.3 on
$g^{(i)}_n$
are satisfied.
Next, we seek a robust representation of
$\mathcal{E}^g$
when g is concave. Let
$\mathcal{C}$
denote the set of
$g = \{g_n\} \in \mathcal{B}$
with
$g_n(z)$
being concave in z for all n.
Lemma 2.5. Let
$g = \{g_n\} \in \mathcal{A}$
. Then
$g \in \mathcal{C}$
if and only if
where
Proof. If
$g_n$
is concave, then it is continuous on the interior of its domain that is
$\mathbb{R}^d$
. Therefore by a well-known fact on the Legendre transform, we have
Let
$x \notin \Theta$
. Since
$\Theta$
is a closed convex set of
$\mathbb{R}^d$
, by the Hahn–Banach theorem (or the separating hyperplane theorem), there exists
$z_0 \in \mathbb{R}^d$
such that
Using (2.10), for
$z = \alpha z_0$
,
$\alpha \gt 0$
,
Since the last term is positive, letting
$\alpha \to \infty$
, we conclude
$b_n(x)=\infty$
. This implies
where
$A_n = \{(\theta,b) \in \Theta \times \mathbb{R} \, ; \, g_n(w) \leq w^\top \theta + b \text{ for all } w \in \mathbb{R}^d\}$
. Fix n and z and then take a sequence
$\{(\theta_k,b_k)\} \subset A_n$
such that
$z^\top \theta_k + b_k \to g_n(z)$
. Since
$\Theta$
is compact, there exists a converging subsequence
$\{\theta_{k_j}\}$
with limit
$\theta_\ast \in \Theta$
. We have
$b_{k_j} = z^\top\theta_{k_j} + b_{k_j} - z^\top \theta_{k_j} \to g_n(z) - z^\top \theta_\ast = \! : \, b_\ast$
. Also,
$w^\top \theta_{k_j} + b_{k_j} \geq g_n(w)$
for all w implies
$w^\top \theta_\ast + b_\ast \geq g_n(w)$
for all w, hence
$(\theta_\ast,b_\ast) \in A_n$
. Thus we obtain (2.13). Conversely, if (2.13) is true, then
$g_n(z)$
is concave, being the minimum of concave (affine) functions. Since
Let
$\mathcal{P}_N$
denote the set of the probability measures on
$(\Omega,\mathscr{F}_N)$
absolutely continuous with respect to
$\mathbb{P}$
. For
$\hat{\mathbb{P}} \in \mathcal{P}_N$
, there corresponds a
$\Delta_d$
-valued predictable process
$\{\hat{P}_n\}$
by (2.4). The measure
$\hat{\mathbb{P}}$
is recovered from
$\{\hat{P}_n\}$
by (2.5). Let
$\hat{\mathbb{E}}$
denote the expectation under
$\hat{\mathbb{P}}$
and define
\begin{equation*} c^g_n(\hat{\mathbb{P}}) = \hat{\mathbb{E}}\Biggl[\sum_{i=n+1}^N b_i(\mathbf{v}\hat{P}_i)\bigg|\mathscr{F}_n\Biggr],\end{equation*}
where
$b_i$
is associated with
$g = \{g_n\} \in \mathcal{C}$
via (2.13). The following theorem shows the nature of the g-expectation with the gradient constraint as a nonlinear expectation, taking care of Knightian uncertainty, refining a general convex duality result in [Reference Detlefsen and Scandolo18] for an explicit representation of a penalty function.
Theorem 2.2. Let
$g \in \mathcal{C}$
. Then, for any
$\mathscr{F}_N$
-measurable random variable Y,
Proof. By (2.13), we have
for any
$\hat{\mathbb{P}} \in \mathcal{P}_N$
. Therefore, by Propositions 2.2 and 2.3, we have
for any
$\hat{\mathbb{P}} \in \mathcal{P}_N$
. On the other hand, for any
$Y \in \mathscr{F}_N$
, there exists the solution
$\{(Y_n,Z_n)\}$
of (2.7) with
$Y_N= Y$
. By (2.13), there exists
$\hat{P}_n$
such that
for each n. Since
$\{g_n\}$
and
$\{Z_n\}$
are predictable,
$\{\hat{P}_n\}$
is a
$\Delta_d$
-valued predictable process. Let
$\hat{\mathbb{P}} \in \mathcal{P}_N$
be associated with
$\{\hat{P}_n\}$
. Then
$\{(Y_n,Z_n)\}$
solves the BSDE with
$\hat{g}_n(z) = z^\top \mathbf{v}\hat{P}_n + b_n(\mathbf{v}\hat{P}_n)$
as well, and so by Proposition 2.2,
which implies (2.14).
Corollary 2.1. Let
$g \in \mathcal{C}$
. Let Y and
$Y^\prime$
be
$\mathscr{F}_N$
-measurable random variables.
