PDF

DISS. ETH NO. 19272
OPTIMAL CONSUMPTION AND INVESTMENT WITH
POWER UTILITY
A dissertation submitted to
ETH ZURICH
for the degree of
Doctor of Sciences
presented by
MARCEL FABIAN NUTZ
Dipl. Math. ETH
born October 2, 1982
citizen of Basel BS, Switzerland
accepted on the recommendation of
Prof. Dr. Martin Schweizer
examiner
Prof. Dr. Huyên Pham
co-examiner
Prof. Dr. H. Mete Soner
co-examiner
Prof. Dr. Nizar Touzi
co-examiner
2010
ii
To Laura, whom I love.
iv
Abstract
In this thesis we study the utility maximization problem for power utility
random elds in a general semimartingale nancial market, with and without
intermediate consumption.
The notion of an opportunity process is intro-
duced as a reduced form of the value process for the resulting stochastic
control problem. This process is shown to describe the key objects: the optimal strategy, the value function, and the convex-dual problem. We show
that the existence of an optimal strategy implies that the opportunity process solves the so-called Bellman equation. The optimal strategy is described
pointwise in terms of the opportunity process, which is also characterized as
the minimal solution of the Bellman equation.
Furthermore, we provide
verication theorems for this equation. As an example, we consider exponential Lévy models, for which we construct an explicit solution in terms of
the Lévy triplet. Finally, we study the asymptotic properties of the optimal
strategy as the relative risk aversion tends to innity or to one. The convergence of the optimal consumption is obtained for the general case, while
the convergence of the optimal trading strategy is obtained for continuous
semimartingale models.
vi
Kurzfassung
Diese Dissertation beschäftigt sich mit dem Nutzenmaximierungs-Problem
für Potenznutzen, mit oder ohne Konsum, in einem allgemeinen Semimartingalmodell eines Finanzmarktes. Der so genannte Opportunitätsprozess wird
eingeführt als reduzierte Form des Wertprozesses des zugehörigen stochastischen Kontrollproblems. Dieser Prozess beschreibt die fundamentalen Grössen:
die optimale Strategie, die Wertfunktion und das konvex-duale Problem. Der
Opportunitätsprozess erfüllt die zugehörige Bellman-Gleichung, sobald eine
optimale Strategie existiert.
Umgekehrt wird dieser Prozess als die min-
imale Lösung dieser Gleichung charakterisiert, und die optimale Strategie
wird punktweise mit Hilfe dieses Prozesses beschrieben. Desweiteren zeigen
wir verschiedene Verikationstheoreme für die Bellman-Gleichung. Für den
Spezialfall eines exponentiellen Lévy-Modells leiten wir die explizite Lösung
des Problems unter minimalen Annahmen her. Schliesslich untersuchen wir
das asymptotische Verhalten der optimalen Strategien, wenn die relative
Risikoaversion gegen unendlich oder gegen eins strebt. Wir zeigen die Konvergenz des optimalen Konsums im allgemeinen Fall und die Konvergenz der
optimalen Handelsstrategie für stetige Semimartingalmodelle.
viii
Acknowledgements
It is my pleasure to thank Martin Schweizer for supervising this thesis and
Huyên Pham, Mete Soner and Nizar Touzi for readily accepting to act as
co-examiners.
During my time as a doctoral student, I greatly beneted
from discussions with Christoph Czichowsky, Freddy Delbaen, Christoph
Frei, Kostas Kardaras, Semyon Malamud, Johannes Muhle-Karbe, Alexandre Roch, Martin Schweizer, Mete Soner, Josef Teichmann, Nick Westray,
Johannes Wissel and Gordan ยšitkovi¢. Furthermore, I am indebted to Peter
Imkeller, Huyên Pham and Walter Schachermayer, who hosted stays at their
respective universities. I thank my oces mates and the entire Group 3 of
the Mathematics Department at ETH for creating such an enjoyable and inspiring atmosphere. I am deeply grateful to my family and friends for their
support and, last but not least, to Laura, for all the love.
Financial support by Swiss National Science Foundation Grant PDFM2120424/1 is gratefully acknowledged.
x
Contents
I Introduction
1
I.1
Power Utility Maximization
I.2
Classication of the Literature
. . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . .
2
I.3
Overview of the Thesis . . . . . . . . . . . . . . . . . . . . . .
3
II Opportunity Process
1
7
II.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
II.2
The Optimization Problem
II.3
The Opportunity Process
II.4
Relation to the Dual Problem . . . . . . . . . . . . . . . . . .
14
II.5
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . .
20
II.6
Appendix A: Dynamic Programming . . . . . . . . . . . . . .
23
II.7
Appendix B: Martingale Property of the Optimal Processes
25
. . . . . . . . . . . . . . . . . . .
8
. . . . . . . . . . . . . . . . . . . .
11
.
III Bellman Equation
27
III.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
III.2 Preliminaries
7
27
. . . . . . . . . . . . . . . . . . . . . . . . . . .
28
III.3 The Bellman Equation . . . . . . . . . . . . . . . . . . . . . .
35
III.4 Minimality of the Opportunity Process . . . . . . . . . . . . .
43
III.5 Verication
47
. . . . . . . . . . . . . . . . . . . . . . . . . . . .
III.6 Appendix A: Proof of Lemma 3.8 . . . . . . . . . . . . . . . .
57
III.7 Appendix B: Parametrization by Representative Portfolios . .
63
III.8 Appendix C: Omitted Calculation . . . . . . . . . . . . . . . .
66
IV Lévy Models
69
IV.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
IV.2 Preliminaries
69
. . . . . . . . . . . . . . . . . . . . . . . . . . .
70
IV.3 Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . .
74
IV.4 Transformation of the Model
. . . . . . . . . . . . . . . . . .
77
IV.5 Proof of Theorem 3.2 . . . . . . . . . . . . . . . . . . . . . . .
80
IV.6
๐‘ž -Optimal
Martingale Measures . . . . . . . . . . . . . . . . .
84
IV.7 Extensions to Non-Convex Constraints . . . . . . . . . . . . .
86
IV.8 Related Literature
88
. . . . . . . . . . . . . . . . . . . . . . . .
xii
Contents
V Risk Aversion Asymptotics
89
V.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
V.2
Preliminaries
. . . . . . . . . . . . . . . . . . . . . . . . . . .
90
V.3
Main Results
. . . . . . . . . . . . . . . . . . . . . . . . . . .
94
V.4
Tools and Ideas for the Proofs . . . . . . . . . . . . . . . . . .
99
V.5
Auxiliary Results . . . . . . . . . . . . . . . . . . . . . . . . . 102
V.6
The Limit
V.7
The Limit
V.8
Appendix A: Convergence of Stochastic Exponentials . . . . . 126
V.9
Appendix B: Proof of Lemma 6.12
๐‘ โ†’ โˆ’โˆž .
๐‘โ†’0 . .
89
. . . . . . . . . . . . . . . . . . . . . . . 108
. . . . . . . . . . . . . . . . . . . . . . . 117
. . . . . . . . . . . . . . . 128
Chapter I
Introduction
This chapter describes the optimization problem at hand, embeds the thesis
in the literature and gives an overview of the main results.
I.1 Power Utility Maximization
Expected utility criteria are used to describe the preferences of a rational
economic agent.
We shall start with a given utility function
๐‘ˆ
and refer
to Föllmer and Schied [22] for a general introduction to the modeling of
preferences and the relations to economic axioms for the latter.
We consider an agent who begins with an initial endowment
๐‘ฅ0 > 0
and
invests in a nancial market in continuous time. Her preferences are modeled
by an increasing and concave function
๐‘ˆ
and her aim is to maximize a certain
utility functional. We shall consider two problems of this type. The rst one
is the maximization of expected utility from terminal wealth, i.e, the agent
chooses her trading strategy
๐œ‹
such as to maximize the expectation
[ (
)]
๐ธ ๐‘ˆ ๐‘‹๐‘‡ (๐œ‹) ,
๐‘‹๐‘‡ (๐œ‹) denotes the wealth resulting from ๐‘ฅ0 and ๐œ‹ at a given time
๐‘‡ . In the second problem the agent is also allowed to consume
during the interval [0, ๐‘‡ ]; i.e., she chooses a trading strategy ๐œ‹ as well as a
consumption rate ๐‘ with the aim of maximizing
]
[โˆซ ๐‘‡
(
)
หœ
๐ธ
๐‘ˆ (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ ๐‘‹๐‘‡ (๐œ‹, ๐‘) ,
where
horizon
0
where
หœ
๐‘ˆ
is again some utility function. We shall treat both of these speci-
cations in a unied notation, or more precisely, a slight extension involving a
time-dependent (and possibly random) utility function
๐‘ˆ๐‘ก (โ‹…).