-
(i) If
$Y \geq Y^\prime$
, then
$\mathcal{E}^g_n(Y) \geq \mathcal{E}^g_n(Y^\prime)$
,
$n=0,1,\ldots, N$
. -
(ii) For any
$\mathscr{F}_n$
-measurable [0,1]-valued random variable
$\lambda$
,
\begin{equation*} \mathcal{E}^g_n(\lambda Y + (1-\lambda)Y^\prime) \geq \lambda\mathcal{E}^g_n(Y) + (1-\lambda) \mathcal{E}^g_n(Y^\prime), \quad n=0,1,\ldots, N. \end{equation*}
Example 2.5. Let
$\Theta_n \subset \Theta$
and
Here, the set
$\Theta_n$
can be random in such a way that
$g_n$
is
$\mathscr{F}_{n-1}\otimes \mathscr{B}(\mathbb{R}^d)$
-measurable. Then we have (2.13) with
$b_n$
such that
$b_n(\theta) = 0$
if
$\theta \in \bar{\Theta}_n$
while
$b_n(\theta) = \infty$
otherwise, where
$\bar{\Theta}_n$
is the closure of
$\Theta_n$
. In particular when
$\Theta_n = \Theta$
, by Theorem 2.2,
for any
$\mathscr{F}_N$
-measurable random variable Y. Note also that
When
$\Theta_n = \{\mathbf{v}P_n^{(1)}, \ldots, \mathbf{v}P^{(m)}_n\}$
for a
$\Delta_d$
-valued predictable process
$\{P^{(i)}_n\}$
, letting
$\mathbb{E}^{(x)}$
denote the expectation under the measure determined by
$\{P^{(x_n)}_n\}$
for
$x = (x_1,\ldots,x_N) \in \{1,\ldots,m\}^N$
, by Theorem 2.2, we have
2.5. Filtration consistent nonlinear expectations
Inspired by Coquet et al. [Reference Coquet, Hu, Memin and Peng16], we call
$\mathcal{E}\colon L^0(\Omega,\mathscr{F}_N,\mathbb{P}) \to \mathbb{R}$
a filtration consistent nonlinear expectation if:
-
(i)
$Y \geq Y^\prime \Rightarrow \mathcal{E}(Y) \geq \mathcal{E}(Y^\prime)$
, -
(ii)
$Y \geq Y^\prime$
and
$ \mathcal{E}(Y) = \mathcal{E}(Y^\prime) \Rightarrow Y = Y^\prime$
, -
(iii)
$\mathcal{E}(c) = c$
for any constant
$c \in \mathbb{R}$
, and -
(iv) for any
$n=1,\ldots,N$
and Y, there exists an
$\mathscr{F}_n$
-measurable
$\eta$
such that
$\mathcal{E}(Y1_A) = \mathcal{E}(\eta 1_A)$
for any
$A \in \mathscr{F}_n$
.
Further,
$\eta$
is uniquely determined as shown in [Reference Coquet, Hu, Memin and Peng16]. Let it be denoted by
$\mathcal{E}_n(Y)$
. It follows that
for any
$m \geq n$
, and
for any
$A \in \mathscr{F}_n$
.
Proposition 2.4. Let
$g = \{g_n \} \in \mathcal{A}$
, and assume that for
$n = 1,\ldots, N$
,
-
(i)
$g_n(0) = 0$
and -
(ii) for any
$z_1, z_2 \in \mathbb{R}^d$
with equality holding only if
\begin{equation*} g_n(z_1) -g_n(z_2) \geq \min_{\theta \in \Theta} \theta^\top (z_1-z_2),\end{equation*}
$z_1 = z_2$
.
Then
$\mathcal{E}^g_0$
is a filtration consistent nonlinear expectation with
$\mathcal{E}_n = \mathcal{E}^g_n$
and a translation invariance property,
for any
$\mathscr{F}_N$
-measurable random variable Y,
$\mathscr{F}_n$
-measurable random variable
$\eta$
, and
$n=1,\ldots, N$
.
Proof. From Proposition 2.3 and its proof, we observe the first two properties of filtration consistent nonlinear expectation. By
$g_n(0) = 0$
we derive (2.15) and (2.16), from which the other properties follow.
The following theorem is a discrete analog of Theorem 7.1 of [Reference Coquet, Hu, Memin and Peng16].
Proposition 2.5. Let
$\mathcal{E}$
be a filtration consistent nonlinear expectation with the translation invariance property (2.17). Let
$g_n(z) = \mathcal{E}_{n-1}(z^\top \Delta X_n)$
. Then
$\mathcal{E}_n = \mathcal{E}^g_n$
.
Proof. We have
$\mathcal{E}_N(Y) = Y$
for any Y, so the claim is true for
$n=N$
. Assume
$\mathcal{E}_k(Y) = \mathcal{E}^g_k(Y)$
for
$k \geq n$
. Then, by (2.2), there exists
$\mathscr{F}_{n-1}$
-measurable
$A_n$
and
$Z_n$
such that
The last term is
$g_n(Z_n)$
by (2.16). Therefore
which implies that
$\mathcal{E}_{n-1}(Y) = \mathcal{E}_{n-1}^g(Y)$
. The result follows by induction.
3. Market equilibrium analysis
3.1. Monetary utility maximization
Now we consider
$\{X_n\}$
to be a d-dimensional asset price process. The lattice L is then understood as the price grid. For any
$\{\mathscr{F}_n\}$
-predictable
$\mathbb{R}^d$
-valued process
$Z = \{Z_n\}$
and
$w \in \mathbb{R}$
,
represents the wealth process associated with the portfolio strategy Z and the initial wealth w. Applying Theorem 2.1 with
$g_n = 0$
,
$n=1,\ldots, N$
, we observe that the market is complete, that is, for any
$\mathscr{F}_N$
-measurable random variable Y, there exists a predictable process Z and
$w \in \mathbb{R}$
such that
$Y = W_N( w,Z)$
.
Consider an agent whose utility functional is
$\mathcal{E}^g_0$
for
$g \in \mathcal{C}$
. This utility is monetary in the sense that for all
$n = 0, \ldots, N$
,
$\mathcal{E}_n^g(Y+A) = \mathcal{E}_n^g(Y)+A$
for any
$\mathscr{F}_N$
-measurable random variable Y and
$\mathscr{F}_n$
-measurable random variable A. When assuming
$g_n(0) = 0$
for all n in addition, the utility is normalized in the sense that
$\mathcal{E}^g_n(0)=0$
,
$n =0,\ldots, N$
, and it is time-consistent in the sense that
$\mathcal{E}_m^g(\mathcal{E}_n^g(Y))=\mathcal{E}_m^g(Y)$
for any
$m \leq n$
and for any
$\mathscr{F}_N$
-measurable random variable Y. The simplest example is
$\mathcal{E}^g_n(Y) = \mathbb{E}[Y\mid \mathscr{F}_n]$
corresponding to
$g_n(z) = z^\top \mathbb{E}[\Delta X_n \mid \mathscr{F}_{n-1}]$
. More generally,
$\mathcal{E}^g_n(Y)$
is a conditional expectation with respect to a probability measure when
$g_n$
are linear for all n by Proposition 2.2. The driver
$\{g_n\}$
should reflect the agent’s belief in the distribution of the price process
$\{X_n\}$
. For example, if
$g_n = 0$
for all n, then
$\mathcal{E}^g_n(Y) = \mathbb{E}_\mathbb{Q}[Y\mid \mathscr{F}_n]$
irrespective of
$\mathbb{P}$
. The choice of nonlinear
$\{g_n\}$
accommodates a nonlinear evaluation of risk, extending the exponential utility (2.9). In light of Theorem 2.2, our problem can be interpreted as a robust utility maximization.