For simplicity,
we restrict our attention to classical utility functions in this introduction. As
usual, the problem is simplied by assuming that the agent is small in the
sense that her actions do not inuence the nancial market. Moreover, the
I Introduction
2
market is supposed to be frictionless, i.e., free of transaction costs, liquidity
eects, and so on.
๐‘ˆ (๐‘ฅ) (de๐‘ฅ > 0) are those for which the relative risk aversion โˆ’๐‘ฅ๐‘ˆ โ€ฒโ€ฒ (๐‘ฅ)/๐‘ˆ โ€ฒ (๐‘ฅ)
The most frequently used explicit examples of utility functions
ned for
is constant. This choice is a compromise between tractability and generality, and also due to the lack of other canonical examples. The logarithmic
utility corresponds to unit relative risk aversion and yields the most explicit
1
results as it leads to a myopic behavior ; but this also means that many
interesting properties of general utility functions cannot be observed there.
(0, โˆž) โˆ– {1} are called
๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1). They
The functions with constant relative risk aversion in
power utilities
and of the form
๐‘ˆ (๐‘ฅ) = ๐‘1 ๐‘ฅ๐‘
with
entail non-myopic phenomena while their scaling properties still lead to a
substantial simplication: in a suitable parametrization, the optimal strategies do not depend on the level of wealth. The frequent use of power utilities
motivates their study in a general nancial market model, and this is the
content of the present thesis.
I.2 Classication of the Literature
There is a vast literature on the maximization of expected utility and we
conne ourselves to a brief classication of the approaches.
The book of
Karatzas and Shreve [42, pp. 153] contains a survey of the literature up to
its date of writing.
Existence.
The existence of an optimal strategy in a frictionless semi-
martingale nancial market has been established in satisfactory generality
using martingale theory and convex duality.
This works for general util-
ity functions; see Kramkov and Schachermayer [49] for the case of terminal
wealth (on
โ„+ ) and Karatzas and ยšitkovi¢ [43] for the case with consumption
(and random endowment).
In the light of these results, the focus of mathematical research in (singleagent) expected utility maximization has shifted to the study of the properties of the optimal strategy as well as to markets with friction. In the sequel,
we focus on the rst aspect and power utility.
Explicit Solutions.
The most evident approach is to explicitly solve the
utility maximization problem for suitable market models. In the case with
consumption this turns out to be dicult: except for complete markets, an
explicit solution can be expected only for Lévy models (see Chapter IV).
In the case of terminal wealth, certain algebraic properties of a market
1
No doubt some will say: `I'm not sure of my taste for risk. I lack a rule to act on.
So I grasp at one that at least ends doubt: better to act to make the odds big that I win
than to be left in doubt?' Not so. There is more than one rule to end doubt. Why pick
on one odd one? (from Samuelson's comment [66] on logarithmic utilitya paper whose
most distinctive feature is to consist of words of one syllable.)
I.3 Overview of the Thesis
model can lead to an explicit solution also in the incomplete case.
3
The
exponentially-quadratic model of Kim and Omberg [47] was among the rst
of this kind; other examples include certain exponentially-ane specications as in Kallsen and Muhle-Karbe [40] and Muhle-Karbe [58].
Markovian Dynamic Programming and PDEs.
When the asset prices
follow a Markov process, dynamic programming is often used to show that
the value function satises the corresponding Hamilton-Jacobi-Bellman partial dierential equation (PDE) in the viscosity sense (or integro-PDE in the
case with jumps). In turn, the value function can be used to describe the optimal strategyat least formally, in the sense that the expression at hand involves the derivatives of the value function while its regularity is known only
under strong assumptions on the model. We refer to Fleming and Soner [21]
for background and general references. Of course, the important early work
of Merton [55, 56] falls in this category.
Stoikov and Zariphopoulou [72]
study optimal consumption in a diusion model with constant correlation;
we extend some of their results to semimartingale models in Chapter II.
Non-Markovian Dynamic Programming and BSDEs.
In a suitable
formulation, dynamic programming can be applied also when the asset prices
are not Markovian, and this is the approach taken in this thesis. In contrast
to the Markov case, the corresponding local equation is stochastic.
In
Chapter III we shall state it as a backward stochastic dierential equation
(BSDE) in the context of a general semimartingale model. For the case of
terminal wealth, this was previously obtained for certain continuous models
by Mania and Tevzadze [54] and Hu et al. [33]. It is worth noting that the
BSDE formulation does not extend to general utility functions, whereas in
the Markov case, the corresponding PDE (or integro-PDE) is standard.
I.3 Overview of the Thesis
In view of the general existence results mentioned above, the problem at the
center of the thesis is
model.
the description of the optimal strategy in a general
The results are divided into four chapters which correspond to the
articles [61, 59, 60, 62]. The interdependencies are as follows:
I Introduction
II Opportunity Process
III Bellman Equation
iiii
iiii
i
i
i
i
i
t iii
IV Lévy Models
UUUU
UUUU
UUUU
UUUU
U*
V Risk Aversion Asymptotics
4
I Introduction
However, each part is written in a self-contained way, i.e., necessary results
from other chapters are always recalled.
The following paragraphs give a
synthesis of each part.
Opportunity Process.
The concept of
dynamic programming
plays an
important role in this thesis. Its fundamental quantity is the value function,
i.e., the maximal expected utility that the agent can achieve, given certain
initial conditions.
๐œ‹
In a non-Markovian context, this means that we freeze
๐‘ก in time. Then
we consider the maximal (conditional) expected utility ๐ฝ๐‘ก (๐œ‹) that can be
achieved by optimizing the strategy on the remaining time interval [๐‘ก, ๐‘‡ ].
The stochastic process ๐‘ก 7โ†’ ๐ฝ๐‘ก (๐œ‹) is called the value process corresponding
to ๐œ‹ .
some strategy
(resp.
(๐œ‹, ๐‘))
to be used until some point
The scaling property of our power utility functional leads to a factorization of the value process into one part which depends on the current wealth,
and a process
๐ฟ.
It is called
opportunity process
as
๐ฟ๐‘ก
encodes the maxi-
mal conditional expected utility which can be attained from time
๐‘ก,
starting
from one unit of endowment. The factorization itself is very classicalfor
instance, it can already be found in Merton's workand an opportunity process is present (in a more or less explicit form) in almost any paper dealing
with so-called isoelastic utility functions.
general study of
๐ฟ
for power utility.
However, there was thus far no
We shall not indicate all the related
literature but merely mention that the name opportunity process was introduced by ยƒerný and Kallsen [11] for an analogous object in the context
of mean-variance hedging.
In Chapter II we rst introduce rigorously the opportunity process and
then proceed to establish the connection to the dual problem in the sense
of convex duality. On the one hand, the dual problem has a scaling property
similar to the one of the (primal) utility maximization problem and this
gives rise to a dual opportunity process denoted by
๐ฟโˆ— is simply a power of
๐ฟ.
๐ฟโˆ— .
It turns out that
This allows us to relate uniform bounds for
๐ฟ
to so-called reverse Hölder inequalities for processes belonging to the dual
domain. On the other hand, the optimal supermartingale solving the dual
problem is expressed via
๐ฟ
and the optimal wealth process.
In the case
with consumption, it is known that this supermartingale coincides with the
marginal utility of the optimal consumption rate
feedback formula for
๐‘ห† in
terms of
๐ฟ.
๐‘ห†.
Therefore, we obtain a
We exploit this connection to obtain
model-independent bounds for the optimal consumption and to study how
it is aected by certain changes in the model.
Bellman Equation.
In contrast to the optimal consumption, nothing is
said in Chapter II about the
trading strategy,
which forms the second part
of the optimal strategy. Its description via the opportunity process requires
a more involved stochastic analysis approach, which is the content of Chapter III. We present a local representation of the optimization problem that
I.3 Overview of the Thesis
we call
Bellman equation
5
in analogy to classical Markovian control prob-
lems. Its main ingredient is a random function which is dened in terms of
the semimartingale characteristics of the asset prices and the opportunity
process. Optimal trading strategies are characterized as maximizers for this
function. We also recover, with a quite dierent proof, the feedback formula
for the optimal consumption.
The Bellman equation is stated in two forms: rst as an equation of differential semimartingale characteristics and then as a backward stochastic
dierential equation (BSDE). The main result is that whenever an optimal strategy exists, the opportunity process (resp. the joint characteristics
with the price process) solves the Bellman equation. Our construction relies purely on dynamic programming and necessitates neither additional noarbitrage assumptions nor duality theory. This allows us to formulate the
problem under portfolio constraints that need not be convex.
We can see the Bellman equation as a description for the opportunity
process. In certain models the equation can be solved directly using existence
results for BSDEs with quadratic growth. However, a verication result is
needed to show that a solution of the equation corresponds to the solution
of our optimization problem.
We provide sucient (and also necessary)
conditions for this to hold. This is also a rst answer to the question whether
the opportunity process is fully described by the Bellman equation. While
there are no general uniqueness results for BSDEs driven by semimartingales,
we show that the opportunity process can be characterized as the minimal
solution of the Bellman equation.
Lévy Models.
In Chapter IV we consider the special case when the asset
prices follow an exponential Lévy process. In the continuous case this corresponds to a drifted geometric Brownian motion, which is the specication
in the original Merton problem.
A classical observation in various models
is that when the asset returns are i.i.d., the optimal portfolio and consumption are given by a constant and a deterministic function, respectively, in a
suitable parametrization. The aim of Chapter IV is to establish this fact for
convex-constrained Lévy models under minimal assumptions.
The optimal portfolio is characterized as the maximizer of a
function
๐”ค
๐”ค
deterministic
dened in terms of the Lévy triplet; and the maximum value of
yields the optimal consumption. The function
๐”ค
is closely related to the
random function mentioned above. While it is clear that the niteness of the
value function is a necessary requirement to study the utility maximization
problem, it is in general impossible to describe this condition directly in terms
of the model primitives. In the present special case, we succeed to state a
description in terms of the Lévy triplet.
We also consider the
๐‘ž -optimal
equivalent martingale measures that are linked to utility maximization by
convex duality (๐‘ž
โˆˆ (โˆ’โˆž, 1) โˆ– {0});
this results in an explicit existence
characterization and a formula for the density process.
Finally, we study
6
I Introduction
some generalizations to non-convex constraints.
The approach in Chapter IV is classical and consists in solving the Bellman equation, which reduces to an ordinary dierential equation in the Lévy
setting. The main diculty is to construct the maximizer for
๐”ค;
once this is
achieved, we can apply the general verication results from Chapter III. The
necessary compactness is obtained from a minimal no-free-lunch condition
via scaling arguments which were developed by Kardaras [44] for
log-utility.
In our case, these arguments require certain additional integrability properties of the asset returns. Without compromising the generality, integrability
is achieved by a transformation which replaces the given assets by certain
portfolios.
In fact, these portfolios are special cases of the representative
portfolios that are introduced in the appendix of Chapter III to clarify certain technical issues (which we shall not detail in this overview).
Risk Aversion Asymptotics.
(๐‘) (๐‘ฅ)
utility function ๐‘ˆ
In Chapter V we vary
optimal strategies
๐‘
=
Thus far, we have considered the power
1 ๐‘
๐‘ ๐‘ฅ for a xed parameter
๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1).
and our main interest concerns the behavior of the
in the limits ๐‘ โ†’ โˆ’โˆž and ๐‘ โ†’ 0.
The relative risk aversion of
๐‘ˆ (๐‘)
tends to innity for
๐‘ โ†’ โˆ’โˆž,
hence
we guess by the economic interpretation of this quantity that the optimal
investment portfolio tends to zero.
If there is no trading, optimizing the
consumption becomes a deterministic problem that is readily solved.
We
prove (in a general semimartingale model) that the optimal consumption,
expressed as a proportion of wealth, converges pointwise to a deterministic
function.
This function corresponds to the consumption which would be
optimal in the case where trading is not allowed. In the continuous semimartingale case, we show that the optimal trading strategy tends to zero.
Our second result pertains to the same limit
๐‘ โ†’ โˆ’โˆž but concerns utility
from terminal wealth only. It can be seen as a rst-order asymptotic: in the
continuous case, we show that the optimal trading strategy scaled by
1โˆ’๐‘
converges to a strategy which is optimal for exponential utility.
For the limit
๐‘ โ†’ 0,
we note that
๐‘ = 0
formally corresponds to the
logarithmic utility function. Again, we establish the convergence of the corresponding optimal consumption in the general case, and the convergence of
the trading strategy in the continuous case.
In view of the feedback formula for the optimal consumption mentioned
before, we study the dependence of the (primal and dual) opportunity processes on
๐‘ and their convergence.
This uses control-theoretic arguments and
convex analysis. To obtain the convergence of the strategies, we study the
asymptotics of the Bellman equation. The continuity assumption simplies
the equation and renders the optimal portfolio relatively explicit (in terms
of the Kunita-Watanabe decomposition of the opportunity process). Despite
this, we are not in a standard framework for quadratic BSDEs, and therefore
we give the proofs by direct arguments.
Chapter II
Opportunity Process
In this chapter, which corresponds to the article [61], we lay the foundations
for the rest of the thesis. We present the basic dynamic programming for
the power utility maximization problem and we introduce the opportunity
process. We also give applications to the study of the optimal consumption
strategy.
II.1 Introduction
We consider the utility maximization problem in a semimartingale model for
a nancial market, with and without intermediate consumption. While the
model is general, we focus on
power utilities.
If the maximization is seen as
a stochastic control problem, the power form leads to a factorization of the
value process into a part which depends on the current wealth and a process
๐ฟ
around which our analysis is built. It is called
opportunity process
as
๐ฟ๐‘ก
encodes the maximal conditional expected utility that can be attained from
time ๐‘ก. This name was introduced by ยƒerný and Kallsen [11] for an analogous
object in the context of mean-variance hedging. Surprisingly, there exists no
general study of
๐ฟ
for the case of power utility, which is a gap we try to ll
here.
The opportunity process is a suitable tool to derive
qualitative
results
about the optimal consumption strategy. We present monotonicity properties and bounds which are quite explicit despite the generality of the model.
This chapter is organized as follows. After the introduction, we specify
the optimization problem in detail. Section II.3 introduces the opportunity
process
๐ฟ
via dynamic programming and examines its basic properties. Sec-
tion II.4 relates
๐ฟ
to convex duality theory and reverse Hölder inequalities,
which is useful to obtain bounds for the opportunity process. Section II.5
gives applications to the study of the optimal consumption.
a feedback formula in terms of
๐ฟ
We establish
and use it to study how certain changes
in the model aect the optimal consumption. These applications illustrate
II Opportunity Process
8
the usefulness of the opportunity process: they are general but have very
simple proofs.
Two appendices supply facts about dynamic programming
and duality theory.
We refer to Jacod and Shiryaev [34] for unexplained notation.
II.2 The Optimization Problem
Financial Market.
probability space
We x the time horizon
(ฮฉ, โ„ฑ, (โ„ฑ๐‘ก )๐‘กโˆˆ[0,๐‘‡ ] , ๐‘ƒ )
๐‘‡ โˆˆ (0, โˆž)
and a ltered
satisfying the usual assumptions of
โ„ฑ0 = {โˆ…, ฮฉ} ๐‘ƒ -a.s. We consider
๐‘… with ๐‘…0 = 0. The (componentwise)
right-continuity and completeness, as well as
๐‘‘
an โ„ -valued càdlàg semimartingale
๐‘† = โ„ฐ(๐‘…) represents the discounted price processes
๐‘‘ risky assets, while ๐‘… stands for their returns. Our agent also has a bank
stochastic exponential
of
account paying zero interest at his disposal.