The agent’s objective is to maximize
$\mathcal{E}^g_0(H + W_N(w,\pi))$
among the predictable process
$\pi = \{\pi_n\}$
, where H is a given
$\mathscr{F}_N$
-measurable random variable representing an initial endowment of the agent, i.e. an initially endowed asset or a scheduled random cashflow. Since the utility is monetary, it suffices to treat the case
$w = 0$
. Let
Then the problem is equivalent to maximizing
$Y^\pi_0$
among
$\pi$
. The following theorem characterizes the maximizer.
Theorem 3.1. Assume that there exists a predictable process
$\{Z^\dagger_n\}$
such that
Then
where
$\pi^\ast_n = Z^\dagger_n - Z^\ast_n$
,
$ Z^\ast_n = Z^H_n + Z^g_n$
,
and
\begin{equation*} G_n = \sum_{i=n+1}^N \mathbb{E}_\mathbb{Q}[ g_i(Z_i^\dagger) \mid \mathscr{F}_n]\end{equation*}
for
$n=1,\ldots, N$
. Moreover,
for
$n=0,1,\ldots, N$
. If in addition
$Z^\dagger_n$
is unique, then
$\pi^\ast_n$
is the unique maximizer.
Proof. Let
$Y^\ast_n$
denote the right-hand side of (3.4). Then
$\{(Y^\ast_n,Z^\ast_n)\}$
is the solution of the BS
$\Delta$
E
by Theorem 2.1. Note that
for a predictable process
$\{Z_n\}$
by (3.1). This means
$\{(Y^\pi_n,Z^\pi_n)\}$
,
$Z^\pi_n = Z_n -\pi_n$
solves the BS
$\Delta$
E
where
$h^\pi_n(z) = g_n(z+ \pi_n)$
. For any predictable process
$\pi$
, we have
$h^\pi_n \leq g_n(Z^\dagger_n)$
. Therefore
$Y^\pi_n \leq Y^\ast_n$
by Proposition 2.3. Notice also that by choosing
$\pi^\ast_n = Z^\dagger_n -Z^\ast_n$
we have
$g_n(Z^\dagger) = h_n^{\pi^\ast}(Z^\ast_n)$
, so that
$\{(Y^\ast_n,Z^\ast_n)\}$
satisfies the same BS
$\Delta$
E as
$\{(Y^{\pi^\ast}_n,Z^{\pi^\ast}_n)\}$
. Hence
$Y^\ast_n = Y^{\pi^\ast}_n$
.
Remark 3.1. Applying Theorem 2.1 to
$g_n = 0$
and
$Y_N = H$
, we have
\begin{equation*} H = \mathbb{E}_\mathbb{Q}[H] + \sum_{n=1}^N Z^H_n \Delta X_n \end{equation*}
with
$\{Z^H_n\}$
defined by (3.3). Therefore the optimal strategy
$\pi^\ast = Z^\dagger_n - Z^\ast_n = -Z^H_n + Z^\dagger_n - Z^g_n$
of Theorem 3.1 is decomposed into the hedging part
$-Z^H_n$
and the optimal investment part
$Z^\dagger_n - Z^g_n$
.
Example 3.1. (Locally entropic monetary utility.) Consider (2.12). Let
$\Delta_d^\circ$
denote the interior of
$\Delta_d$
and assume
$\{(\hat{P}_{n,1},\ldots, \hat{P}_{n,d})^\top\}$
to be
$\Delta_d^\circ$
-valued. Then the map
$z \mapsto g_n(z)$
is strictly concave and its unique maximizer is given by
$Z^\dagger_n = b(y)$
of (2.3) for
or equivalently
Indeed,
$v_j^\top Z^\dagger_n = \Gamma_n^{-1}\log \hat{P}_{n,j} - a(y)$
implies
\begin{equation*} \sum_{j=0}^d v_j \,{\mathrm{e}}^{-\Gamma_n v_j^\top Z^\dagger_n} \hat{P}_{n,j}= 0, \end{equation*}
and thus
$\nabla g_n(Z^\dagger_n) =0$
for all n. We also have
\begin{align} \frac{1}{\Gamma_n}\log B_n + g_n(Z^\dagger_n) &= -\frac{1}{\Gamma_n} \log \Biggl( \sum_{j=0}^d \,{\mathrm{e}}^{-\Gamma_n v_j^\top Z^\dagger_n} \hat{P}_{n,j}\Biggr)\notag \\*& =-\frac{1}{\Gamma_n} \log \Bigl((d+1) \,{\mathrm{e}}^{\frac{1}{d+1}\mathbf{1}^\top \log \hat{P}_n }\Bigr)\notag \\ &= \frac{1}{\Gamma_n} \frac{1}{d+1}\sum_{j=0}^d \log \frac{1}{(d+1)\hat{P}_{n,j}} \notag \\*&= \frac{1}{\Gamma_n}D_{\mathrm{KL}}(Q_n ||\hat{P}_n) \notag \\*&\geq 0, \end{align}
where
$D_{\mathrm{KL}}$
denotes the Kullback–Leibler divergence on
$\Delta_d$
and
$Q_n = \mathbf{1}/(d+1)$
. By (3.3) and (3.6),
where
$\hat{Y}_n = (\hat{Y}_{n,0},\ldots, \hat{Y}_{n,d})^\top$
and
$ \hat{Y}_{n,j} = \mathbb{E}_\mathbb{Q}[G_n \mid \mathscr{F}_{n-1}, \Delta X_n = v_j]$
. The first term
$-Z^H_n$
is the hedging term as noted in Remark 3.1. The term
can be interpreted as the discrete counterpart of the Merton portfolio (see e.g. Remark 8.9 of [Reference Karatzas and Shreve27]). The term
$\hat{Y}_n$
adjusts the expected return depending on the stochastic dynamics of
$D_{\mathrm{KL}}(Q_i ||\hat{P}_i)$
for
$i \geq n+1$
. Indeed, when
$B_n = 1$
and
$\Gamma_n = \gamma$
for all n and a constant
$\gamma \gt 0$
as in (2.8), we have
\begin{align*} G_n &= \sum_{i=n+1}^n \mathbb{E}_\mathbb{Q}[g_i(Z^\dagger_i) \mid \mathscr{F}_n]\\ &= \frac{1}{\gamma}\sum_{i=n+1}^n \mathbb{E}_\mathbb{Q}[D_{\mathrm{KL}}(Q_i||\hat{P}_i) | \mathscr{F}_n]\\ &= \frac{1}{\gamma} \mathbb{E}_\mathbb{Q}\biggl[ \log \frac{\mathrm{d}\mathbb{Q}}{\mathrm{d}\hat{\mathbb{P}}} - \log \mathbb{E}_\mathbb{Q}\biggl[ \frac{\mathrm{d}\mathbb{Q}}{\mathrm{d}\hat{\mathbb{P}}}\,\bigg|\, \mathscr{F}_n\biggr]\, \bigg|\, \mathscr{F}_n\biggr]\end{align*}
by (3.7), where
$\hat{\mathbb{P}}$
is associated with
$\{\hat{P}_n\}$
by (2.5). Note also that
$Z^g_n = 0$
if
$G_n$
is deterministic, which is the case when
$\{\hat{P}_n\}$
,
$\{\Gamma_n\}$
, and
$\{B_n\}$
are deterministic.