Trading Strategies and Consumption. The agent is endowed with a
deterministic initial capital ๐‘ฅ0 > 0. A trading strategy is a predictable ๐‘…integrable
โ„๐‘‘ -valued
process
๐œ‹,
where the
๐‘–th
component is interpreted as
the fraction of wealth (or the portfolio proportion) invested in the
asset. A
โˆซ๐‘‡
0
consumption strategy
๐‘๐‘ก ๐‘‘๐‘ก < โˆž ๐‘ƒ -a.s.
is a nonnegative optional process
We want to consider two cases.
occurs only at the terminal time
๐‘‡
๐‘
๐‘–th
risky
such that
Either consumption
(utility from terminal wealth only);
or there is intermediate consumption plus a bulk consumption at the time
horizon. To unify the notation, dene the measure
{
0
๐œ‡(๐‘‘๐‘ก) :=
๐‘‘๐‘ก
๐œ‡
on
[0, ๐‘‡ ]
by
in the case without intermediate consumption,
in the case with intermediate consumption.
๐œ‡โˆ˜ := ๐œ‡ + ๐›ฟ{๐‘‡ } , where ๐›ฟ{๐‘‡ } is the unit Dirac measure at ๐‘‡ .
process ๐‘‹(๐œ‹, ๐‘) corresponding to a pair (๐œ‹, ๐‘) is described by the
We also dene
The
wealth
linear equation
โˆซ
๐‘ก
โˆซ
๐‘‹๐‘ โˆ’ (๐œ‹, ๐‘)๐œ‹๐‘  ๐‘‘๐‘…๐‘  โˆ’
๐‘‹๐‘ก (๐œ‹, ๐‘) = ๐‘ฅ0 +
0
and the set of
admissible
๐‘ก
๐‘๐‘  ๐œ‡(๐‘‘๐‘ ),
0โ‰ค๐‘กโ‰ค๐‘‡
(2.1)
0
trading and consumption pairs is
{
๐’œ(๐‘ฅ0 ) = (๐œ‹, ๐‘) : ๐‘‹(๐œ‹, ๐‘) > 0, ๐‘‹โˆ’ (๐œ‹, ๐‘) > 0
and
}
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘) .
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘) means that all the remaining wealth is consumed at time ๐‘‡ ; it is merely for notational convenience. Indeed, ๐‘‹(๐œ‹, ๐‘)
does not depend on ๐‘๐‘‡ , hence any given consumption strategy ๐‘ can be redened to satisfy ๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘). We x the initial capital ๐‘ฅ0 and usually write
๐’œ for ๐’œ(๐‘ฅ0 ). A consumption strategy ๐‘ is called admissible if there exists ๐œ‹
The convention
II.2 The Optimization Problem
such that
(๐œ‹, ๐‘) โˆˆ ๐’œ;
we write
๐‘โˆˆ๐’œ
for brevity. The meaning of
๐œ‹โˆˆ๐’œ
9
is
analogous.
Sometimes it is convenient to parametrize the consumption strategies as
fractions of wealth. Let
(๐œ‹, ๐‘) โˆˆ ๐’œ and let ๐‘‹ = ๐‘‹(๐œ‹, ๐‘) be the corresponding
wealth process. Then
๐œ… :=
is called the
๐‘
๐‘‹
(2.2)
propensity to consume corresponding to (๐œ‹, ๐‘).
due to our convention that
Remark 2.1.
Note that
๐œ…๐‘‡ = 1
๐‘๐‘‡ = ๐‘‹๐‘‡ .
(i) The parametrization
(๐œ‹, ๐œ…)
allows to express wealth pro-
cesses as stochastic exponentials: by (2.1),
(
)
๐‘‹(๐œ‹, ๐œ…) = ๐‘ฅ0 โ„ฐ ๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡
(2.3)
๐‘‹(๐œ‹, ๐‘) for ๐œ… := ๐‘/๐‘‹(๐œ‹, ๐‘), where we have used that ๐‘‹(๐œ‹, ๐‘) =
๐‘‹(๐œ‹, ๐‘)โˆ’โˆซ๐œ‡-a.e. because it is càdlàg. The symbol โˆ™ indicates an integral, e.g.,
๐œ‹ โˆ™ ๐‘… = ๐œ‹๐‘  ๐‘‘๐‘…๐‘  .
coincides with
(ii) Relation (2.2) induces a one-to-one correspondence between the pairs
(๐œ‹, ๐‘) โˆˆ ๐’œ
(๐œ‹, ๐œ…) such that ๐œ‹ โˆˆ ๐’œ and ๐œ… is a nonnegative
โˆซ๐‘‡
optional process satisfying
0 ๐œ…๐‘  ๐‘‘๐‘  < โˆž ๐‘ƒ -a.s. and ๐œ…๐‘‡ = 1. Indeed, given
(๐œ‹, ๐‘) โˆˆ ๐’œ, dene ๐œ… by (2.2) with ๐‘‹ = ๐‘‹(๐œ‹, ๐‘). As ๐‘‹, ๐‘‹โˆ’ > 0 and as ๐‘‹
is càdlàg, almost every path of ๐‘‹ is bounded away from zero and ๐œ… has
the desired integrability. Conversely, given (๐œ‹, ๐œ…), dene ๐‘‹ via (2.3) and
๐‘ := ๐œ…๐‘‹ ; then ๐‘‹ = ๐‘‹(๐œ‹, ๐‘). From admissibility we deduce ๐œ‹ โŠค ฮ”๐‘… > โˆ’1 up
to evanescence, which in turn shows ๐‘‹ > 0. Now ๐‘‹โˆ’ > 0 by a standard
property of stochastic exponentials [34, II.8a], so (๐œ‹, ๐‘) โˆˆ ๐’œ.
and the pairs
Preferences.
[โˆซ
๐ธ
๐‘‡
0
๐ท๐‘ 
๐œ‡โˆ˜ (๐‘‘๐‘ )
]Let ๐ท be a càdlàg adapted strictly positive process such that
< โˆž and x ๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1). We dene the utility
random eld
๐‘ˆ๐‘ก (๐‘ฅ) := ๐ท๐‘ก ๐‘1 ๐‘ฅ๐‘ ,
where
1/0 := โˆž.
To wit, this is
๐‘ฅ โˆˆ [0, โˆž), ๐‘ก โˆˆ [0, ๐‘‡ ],
any ๐‘-homogeneous
utility random eld
such that a constant consumption yields nite expected utility. The positive
number
1โˆ’๐‘
is called the relative risk aversion of
assume that there are constants
0 < ๐‘˜1 โ‰ค ๐‘˜2 < โˆž
๐‘˜1 โ‰ค ๐ท๐‘ก โ‰ค ๐‘˜2 ,
The
expectedโˆซ utility
๐‘‡
0
๐‘ˆ.
Sometimes we shall
such that
๐‘ก โˆˆ [0, ๐‘‡ ].
(2.4)
๐‘ โˆˆ ๐’œ
๐ธ[๐‘ˆ๐‘‡ (๐‘๐‘‡ )]
corresponding to a consumption strategy
) ๐œ‡โˆ˜ (๐‘‘๐‘ก)].
is
๐ธ[ ๐‘ˆ๐‘ก (๐‘๐‘ก
We recall that this is either
or
โˆซ๐‘‡
๐ธ[ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ๐‘‡ (๐‘๐‘‡ )]. In the case without intermediate consumption, ๐‘ˆ๐‘ก
is irrelevant for ๐‘ก < ๐‘‡ . We remark that Zariphopoulou [74] and Tehranchi [73]
given by
have used utility functions modied by a multiplicative random variable, in
the case where utility is obtained from terminal wealth.
II Opportunity Process
10
Remark 2.2.
The process
๐ท
can be used for discounting utility and con-
sumption, or to determine the weight of intermediate consumption compared
to terminal wealth. Our utility functional can also be related to the usual
1 ๐‘
๐‘ ๐‘ฅ in the following ways. If we write
power utility function
๐ธ
๐‘‡
[โˆซ
๐‘‡
]
[โˆซ
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) = ๐ธ
1 ๐‘
๐‘ ๐‘๐‘ก ๐‘‘๐พ๐‘ก
0
0
]
๐ท๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ก), we have the usual power utility, but with a
๐‘‘๐พ๐‘ก :=
clock ๐พ (cf.
for
Goll and Kallsen [25]).
taxation
To model
๐ท := (1 +
stochastic
๐œš)โˆ’๐‘ . If
๐‘
of the consumption, let
๐œš > โˆ’1
be the tax rate and
represents the cashow out of the portfolio,
๐‘/(1 + ๐œš)
is the eectively obtained amount of the consumption good, yielding the
1
๐‘ (๐‘๐‘ก /(1
instantaneous utility
multiplicative
+ ๐œš๐‘ก ))๐‘ = ๐‘ˆ๐‘ก (๐‘๐‘ก ).