Example 3.2. Let
$d=2$
and
where
$c \in \mathbb{R}$
. Then we have
\begin{equation*} X_{n,2} = c\Biggl(3 \sum_{k=1}^n |\Delta X_{k,1}|^2 - 2n\Biggr).\end{equation*}
Indeed, if
$X_n = (n-a-b)v_0 + av_1 + bv_2$
for
$(a,b) \in \mathbb{N}^2$
, then
Regarding
$\{X_{n,1}\}$
as a price process of an asset, the above identity allows us to interpret
$X_{N,2}$
as an affine transform of the variance swap payoff of the asset. Further, regarding
$\mathbb{Q}$
as the pricing measure,
$\{X_{n,2}\}$
corresponds to the price process of the variance swap payoff. Note that
$\{X_{n,1}\}$
describes a trinomial model for the asset, which is not complete. The variance swap trading makes the two-dimensional market
$\{X_n\}$
complete.
3.2. General equilibrium
Consider m agents who maximize respective utilities
$\mathcal{E}_0^{(i)}(H^{(i)} + W_N(0,\pi^{(i)}))$
,
$i=1,\ldots, m$
through trading strategies
$\pi^{(i)}$
of the d-dimensional asset
$\{X_n\}$
, where
$\mathcal{E}^{(i)}$
is the solution map of the BS
$\Delta$
E (2.7) with
$g = g^{(i)} \in \mathcal{C}$
, and
$H^{(i)}$
is an
$\mathscr{F}_N$
-measurable random variable representing an endowment for the agent i,
$i=1,\ldots,m$
. Let
$H_n$
denote the total supply vector of the asset vector
$X_n$
and assume
$\{H_n\}$
to be an
$\mathbb{R}^d$
-valued predictable process. We say the market is in general equilibrium if there exist predictable processes
$\pi^{(i)} = \{\pi^{(i)}_n\}$
,
$i=1,\ldots, m$
such that
-
(i)
$\mathcal{E}_0^{(i)}(H^{(i)} + W_N(0,\pi^{(i)})) = \max_\pi \mathcal{E}_0^{(i)}(H^{(i)} + W_N(0,\pi))$
for all
$i=1,\ldots, m$
, and -
(ii)
$\sum_{i=1}^m \pi^{(i)}_n = H_n $
for all
$n=1,\ldots, N$
,
where the maximum is among all predictable processes
$\pi$
.
Proposition 3.1. Let
$H^{(i)}$
and
$\tilde{H}^{(i)}$
,
$i=1,\ldots,m$
be
$\mathscr{F}_N$
-measurable random variables, and let
$\{H_n\}$
and
$\{\tilde{H}_n\}$
be
$\mathbb{R}^d$
-valued predictable processes satisfying
\begin{equation*} \sum_{i=1}^m H^{(i)} + \sum_{n=1}^N H_n^\top \Delta X_n = \sum_{i=1}^m \tilde{H}^{(i)} + \sum_{n=1}^N \tilde{H}_n^\top \Delta X_n. \end{equation*}
Then the market with the endowments
$H^{(i)}$
and the total supply
$\{H_n\}$
is in general equilibrium if and only if the market with the endowments
$\tilde{H}^{(i)}$
and total supply
$\{\tilde{H}_n\}$
is in general equilibrium.
Proof. Let
$Z^{\dagger(i)}$
,
$Z^{H(i)}$
, and
$Z^{g(i)}$
, respectively, denote
$Z^{\dagger}$
,
$Z^{H}$
, and
$Z^{g}$
in Theorem 3.1 with
$g = g^{(i)}$
and
$H = H^{(i)}$
. Then, by Theorem 3.1,
where
$Z^H$
is defined by (3.3) with
\begin{equation*} H = \sum_{i=1}^m H^{(i)} + \sum_{n=1}^N H_n^\top \Delta X_n.\end{equation*}
Therefore, whether or not
$\sum_{i=1}^m \pi^{(i)}_n =H_n$
depends on
$H^{(i)}$
and
$\{H_n\}$
only through H.
We are interested in conditions on
$g^{(i)}$
for the market to be in general equilibrium. In light of Proposition 3.1, we assume hereafter
$H^{(i)}=0$
(
$i \geq 2)$
and
$H_n =0$
(
$n\geq 1$
), without loss of generality. Let H denote
$H^{(1)}$
.