๐ท๐‘‡
Similarly,
bonus payment.
can model a
For yet another alternative, assume either that there is no intermediate
consumption or that
๐ธ
๐ท
๐‘‡
[โˆซ
๐ธ[๐ท๐‘‡ ] = 1.
is a martingale, and that
]
หœ
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) = ๐ธ ๐‘ƒ
[โˆซ
0
0
with the equivalent probability
๐‘ƒหœ
dened by
๐‘‡
1 ๐‘ โˆ˜
๐‘ ๐‘๐‘ก ๐œ‡ (๐‘‘๐‘ก)
๐‘ƒหœ
instead of the objective probability
]
๐‘‘๐‘ƒหœ = ๐ท๐‘‡ ๐‘‘๐‘ƒ .
standard power utility problem for an agent with
uses
Then
This is the
subjective beliefs, i.e., who
๐‘ƒ.
Of course, these applications can be combined in a multiplicative way.
We assume that the value of the utility maximization problem is nite:
๐‘ข(๐‘ฅ0 ) := sup ๐ธ
standing assumption for the entire chapter.
because then
๐‘ˆ < 0.
If
๐‘ > 0,
case basis (see also Remark 4.7).
๐ธ
[โˆซ๐‘‡
0
]
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) < โˆž.
๐‘ˆ๐‘ก (๐‘๐‘ก
) ๐œ‡โˆ˜ (๐‘‘๐‘ก)
]
= ๐‘ข(๐‘ฅ0 ).
to guarantee its existence.
measures for
๐‘†.
(2.5)
0
๐‘โˆˆ๐’œ(๐‘ฅ0 )
This is a
๐‘‡
[โˆซ
It is void if
๐‘<0
it needs to be checked on a case-byA strategy
(ห†
๐œ‹ , ๐‘ห†) โˆˆ ๐’œ(๐‘ฅ0 )
is
optimal
if
Of course, a no-arbitrage property is required
Let
M๐‘†
be the set of equivalent
๐œŽ -martingale
If
M ๐‘† โˆ•= โˆ…,
(2.6)
arbitrage is excluded in the sense of the NFLVR condition (see Delbaen and
Schachermayer [17]). We can cite the following existence result of Karatzas
and ยšitkovi¢ [43]; it was previously obtained by Kramkov and Schachermayer [49] for the case without intermediate consumption.
Proposition 2.3. Under (2.4) and (2.6), there exists an optimal strategy
ห† = ๐‘‹(ห†
(ห†
๐œ‹ , ๐‘ห†) โˆˆ ๐’œ. The corresponding wealth process ๐‘‹
๐œ‹ , ๐‘ห†) is unique. The
consumption strategy ๐‘ห† can be chosen to be càdlàg and is unique ๐‘ƒ โŠ— ๐œ‡โˆ˜ -a.e.
II.3 The Opportunity Process
11
๐‘ห† denotes a càdlàg version. We note that under (2.6), the
๐‘‹(๐œ‹, ๐‘)โˆ’ > 0 in the denition of ๐’œ is automatically satised
๐‘‹(๐œ‹, ๐‘) > 0, because ๐‘‹(๐œ‹, ๐‘) is then a positive supermartingale
In the sequel,
requirement
as soon as
under an equivalent measure.
Remark 2.4.
In Proposition 2.3, the assumption on
๐ท
can be weakened
by exploiting that (2.6) is invariant under equivalent changes of measure.
Suppose that
๐ท = ๐ทโ€ฒ ๐ทโ€ฒโ€ฒ ,
with unit expectation.
the probability
instead of
๐ท,
where
๐ทโ€ฒ
meets (2.4) and
๐ทโ€ฒโ€ฒ
is a martingale
As in Remark 2.2, we consider the problem under
๐‘‘๐‘ƒหœ = ๐ท๐‘‡โ€ฒโ€ฒ ๐‘‘๐‘ƒ ,
๐‘ƒหœ with ๐ทโ€ฒ
under ๐‘ƒ .
then Proposition 2.3 applies under
and we obtain the existence of a solution also
II.3 The Opportunity Process
This section introduces the main object under discussion.
We do not yet
impose the existence of an optimal strategy, but recall the standing assumption (2.5). To apply dynamic programming, we introduce for each
and
๐‘ก โˆˆ [0, ๐‘‡ ]
(๐œ‹, ๐‘) โˆˆ ๐’œ
the set
{
๐’œ(๐œ‹, ๐‘, ๐‘ก) = (หœ
๐œ‹ , ๐‘หœ) โˆˆ ๐’œ : (หœ
๐œ‹ , ๐‘หœ) = (๐œ‹, ๐‘)
on
}
[0, ๐‘ก] .
(3.1)
(๐‘ก, ๐‘‡ ] after having used (๐œ‹, ๐‘) until ๐‘ก. The
๐‘หœ โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก) means that there exists ๐œ‹
หœ such that (หœ
๐œ‹ , ๐‘หœ) โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก).
(๐œ‹, ๐‘) โˆˆ ๐’œ, we consider the value process
]
[โˆซ ๐‘‡
๐ฝ๐‘ก (๐œ‹, ๐‘) := ess sup ๐ธ
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก .
(3.2)
These are the controls available on
notation
Given
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
๐‘ก
We choose the càdlàg version of this process (see Proposition 6.2 in the
๐‘-homogeneity of the utility functional leads to the following
๐ฝ.
Appendix). The
factorization of
Proposition 3.1. There exists a unique càdlàg semimartingale ๐ฟ, called
opportunity process, such that
๐ฟ๐‘ก
1
๐‘
(
๐‘‹๐‘ก (๐œ‹, ๐‘)
)๐‘
= ๐ฝ๐‘ก (๐œ‹, ๐‘) = ess sup ๐ธ
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
[โˆซ
๐‘ก
๐‘‡
]
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
(3.3)
for any admissible strategy (๐œ‹, ๐‘) โˆˆ ๐’œ. In particular, ๐ฟ๐‘‡ = ๐ท๐‘‡ .
Proof.
ห‡ := ๐‘‹(ห‡
(๐œ‹, ๐‘), (ห‡
๐œ‹ , ๐‘ห‡) โˆˆ ๐’œ and ๐‘‹ := ๐‘‹(๐œ‹, ๐‘), ๐‘‹
๐œ‹ , ๐‘ห‡). We claim that
โˆซ
]
[ ๐‘‡
1
โˆ˜
ess
sup
๐ธ
๐‘ˆ
(หœ
๐‘
)
๐œ‡
(๐‘‘๐‘ )
(3.4)
โ„ฑ๐‘ก
๐‘  ๐‘ 
ห‡ ๐‘ ๐‘หœโˆˆ๐’œ(ห‡๐œ‹,ห‡๐‘,๐‘ก)
๐‘‹
๐‘ก
๐‘ก
]
[โˆซ ๐‘‡
1
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก .
= ๐‘ ess sup ๐ธ
๐‘‹๐‘ก ๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
๐‘ก
Let
II Opportunity Process
12
Indeed, using the lattice property given in Fact 6.1, we can nd a sequence
(๐‘๐‘› )
in
๐’œ(ห‡
๐œ‹ , ๐‘ห‡, ๐‘ก)
such that, with a monotone increasing limit,
]
[โˆซ ๐‘‡
๐‘‹๐‘ก๐‘
โˆ˜
๐‘ˆ
(หœ
๐‘
)
๐œ‡
(๐‘‘๐‘ )
ess
sup
๐ธ
โ„ฑ๐‘ก =
๐‘ 
๐‘ 
ห‡ ๐‘ ๐‘หœโˆˆ๐’œ(ห‡๐œ‹,ห‡๐‘,๐‘ก)
๐‘‹
๐‘ก
๐‘ก
]
[โˆซ ๐‘‡ (
) โˆ˜
๐‘‹๐‘ก ๐‘›
๐‘ˆ๐‘  ๐‘‹ห‡ ๐‘๐‘  ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก โ‰ค
= lim ๐ธ
๐‘›
๐‘ก
๐‘ก
[
๐‘‹๐‘ก๐‘
lim
๐ธ
ห‡๐‘ ๐‘›
๐‘‹
๐‘ก
ess sup ๐ธ
๐‘‡
โˆซ
]
๐‘ˆ๐‘  (๐‘๐‘›๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘ก
๐‘‡
[โˆซ
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
๐‘ก
]
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก ,
โˆ˜
where we have used Fact 6.3 in the last step. The claim follows by symmetry.