For functions
$f^{(1)}$
and
$f^{(2)}$
on
$\mathbb{R}^d$
, define the sup-convolution
$f^{(1)} \square f^{(2)}$
by
For the drivers
$g^{(i)} = \{g^{(i)}_n\}$
,
$i=1,\ldots,m$
of the m agents’ utilities, let
Lemma 3.1. For all n,
Proof. It is trivial that
$g_n(z)$
is upper-bounded by the right-hand side sum for any z, and hence its supremum is upper-bounded. Conversely, let
$\bigl\{z^{(i)}_k\bigr\}$
be a sequence for which
$ \lim_{k \to \infty} g_n^{(i)}\bigl(z^{(i)}_k\bigr) = \sup_{z \in \mathbb{R}^d} g_n^{(i)}(z)$
for each i. Then
\begin{equation*} \sum_{i=1}^m g_n^{(i)}\bigl(z^{(i)}_k\bigr) \leq g_n\Biggl(\sum_{i=1}^m z^{(i)}_k \Biggr) \leq \sup_{z \in \mathbb{R}^d} g_n(z)\end{equation*}
for any k, and the limit is similarly bounded.
A single agent whose utility is
$\mathcal{E}^g_0$
with
$g = \{g_n\}$
defined by (3.8) and whose endowment is H is called the representative agent of the market. The following theorem reduces the general equilibrium problem for the multi-agent market to the one for a single agent market. This extends the idea of the well-known Gorman aggregation theorem.
Proposition 3.2. Assume that there exists a unique predictable process
$\{Z^{\dagger(i)}_n\}$
for each
$i=1,\ldots,m$
such that
for all
$n = 1,\ldots, N$
. Assume further that the maximizer of the map
$z \mapsto g_n(z)(\omega)$
is unique for each n and
$\omega \in \Omega$
. The market for the m agents is in general equilibrium if and only the market for the representative agent is in general equilibrium.
Proof. Let
$\{(Y^{\ast(i)}_n,Z^{\ast(i)}_n)\}$
be the solution of
with
$Y^{\ast (1)}_N = H$
and
$Y^{\ast (i)}_N = 0$
for
$i \geq 2$
. By Theorem 3.1, the unique optimal strategy
$\pi^{(i)}$
for the agent i is given by
$\pi^{(i)}_n = Z^{\dagger (i)}_n -Z^{\ast(i)}_n $
. By (3.9), we have
Therefore
solves
Since
$Z^\dagger_n$
is the unique maximizer of
$g_n$
, the unique optimal strategy for the representative agent is
$\pi^\ast_n = Z^\dagger_n - Z^\ast_n$
, again by Theorem 3.1. Hence
Therefore
$\sum_{i=1}^m \pi^i_n = 0$
if and only if
$\pi^\ast_n = 0$
.
3.3. Equilibrium in a single agent market
By Proposition 3.2, it suffices to consider the case
$m=1$
in order to characterize the general equilibrium. Let
$m=1$
, and put
$g_n = g^{(1)}_n$
and
$\pi_n = \pi^{(1)}_n$
. The market is in general equilibrium if and only if
$\mathcal{E}_0^g (H) = \max_\pi \mathcal{E}_0^{g}(H + W_N(0,\pi))$
, that is,
$\pi^\ast_n \equiv 0$
is the maximizer. The following theorems characterize the general equilibrium by a backward recurrence relation for the BS
$\Delta$
E driver
$\{g_n\}$
.
Theorem 3.2. The market is in general equilibrium if (3.2) holds with
where
\begin{equation*} G_n =\mathbb{E}_\mathbb{Q}\Biggl[ \sum_{i=n+1}^N \sup_{z\in \mathbb{R}^d} g_i(z) \bigg| \mathscr{F}_n\Biggr].\end{equation*}
Conversely, if the market is in general equilibrium and there exist sequences of maximizers
$\{Z^\dagger_n\}$
of
$\{g_n\}$
, then (3.10) is one of them.
Proof. Notice that the right-hand side of (3.10) coincides with
$Z^\ast_n$
defined in Theorem 3.1. If (3.2) holds with (3.10), then we conclude that
$\pi^\ast_n = 0$
is optimal by Theorem 3.1, which means that the market is in general equilibrium. Conversely, if the market is in general equilibrium, then by Theorem 3.1,
$Z^\ast_n$
should coincide with a maximizer of
$g_n$
if any.
Proposition 3.3. Let
$f_n\colon \Omega \times \mathbb{R}^d \times \Delta_d \to \mathbb{R}$
be
$\mathscr{F}_{n-1}\otimes \mathscr{B}(\mathbb{R}^d \times \Delta_d)$
-measurable, concave on
$\mathbb{R}^d$
and continuously differentiable on
$\mathbb{R}^d$
with
$\nabla f_n$
taking values in
$\Theta$
, and
$0 \in \nabla f_n(z,\Delta_d)$
for all
$z \in \mathbb{R}^d$
,
$n=1,\ldots, N$
. Then there exists a
$\Delta_d$
-valued predictable process
$\{\hat{P}_n\}$
such that the market with
$g =\{g_n\}$
,
$g_n(z) = f_n(z,\hat{P}_n)$
, is in general equilibrium.
Proof. Note first that
$g \in \mathcal{C}$
by Remark 2.3. We construct
$\{\hat{P}_n\}$
inductively. By the assumption, there exists
$\hat{P}_N$
such that
$\nabla f_N(Z^\dagger_N,\hat{P}_N) = 0$
for
$Z^\dagger_N$
defined by (3.10) for
$n=N$
. Given
$\hat{P}_k$
for
$k \geq n+1 $
, let
$Z^\dagger_n$
be defined by (3.10) with
$g_k(z) = f_k\bigr(z,\hat{P}_k\bigr)$
. By the assumption, there exists
$\hat{P}_n$
such that
$\nabla f_n(Z^\dagger_n,\hat{P}_n) = 0$
. Since
$z \mapsto f_n(z,\hat{P}_n)$
is concave,
$Z^\dagger_n$
is a maximizer of
$g_n(z) = f_n(z,\hat{P}_n)$
. The result then follows from Theorem 3.2.
Example 3.3.