)๐‘ ]
๐‘‹
(๐œ‹,
๐‘)
, ๐ฟ does not depend on
๐‘ก
๐‘
choice of (๐œ‹, ๐‘) โˆˆ ๐’œ and inherits the properties of ๐ฝ(๐œ‹, ๐‘) and ๐‘‹(๐œ‹, ๐‘) > 0.
๐ฟ๐‘ก := ๐ฝ๐‘ก (๐œ‹, ๐‘)/
Thus, if we dene
[1(
the
The opportunity process describes (๐‘ times) the maximal amount of conditional expected utility that can be accumulated on
[๐‘ก, ๐‘‡ ]
from one unit of
wealth. Note that the value function (2.5) can be expressed as
๐‘ข(๐‘ฅ) = ๐ฟ0 ๐‘1 ๐‘ฅ๐‘ .
In a Markovian setting, the factorization of the value function (which
then replaces the value process) is very classical; for instance, it can already
be found in Merton [56]. Mania and Tevzadze [54] study power utility from
terminal wealth in a continuous semimartingale model; that paper contains
some of the basic notions used here as well.
Remark 3.2.
Let
Remark 3.3.
We can now formalize the fact that the optimal strategies (in
๐ท
๐ท0 = 1 and ๐‘ƒหœ as in Remark 2.2.
หœ
Bayes' rule and (3.3) show that ๐ฟ := ๐ฟ/๐ท can be understood as opportunity
หœ for the standard power utility function.
process under ๐‘ƒ
be a martingale with
a suitable parametrization) do not depend on the current level of wealth, a
special feature implied by the choice of power utility. If
ห† = ๐‘‹(ห†
ห†
๐‘‹
๐œ‹ , ๐‘ห†), and ๐œ…
ห† = ๐‘ห†/๐‘‹
(ห†
๐œ‹ , ๐‘ห†) โˆˆ ๐’œ is optimal,
(ห†
๐œ‹, ๐œ…
ห†)
is the optimal propensity to consume, then
denes a conditionally optimal strategy for the problem
ess sup ๐ธ
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
[โˆซ
๐‘ก
๐‘‡
]
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก ;
for
any (๐œ‹, ๐‘) โˆˆ ๐’œ, ๐‘ก โˆˆ [0, ๐‘‡ ].
(๐œ‹, ๐‘) โˆˆ ๐’œ and ๐‘ก โˆˆ [0, ๐‘‡ ]. Dene the pair (¯
๐œ‹ , ๐‘¯) by ๐œ‹
¯ =
๐‘‹๐‘ก (๐œ‹,๐‘)
¯
๐œ‹1[0,๐‘ก] + ๐œ‹
ห† 1(๐‘ก,๐‘‡ ] and ๐‘¯ = ๐‘1[0,๐‘ก] + ห† ๐‘ห†1(๐‘ก,๐‘‡ ] and let ๐‘‹ := ๐‘‹(¯
๐œ‹ , ๐‘¯). Note
๐‘‹๐‘ก
that (ห†
๐œ‹ , ๐‘ห†) is conditionally optimal in ๐’œ(ห†
๐œ‹ , ๐‘ห†, ๐‘ก), as otherwise Fact 6.1 yields a
contradiction to the global optimality of (ห†
๐œ‹ , ๐‘ห†). Now (3.4) with (ห‡
๐œ‹ , ๐‘ห‡) := (ห†
๐œ‹ , ๐‘ห†)
shows that (¯
๐œ‹ , ๐‘¯) is conditionally optimal in ๐’œ(๐œ‹, ๐‘, ๐‘ก). The result follows as
¯ = ๐‘ห†/๐‘‹
ห† =๐œ…
๐‘¯/๐‘‹
ห† on (๐‘ก, ๐‘‡ ] by Fact 6.3.
To see this, x
The martingale optimality principle of dynamic programming takes the
following form in our setting.
II.3 The Opportunity Process
13
Proposition
3.4. Let (๐œ‹, ๐‘) โˆˆ ๐’œ be an admissible strategy and assume that
โˆซ๐‘‡
โˆ˜
๐ธ[ 0 ๐‘ˆ๐‘  (๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ )] > โˆ’โˆž. Then the process
๐ฟ๐‘ก ๐‘1
(
)๐‘
๐‘‹๐‘ก (๐œ‹, ๐‘) +
๐‘ก
โˆซ
๐‘ˆ๐‘  (๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ),
๐‘ก โˆˆ [0, ๐‘‡ ]
0
is a supermartingale; it is a martingale if and only if (๐œ‹, ๐‘) is optimal.
Proof.
Combine Proposition 3.1 and Proposition 6.2.
The following lemma collects some elementary properties of
๐ฟ.
The
bounds are obtained by comparison with no-trade strategies, hence they are
๐ท is deterministic or if there are con๐‘˜1 , ๐‘˜2 > 0 as in (2.4), we obtain bounds which are model-independent;
independent of the price process. If
stants
they depend only on the utility function and the time to maturity.
Lemma 3.5. The opportunity process ๐ฟ is a special semimartingale.
(i) If ๐‘ โˆˆ (0, 1), ๐ฟ is a supermartingale satisfying
(
)โˆ’๐‘ [
๐ฟ๐‘ก โ‰ฅ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]
๐ธ
โˆซ
๐‘‡
]
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก ,
๐‘ก
0โ‰ค๐‘กโ‰ค๐‘‡
(3.5)
and ๐ฟ, ๐ฟโˆ’ > 0. In particular, ๐ฟ โ‰ฅ ๐‘˜1 if ๐ท โ‰ฅ ๐‘˜1 .
(ii) If ๐‘ < 0, ๐ฟ satises
)โˆ’๐‘ [
0 โ‰ค ๐ฟ๐‘ก โ‰ค ๐œ‡ [๐‘ก, ๐‘‡ ]
๐ธ
(
โˆ˜
โˆซ
๐‘‡
๐‘ก
]
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก ,
0โ‰ค๐‘กโ‰ค๐‘‡
(3.6)
and in particular ๐ฟ๐‘ก โ‰ค ๐‘˜2 ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] 1โˆ’๐‘ if ๐ท โ‰ค ๐‘˜2 . In the case without
intermediate consumption, ๐ฟ is a submartingale. Moreover, in both
cases, ๐ฟ, ๐ฟโˆ’ > 0 if there exists an optimal strategy (ห†๐œ‹, ๐‘ห†).
(
Proof.
)
๐‘ > 0, or ๐‘ < 0 and there is no
intermediate consumption. Then ๐œ‹ โ‰ก 0, ๐‘ โ‰ก ๐‘ฅ0 1{๐‘‡ } is an admissible stratโˆซ๐‘ก
1 ๐‘
1 ๐‘
egy and Proposition 3.4 shows that ๐ฟ๐‘ก ๐‘ฅ0 +
๐‘
0 ๐‘ˆ๐‘  (0) ๐œ‡(๐‘‘๐‘ ) = ๐ฟ๐‘ก ๐‘ ๐‘ฅ0 is a
Consider the cases where either
supermartingale, proving the super/submartingale properties in (i) and (ii).
Let
๐‘
be arbitrary and assume there is no intermediate consumption.
๐œ‹ โ‰ก 0 and ๐‘ โ‰ก ๐‘ฅ0 1{๐‘‡ } , we get ๐ฟ๐‘ก ๐‘1 ๐‘ฅ๐‘0 โ‰ฅ ๐ธ[๐‘ˆ๐‘‡ (๐‘๐‘‡ )โˆฃโ„ฑ๐‘ก ] =
Hence ๐ฟ๐‘ก โ‰ฅ ๐ธ[๐ท๐‘‡ โˆฃโ„ฑ๐‘ก ] if ๐‘ > 0 and ๐ฟ๐‘ก โ‰ค ๐ธ[๐ท๐‘‡ โˆฃโ„ฑ๐‘ก ] if ๐‘ < 0,
Applying (3.3) with
๐ธ[๐ท๐‘‡ โˆฃโ„ฑ๐‘ก ] ๐‘1 ๐‘ฅ๐‘0 .
which corresponds to (3.5) and (3.6) for this case.
๐‘ is arbitrary), we consume
๐‘ก. That is, we use (3.3) with ๐œ‹ โ‰ก 0
]
[โˆซ๐‘‡
1 ๐‘
โˆ’1 1
โˆ˜ (๐‘‘๐‘ )โ„ฑ =
and ๐‘ = ๐‘ฅ0 (๐‘‡ โˆ’ ๐‘ก + 1)
๐‘ˆ
(๐‘
)
๐œ‡
๐‘ 
๐‘ 
๐‘ก
[๐‘ก,๐‘‡ ] to obtain ๐ฟ๐‘ก ๐‘ ๐‘ฅ0 โ‰ฅ ๐ธ
๐‘ก
]
[โˆซ๐‘‡
1 ๐‘
โˆ’๐‘
โˆ˜
๐‘ ๐‘ฅ0 (1 + ๐‘‡ โˆ’ ๐‘ก) ๐ธ
๐‘ก ๐ท๐‘  ๐œ‡ (๐‘‘๐‘ ) โ„ฑ๐‘ก . This ends the proof of (3.5) and (3.6).
In the case ๐‘ < 0, (3.6) shows that ๐ฟ is dominated by a martingale, hence
๐ฟ is of class (D) and in particular a special semimartingale.
If there is intermediate consumption (and
at a constant rate after the xed time
II Opportunity Process
14
If ๐‘ > 0, (3.5) shows ๐ฟ > 0 and
๐ฟโˆ’ > 0 follows by the minimum principle for positive supermartinห† = ๐‘‹(ห†
gales. For ๐‘ < 0, let ๐‘‹
๐œ‹ , ๐‘ห†) be the optimal wealth process. Clearly
๐ฟ > 0 follows
from (3.3) with (ห†
๐œ‹ , ๐‘ห†). From Proposition 3.4 we have that
โˆซ
1 ห†๐‘
ห† ๐‘ ๐ฟ is a positive su๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ) is a negative martingale, hence ๐‘‹
๐‘ ๐‘‹ ๐ฟ + ๐‘ˆ๐‘  (ห†
ห† ๐‘ ๐ฟ๐‘ก > 0] = 1 and it remains to note
permartingale. Therefore ๐‘ƒ [inf 0โ‰ค๐‘กโ‰ค๐‘‡ ๐‘‹
๐‘ก
ห† ๐‘ are ๐‘ƒ -a.s. bounded because ๐‘‹,
ห† ๐‘‹
ห†โˆ’ > 0.
that the paths of ๐‘‹
It remains to prove the positivity.
then
The following concerns the submartingale property in Lemma 3.5(ii).
Example 3.6. Consider the case with intermediate consumption and assume
๐ท โ‰ก 1 and ๐‘† โ‰ก 1. Then an optimal strategy is given by (ห†
๐œ‹ , ๐‘ห†) โ‰ก
(0, ๐‘ฅ0 /(1+๐‘‡ )) and ๐ฟ๐‘ก = (1+๐‘‡ โˆ’๐‘ก)1โˆ’๐‘ is a decreasing function. In particular,
๐ฟ is not a submartingale.
that
Remark 3.7. We can also consider the utility maximization problem under
constraints in the following sense. Suppose that for each (๐œ”, ๐‘ก) โˆˆ ฮฉ × [0, ๐‘‡ ]
C๐‘ก (๐œ”) โŠ† โ„๐‘‘ . We assume that each of these sets contains
the origin. A strategy (๐œ‹, ๐‘) โˆˆ ๐’œ is called C -admissible if ๐œ‹๐‘ก (๐œ”) โˆˆ C๐‘ก (๐œ”) for
C
all (๐œ”, ๐‘ก), and the set of all these strategies is denoted by ๐’œ . The example
โˆ˜
C
(๐œ‹, ๐‘) โ‰ก (0, ๐‘ฅ0 /๐œ‡ [0, ๐‘‡ ]) shows that ๐’œ โˆ•= โˆ….
We do not impose assumptions on the set-valued mapping C at this stage.
we are given a set
For dynamic programming, the relevant point is that the constraints are
(๐œ”, ๐‘ก),
specied as a pointwise condition in
rather than as a set of processes
๐œ‹ . We note that all arguments in this section remain valid if ๐’œ is replaced by
๐’œC throughout. This generalization is not true for the subsequent section,
and existence of an optimal strategy is not guaranteed for general C .
II.4 Relation to the Dual Problem
We discuss how the problem dual to utility maximization relates to the
๐ฟ. We assume (2.4) and (2.6) in the entire Section II.4, hence Proposition 2.3 applies. The dual problem will be dened
opportunity process
on a domain
Y
introduced below.
Since its denition is slightly cumber-
some, we point out that to follow the results in the body of this chapter,
Y are needed. First, the density process
๐‘„ โˆˆ M ๐‘† , scaled by a certain constant ๐‘ฆ0 , is
each element of Y is a positive supermartingale.
only two facts about
of each mar-
tingale measure
contained in
Y.
Second,
Following [43], the
dual problem
inf
๐‘Œ โˆˆY (๐‘ฆ0 )
where
๐‘ฆ0 := ๐‘ขโ€ฒ (๐‘ฅ0 ) = ๐ฟ0 ๐‘ฅ๐‘โˆ’1
0
๐ธ
[โˆซ
is
๐‘‡
]
๐‘ˆ๐‘กโˆ— (๐‘Œ๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) ,
and
๐‘ˆ๐‘กโˆ—
is the convex conjugate of
๐‘ˆ๐‘กโˆ— (๐‘ฆ) := sup ๐‘ˆ๐‘ก (๐‘ฅ) โˆ’ ๐‘ฅ๐‘ฆ = โˆ’ 1๐‘ž ๐‘ฆ ๐‘ž ๐ท๐‘ก๐›ฝ .
{
๐‘ฅ>0
(4.1)
0
}
๐‘ฅ 7โ†’ ๐‘ˆ๐‘ก (๐‘ฅ),
(4.2)
II.4 Relation to the Dual Problem
15
We have denoted by
1
> 0,
1โˆ’๐‘
๐›ฝ :=
๐‘ž :=
๐‘
โˆˆ (โˆ’โˆž, 0) โˆช (0, 1)
๐‘โˆ’1
the relative risk tolerance and the exponent conjugate to
(4.3)
๐‘,
respectively.
These constants will be used very often in the sequel and it is useful to note
sign(๐‘) = โˆ’ sign(๐‘ž).
It remains to dene the domain
X = {๐ป โˆ™ ๐‘† : ๐ป โˆˆ ๐ฟ(๐‘†), ๐ป โˆ™ ๐‘†
be the set of gains processes from trading.
Y = Y (๐‘ฆ0 ).
Let
is bounded below}
The set of supermartingale
densities is dened by
Y โˆ— = {๐‘Œ โ‰ฅ 0
càdlàg
: ๐‘Œ0 โ‰ค ๐‘ฆ0 ,
๐‘Œ๐บ
supermartingale for all
its subset corresponding to probability measures equivalent to
Y M = {๐‘Œ โˆˆ Y โˆ— : ๐‘Œ > 0
is a martingale and
๐บ โˆˆ X };
๐‘ƒ
on
โ„ฑ๐‘‡
is
๐‘Œ0 = ๐‘ฆ0 }.
We place ourselves in the setting of [43] by considering the same dual domain
Y D โŠ† Y โˆ—.
It consists of density processes of (the regular parts of ) the
๐œŽ((๐ฟโˆž )โˆ— , ๐ฟโˆž )-closure of {๐‘Œ๐‘‡ : ๐‘Œ โˆˆ Y M } โŠ‚
โˆž
โˆ—
โŠ† (๐ฟ ) . More precisely, we multiply each density with the constant ๐‘ฆ0 .
D is not important
We refer to [43] for details as the precise construction of Y
nitely additive measures in the
๐ฟ1
here, it is relevant for us only that
Y
Y M โŠ† Y D โŠ† Y โˆ—.