(
Locally entropic monetary utility.) We consider (2.12) again with
$\{\hat{P}_n\}$
,
$\hat{P}_n = (\hat{P}_{n,1},\ldots, \hat{P}_{n,d})^\top$
, being
$\Delta_d^\circ$
-valued. The driver
$g_n$
is of the form
$g_n(z) = (f(\Gamma_nz,\hat{P}_n) -\log B_n)/\Gamma_n $
, where
\begin{equation} f(z,p) = -\log \sum_{j=0}^d \,{\mathrm{e}}^{-z^\top v_j}p_j,\quad z \in \mathbb{R}^d, \quad p = (p_0,\ldots,p_d)^\top \in \Delta_d. \end{equation}
By (3.6), we have
$\nabla f_n(Z_n,\hat{P}_n) = 0$
if and only if
By Theorem 3.2, the market is in general equilibrium if and only if
This is a backward recurrence equation for
$\{\hat{P}_n\}$
because
\begin{equation*} G_n = \mathbb{E}_\mathbb{Q}\Biggl[ \sum_{i=n+1}^N \frac{1}{\Gamma_i} ( D_{\mathrm{KL}}(Q_i||\hat{P}_i) - \log B_i )\, \bigg|\, \mathscr{F}_n \Biggr]\end{equation*}
by (3.7). Since
$\mathbf{v}$
has rank d and
$\mathbf{v}\mathbf{1} = 0$
, an equivalent condition is that
for
$n=1,\ldots, N$
. Let
$ Y^\ast_n = \mathbb{E}_\mathbb{Q}[H\mid \mathscr{F}_n] + G_n$
. Then
\begin{equation*}Y_n = \mathbb{E}_\mathbb{Q}[Y_n\mid \mathscr{F}_{n-1},\Delta X_n] = \sum_{j=0}^d \bar{Y}_{n,j} 1_{\{\Delta X_n = v_j\}}, \end{equation*}
which implies
under (3.12). Therefore the market is in general equilibrium if and only if
\begin{equation*} \frac{\mathrm{d}\hat{\mathbb{P}}}{\mathrm{d}\mathbb{Q}}=\prod_{n=1}^N\frac{{\mathrm{e}}^{\Gamma_nY^\ast_n}}{\mathbb{E}_{\mathbb{Q}}[{\mathrm{e}}^{\Gamma_nY^\ast_n}\mid \mathscr{F}_{n-1}]}, \end{equation*}
where
$\hat{\mathbb{P}}$
is defined by (2.5). Further, by (3.5), we have
\begin{align*} \log \hat{\mathbb{E}}[{\mathrm{e}}^{-\Gamma_n Y^\ast_n}\mid \mathscr{F}_{n-1}]& = -\Gamma_n(Y^\ast_{n-1} - g_n(Z^\dagger_n)) +\log \hat{\mathbb{E}}[{\mathrm{e}}^{-\Gamma_n(\Delta X_n)^\top Z^\dagger_n}\mid \mathscr{F}_{n-1}] \\*&=-\Gamma_n(Y^\ast_{n-1} - g_n(Z^\dagger_n)) -f_n(\Gamma_n Z^\dagger_n, \hat{P}_n) \\*& =-\Gamma_nY^\ast_{n-1} - \log B_n. \end{align*}
This implies
\begin{equation} \frac{\mathrm{d}\mathbb{Q}}{\mathrm{d}\hat{\mathbb{P}}}=\prod_{n=1}^N\frac{{\mathrm{e}}^{-\Gamma_nY^\ast_n}}{\hat{\mathbb{E}}[{\mathrm{e}}^{-\Gamma_n Y^\ast_n}\mid \mathscr{F}_{n-1}]}= \exp\Biggl\{\sum_{n=1}^N -\Gamma_n \Delta Y^\ast_n \Biggr\} \prod_{n=1}^N B_n. \end{equation}
In particular, when
$B_n = 1$
and
$\Gamma_n = \gamma$
for all n and a constant
$\gamma \gt 0$
as in (2.8), we have
which characterizes the equilibrium probability measure
$\hat{\mathbb{P}}$
and is consistent with the well-known computation under exponential utility.
Remark 3.2. In addition to the conditions of Proposition 3.3, if
$f_n$
is of the form
$f_n(z,p) = f(\Gamma_n z,p)/\Gamma_n$
for a smooth function
$f\colon \mathbb{R}^d \times \Delta_d \to \mathbb{R}$
with
$\nabla f(0,p) = p$
and a positive predictable process
$\{\Gamma_n\}$
as in Example 3.3 with
$B_n = 1$
, then we have
considering
$\Gamma_n$
to be small. This approximation implies in turn
in light of Proposition 2.2, where
$\hat{\mathbb{E}}$
is the expectation under
$\hat{\mathbb{P}}$
defined by (2.5). Therefore, in this case, extending Example 3.3, we can interpret
$\Gamma_n$
as a risk-aversion parameter and
$\hat{\mathbb{P}}$
as the belief of the agent. If the market is in general equilibrium,
is interpreted as an equilibrium return.
Example 3.4. If
$H = h(X_N)$
and
$g_n(z) = f_n(X_{n-1},z)$
for deterministic functions h and
$f_n$
, as in Proposition 2.1, and if f(x,z) is strictly concave and continuously differentiable in z, then the condition (3.10) follows from the deterministic identity
where
$\mathcal{N}u_n$
is as in Proposition 2.1. For example, for the market with no random endowments (
$H^{(i)}=0$
) and unit total supply (
$H_n = \mathbf{1}$
), we have
$H = \mathbf{1}^\top X_N$
.
Example 3.5. Consider (2.12) with
$B_n = 1$
,
$\Gamma_n = \gamma_n(X_{n-1})$
and
$\hat{P}_n = p_n(X_{n-1})$
for deterministic functions
$\gamma_n\colon L \to (0,\infty)$
and
$p_n\colon L \to \Delta_d^\circ$
. Assume also
$H = h (X_N)$
as in Example 3.4. Then, from Examples 3.3 and 3.4, the market is in general equilibrium if
The function
$u_n(x)$
is computed backward inductively without using
$p_n(x)$
. For a given function
$\gamma_n(x)$
, there exists a unique
$p_n(x) \in \Delta_d^\circ$
satisfying this equation for each
$x \in L$
. For the sequence of such functions
$p_n(x)$
obtained in the backward manner, the
$\Delta_d$
-valued sequence
$\hat{P}_n = p_n(X_{n-1})$
defines a unique equilibrium probability
$\hat{\mathbb{P}}$
by (2.5) associated with the sequence
$\Gamma_n = \gamma_n(X_{n-1})$
. The equilibrium return is approximated as
using
$\mathbf{v1} = 0$
.
3.4. Equilibrium under heterogeneous beliefs
Here we assume
$m>1$
again and give more explicit computations of the sup-convolution in special cases. First, we consider a homogeneous case, i.e. the case where all of the drivers
$g^{(i)}$
have the same functional form determined by a common
$\Delta_d$
-valued predictable process
$\{\hat{P}_n\}$
as
where
$f_n\colon \Omega \times \mathbb{R}^d \times \Delta_d \to \mathbb{R}$
is as in Proposition 3.3 and
$\{\Gamma^{(i)}_n\}$
,
$i=1,\ldots,m$
are positive predictable processes quantifying each agent’s risk preference; see Remark 3.2. By induction, we can show that
with
Proposition 3.4. Under (3.15), there exists a
$\Delta_d$
-valued predictable process
$\{\hat{P}_n\}$
such that the market is in general equilibrium.
Example 3.6.
(
Locally entropic monetary utility.) Let
$f_n(z,p) = f(z,p)$
defined by (3.11). The predictable processes
$\{\Gamma^{(i)}_n\}$
and
$\{\Gamma_n\}$
are then interpreted as the local risk-aversion parameter for the agent i and for the representative agent respectively. Then (3.12) defines the unique sequence
$\{\hat{P}_n\}$
such that the market is in general equilibrium. The equilibrium probability measure
$\hat{\mathbb{P}}$
satisfies (3.13) with
$B_n = 1$
. When
$\Gamma_n = \gamma$
for all n for a constant
$\gamma \gt 0$
, then the representative agent has the exponential utility (2.9) and the equilibrium probability measure
$\hat{\mathbb{P}}$
is characterized by (3.14).
Now we consider a heterogeneous case. We assume locally entropic monetary utilities
$g^{(i)}_n(z) = f(\Gamma^{(i)}_nz,\hat{P}^{(i)}_n)/\Gamma^{(i)}_n$
, where f is defined by (3.11), and
$\{\Gamma^{(i)}_n\}$
and
respectively, are
$(0,\infty)$
-valued and
$\Delta_d^\circ$
-valued predictable processes for each
$i=1,\ldots,m$
. Each sequence
$\{\hat{P}^{(i)}_n\}$
determines a probability measure
$\hat{P}^{(i)}$
on
$\mathscr{F}_N$
by (2.5), which is interpreted as the agent i’s belief in the law of
$\{X_n\}$
.
Proposition 3.5. Define
$\{\Gamma_n\}$
by (3.16). Then
where
$ \tilde{P}_n = ( \tilde{P}_{n,0},\ldots, \tilde{P}_{n,d})^\top$
and
\begin{equation*} \tilde{P}_{n,j} = \frac{1}{B_n}\prod_{i=1}^m \bigl(\hat{P}^{(i)}_{n,j}\bigr)^{\Gamma_n/\Gamma_n^{(i)}}, \quad B_n = \sum_{j=0}^d \prod_{i=1}^m \bigl(\hat{P}^{(i)}_{n,j}\bigr)^{\Gamma_n/\Gamma_n^{(i)}}. \end{equation*}
Proof. The case
$m=2$
follows by solving the equation
in
$x \in \mathbb{R}^d$
; see Lemma A.1. The general case then follows by induction.
Remark 3.3. By Lemma A.2 we have
$B_n \leq 1$
, with equality holding if and only if
$\hat{P}^{(i)}_n =\hat{P}^{(1)}_n$
for all i.
By Proposition 3.5, the representative agent’s market falls into Example 3.3. In particular, when
$\Gamma_n = \gamma>0$
(a constant), the market is in general equilibrium if and only if
$\tilde{P}_n =\hat{P}_n$
, where
To highlight the outcome of heterogeneous beliefs, let us further assume there are only two agents (
$m=2$
) with constant risk aversion
$\Gamma^{(i)}_n = \gamma_i>0$
and with no endowment (
$H=0$
). In this case,
$\nabla g^{(i)}(0,\mathbb{Q}) = 0$
, and so, if the two agents have a common belief
$\hat{\mathbb{P}}$
, we need
$\hat{\mathbb{P}}=\mathbb{Q}$
for the market to be in general equilibrium. The optimal strategies are simply
$\pi^{(1)}_n = \pi^{(2)}_n = 0$
. On the other hand, for any
$\Delta_d^\circ$
-valued deterministic sequence
$\{\hat{P}^{(1)}_n\}$
, by choosing
$\hat{P}^{(2)}_n$
as
\begin{equation*}\hat{P}^{(2)}_{n,j} = \frac{(\hat{P}^{(1)}_{n,j})^{-\gamma_2/\gamma_1}}{\sum_{k=0}^d (\hat{P}^{(1)}_{n,k})^{-\gamma_2/\gamma_1}},\end{equation*}
we have
$\tilde{P}_{n,j} = 1/(d+1)$
for all n and j, which makes this market with heterogeneous beliefs be in general equilibrium. The individual optimal strategies
$\pi^{(1)}_n = - \pi^{(2)}_n$
are non-zero; the agents bet on their beliefs.
Another observation is that the equilibrium return is mostly affected by the belief of the least risk-averse agent. Indeed, if
$\Gamma^{(1)}_n \ll \Gamma^{(i)}_n$
for
$i \geq 2$
, we have
$\Gamma_n/\Gamma_n^{(1)} \approx 1$
, while
$\Gamma_n/\Gamma_n^{(i)} \approx 0$
for
$i \geq 2$
. Therefore
$\tilde{P}_n \approx \hat{P}^{(1)}_n$
.
Remark 3.4. The product
$\prod_n B_n$
corresponds to the consensus characteristic introduced in a continuous-time framework [Reference Jouini and Napp26] of heterogeneous beliefs. It can be interpreted as a discounting factor, and was further investigated in [Reference Hara21].
Appendix A. Computation of sup-convolution
Lemma A.1. Let
$\alpha \gt 0$
,
$\beta \gt 0$
,
$p_j > 0$
,
$q_j > 0$
,
$j=0,\ldots,d$
, and
$z \in \mathbb{R}^d$
. Then
\begin{align*} & \sup_{x \in \mathbb{R}^d}\Biggl\{-\frac{1}{\alpha}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\alpha x^\top v_j}p_j-\frac{1}{\beta}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\beta (z-x)^\top v_j}q_j\Biggr\}\\* &\quad = -\frac{1}{\gamma} \log \sum_{j=0}^d \,{\mathrm{e}}^{-\gamma z^\top v_j} p_j^{\gamma/\alpha}q_j^{\gamma/\beta},\end{align*}
where
$\gamma = 1/(1/\alpha + 1/\beta)$
.
Proof. The first-order condition is
\begin{equation*} \sum_{j=0}^d v_j \biggl(\frac{{\mathrm{e}}^{-\alpha x^\top v_j} p_j}{S(x,\alpha,\{p_k\})} - \frac{{\mathrm{e}}^{-\beta (z-x)^\top v_j}q_j}{S(z-x,\beta,\{q_k\})}\biggr) = 0,\end{equation*}
where
\begin{equation*} S(x,\alpha,\{p_k\}) = \sum_{j=0}^d \,{\mathrm{e}}^{-\alpha x^\top v_j}p_j.\end{equation*}
Since
$\mathbf{v}\mathbf{1}= 0$
and
$\mathrm{rank}\, \mathbf{v} = d$
, the first-order condition is met if and only if
for a function c. Substituting
we obtain
\begin{align*} -\frac{1}{\alpha}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\alpha x^\top v_j}p_j & = -\frac{1}{\alpha}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\gamma z^\top v_j}p_j^{\gamma/\alpha}q_j^{\gamma/\beta} - \frac{c(x)}{\alpha + \beta}, \\*-\frac{1}{\beta}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\beta (z-x)^\top v_j}q_j& = -\frac{1}{\beta}\log \sum_{j=0}^d \,{\mathrm{e}}^{-\gamma z^\top v_j}p_j^{\gamma/\alpha}q_j^{\gamma/\beta} + \frac{c(x)}{\alpha + \beta},\end{align*}
hence the result.
Lemma A.2. Let
$(p_{i,0},\ldots,p_{i,d})^\top$
,
$i=1,\ldots,m$
be m points in
$\Delta_d^\circ$
. Let
$\gamma_i \gt 0$
for
$i=1,\ldots,m$
and
\begin{equation*} \gamma = \Biggl(\sum_{i=1}^m \frac{1}{\gamma_i}\Biggr)^{-1}. \end{equation*}
Then
\begin{equation*} \sum_{j=0}^d \prod_{i=1}^m p_{i,j}^{\gamma/\gamma_i} \leq 1. \end{equation*}
Proof. The case
$m=1$
is trivial. Let
\begin{equation*} \hat\gamma_k = \Biggl(\sum_{i=1}^k \frac{1}{\gamma_i}\Biggr)^{-1}, \quad k=1,\ldots,m. \end{equation*}
If the inequality is true when
$m=k$
, then
\begin{equation*} \sum_{j=0}^d \prod_{i=1}^{k+1} p_{i,j}^{\hat\gamma_{k+1}/\gamma_i} = \sum_{j=0}^d p_{k+1,j} \biggl(\frac{\prod_{i=1}^{k} p_{i,j}^{\hat\gamma_k/\gamma_i} }{p_{k+1,j}}\biggr)^{\hat\gamma_{k+1}/\hat\gamma_k} \leq \Biggl( \sum_{j=0}^d \prod_{i=1}^{k} p_{i,j}^{\hat\gamma_{k}/\gamma_i} \Biggr)^{\hat\gamma_{k+1}/\hat\gamma_k} \leq 1\end{equation*}
by Jensen’s inequality. We obtain the result by induction.
Appendix B. Numerical experiment
Here we give a brief numerical experiment on the locally entropic monetary utility defined by (2.12). We focus on the case
$B_n=1$
and
$\hat{P}_{n,j} = 1/(d+1)$
for all n and j. The purpose here is to examine numerically the effect of the local risk-aversion process
$\{\Gamma_n\}$
. More specifically, we consider an auto-regressive structure
with constants
$\gamma \gt 0$
,
$\alpha \in [0,1]$
and
$\beta \in \mathbb{R}$
, and compute the g-expectation
$\mathcal{E}^g_0(Y_N)$
for
$Y_N = (Nd)^{-1/2}\mathbf{1}^\top X_N$
. When
$\beta$
is negative, the negative values of
$\mathbf{1}^\top \Delta X_n$
push
$\Gamma_n$
up, that is, the utility becomes more risk-averse when
$\mathbf{1}^\top \Delta X_n$
is negative. We are interested in how such dynamics affects the initial utility value
$Y_0 = \mathcal{E}^g_0(Y_N)$
.
For example, for
$d=100$
,
$N=3$
,
$\gamma=1$
, and
$\Sigma/(d+1)$
being the identity matrix, Figure 1 shows the shapes of
$Y_0=\mathcal{E}^g_0(Y_N)$
as a function of
$\beta \in ({-}0.1,0.1)$
for
$\alpha = 0$
,
$0.5$
, and 1. For the negative region of
$\beta$
, we have monotone shapes, which means that the larger the variance of the process
$\{\Gamma_n\}$
, the less the initial utility. This monotonicity is lost in the positive region of
$\beta$
.

Figure 1.
$Y_0=\mathcal{E}^g_0(Y_N)$
as a function of
$\beta \in ({-}0.1, 0.1)$
for
$\alpha = 0, 0.5$
and 1.
Acknowledgements
The authors are grateful to Professor Chiaki Hara for his helpful comments.
Funding information
Jun Sekine acknowledges Grant-in-Aid for Scientific Research (C), 23K01450.
Competing interests
There were no competing interests to declare which arose during the preparation or publication process of this article.