In particular,
D if we identify measures and their density processes.
๐‘ฆ0 M ๐‘† โŠ†
For notational
reasons, we make the dual domain slightly smaller and let
Y := {๐‘Œ โˆˆ Y D : ๐‘Œ > 0}.
By [43, Theorem 3.10] there exists a unique
๐‘Œห† = ๐‘Œห† (๐‘ฆ0 ) โˆˆ Y
such that the
inmum in (4.1) is attained, and it is related to the optimal consumption
๐‘ห†
via the marginal utility by
๐‘Œห†๐‘ก = โˆ‚๐‘ฅ ๐‘ˆ๐‘ก (๐‘ฅ)โˆฃ๐‘ฅ=ห†๐‘๐‘ก = ๐ท๐‘ก ๐‘ห†๐‘โˆ’1
๐‘ก
on the support of
๐œ‡โˆ˜ .
(4.4)
In the case without intermediate consumption, an
existence result was previously obtained in [49].
Remark 4.1.
All the results stated below remain true if we replace
{๐‘Œ โˆˆ Y โˆ— : ๐‘Œ > 0};
Y
by
i.e., it is not important for our purposes whether we use
the dual domain of [43] or the one of [49]. This is easily veried using the
Y D contains all maximal elements of Y โˆ— (see [43, Theorem 2.10]).
โˆ— is called maximal if ๐‘Œ = ๐‘Œ โ€ฒ ๐ต , for some ๐‘Œ โ€ฒ โˆˆ Y โˆ— and some
Here ๐‘Œ โˆˆ Y
càdlàg nonincreasing process ๐ต โˆˆ [0, 1], implies ๐ต โ‰ก 1.
fact that
II Opportunity Process
16
Proposition 4.2. Let (ห†๐‘, ๐œ‹ห† ) โˆˆ ๐’œ be an optimal strategy and
The solution to the dual problem is given by
ห† = ๐‘‹(ห†
๐‘‹
๐œ‹ , ๐‘ห†).
ห† ๐‘โˆ’1 .
๐‘Œห† = ๐ฟ๐‘‹
Proof.
๐ฟ๐‘‡ = ๐ท๐‘‡
As
ห†๐‘‡ ,
๐‘ห†๐‘‡ = ๐‘‹
and
(4.4) already yields
Moreover, by Lemma 7.1 in the Appendix,
โˆซ
ห†๐‘ก +
๐‘๐‘ก := ๐‘Œห†๐‘ก ๐‘‹
๐‘Œห†
ห† ๐‘โˆ’1 .
๐‘Œห†๐‘‡ = ๐ฟ๐‘‡ ๐‘‹
๐‘‡
has the property that
๐‘ก
โˆซ
0
๐‘ก
๐‘ˆ๐‘  (ห†
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ )
ห†๐‘ก + ๐‘
๐‘Œห†๐‘  ๐‘ห†๐‘  ๐œ‡(๐‘‘๐‘ ) = ๐‘Œห†๐‘ก ๐‘‹
0
is a martingale. By Proposition 3.4,
ห†๐‘ + ๐‘
๐‘หœ๐‘ก := ๐ฟ๐‘ก ๐‘‹
๐‘ก
โˆซ๐‘ก
0
๐‘ˆ๐‘  (ห†
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ )
is also a
martingale. The terminal values of these martingales coincide, hence
We deduce
๐‘Œห† =
The formula
๐ฟ.
ห† ๐‘โˆ’1 as
๐ฟ๐‘‹
หœ = ๐‘.
๐‘
ห† > 0.
๐‘‹
ห† ๐‘โˆ’1
๐‘Œห† = ๐ฟ๐‘‹
could be used to
dene the opportunity process
This is the approach taken in Muhle-Karbe [58] (see also Kallsen and
Muhle-Karbe [40]), where utility from terminal wealth is considered and the
opportunity process is used as a tool to verify the optimality of an explicit
candidate solution. Our approach via the value process has the advantage
that it immediately yields the properties in Lemma 3.5 and monotonicity
results (see Section II.5).
II.4.1 The Dual Opportunity Process
We now introduce the analogue of
and
๐‘ก โˆˆ [0, ๐‘‡ ]
๐ฟ for the dual problem.
Dene for
๐‘Œ โˆˆY
the set
{
Y (๐‘Œ, ๐‘ก) := ๐‘Œหœ โˆˆ Y : ๐‘Œหœ = ๐‘Œ
on
}
[0, ๐‘ก] .
We recall the constants (4.3) and the standing assumptions (2.4) and (2.6).
Proposition 4.3. There exists a unique càdlàg process ๐ฟโˆ— , called
portunity process, such that for all ๐‘Œ โˆˆ Y and ๐‘ก โˆˆ [0, ๐‘‡ ],
โˆ’ 1๐‘ž ๐‘Œ๐‘ก๐‘ž ๐ฟโˆ—๐‘ก = ess inf ๐ธ
๐‘Œหœ โˆˆY (๐‘Œ,๐‘ก)
[โˆซ
๐‘ก
๐‘‡
dual op-
]
๐‘ˆ๐‘ โˆ— (๐‘Œหœ๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก .
An alternative description is
โŽง
]
[โˆซ
๏ฃด
โŽจess sup๐‘Œ โˆˆY ๐ธ ๐‘‡ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘ก
๐ฟโˆ—๐‘ก =
]
[โˆซ
๏ฃด
๐‘‡
โŽฉ
ess inf ๐‘Œ โˆˆY ๐ธ ๐‘ก ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
and the extrema are attained at ๐‘Œ = ๐‘Œห† .
if ๐‘ž โˆˆ (0, 1),
if ๐‘ž < 0
II.4 Relation to the Dual Problem
Proof.
17
Y [43, Theorem 2.10] shows that if ๐‘Œ, ๐‘Œห‡ โˆˆ Y
หœ โˆˆ Y (๐‘Œห‡ , ๐‘ก), then ๐‘Œ 1[0,๐‘ก) + (๐‘Œ๐‘ก /๐‘Œห‡๐‘ก )๐‘Œหœ 1[๐‘ก,๐‘‡ ] is in Y (๐‘Œ, ๐‘ก). It also implies
and ๐‘Œ
1
2 โˆˆ Y (๐‘Œ, ๐‘ก), then ๐‘Œ 1 1 + ๐‘Œ 2 1 ๐‘ โˆˆ Y (๐‘Œ, ๐‘ก). The
that if ๐ด โˆˆ โ„ฑ๐‘ก and ๐‘Œ , ๐‘Œ
๐ด
๐ด
The fork convexity of
proof of the rst claim is now analogous to that of Proposition 3.1.
โˆ—
second part follows by using that ๐ฟ does not depend on
The process
๐ฟโˆ—
is related to
Proposition 4.4. Let ๐›ฝ =
Proof.
1
1โˆ’๐‘
๐ฟ
The
๐‘Œ.
by a simple power transformation.
. Then ๐ฟโˆ— = ๐ฟ๐›ฝ .
โˆซ
ห†๐‘ก ๐‘Œห†๐‘ก + ๐‘ก ๐‘ห†๐‘  ๐‘Œห†๐‘  ๐œ‡(๐‘‘๐‘ ) from Lemma 7.1
๐‘๐‘ก := ๐‘‹
0
]
[โˆซ
โˆซ
ห†๐‘ก ๐‘Œห†๐‘ก = ๐ธ[๐‘๐‘‡ โˆฃโ„ฑ๐‘ก ] โˆ’ ๐‘ก ๐‘ห†๐‘  ๐‘Œห†๐‘  ๐œ‡(๐‘‘๐‘ ) = ๐ธ ๐‘‡ ๐‘ห†๐‘  ๐‘Œห†๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก =
implies that ๐‘‹
0
๐‘ก
]
[โˆซ๐‘‡
๐ธ ๐‘ก ๐ท๐‘ ๐›ฝ ๐‘Œห†๐‘ ๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก , where the last equality is obtained by expressing ๐‘ห†
ห† ๐‘ž ๐ฟโˆ—๐‘ก by Proposition 4.3; so we have
via (4.4). The right hand side equals ๐‘Œ
๐‘ก
ห† ๐‘Œห† = ๐‘Œห† ๐‘ž ๐ฟโˆ— . On the other hand, (๐ฟ๐‘‹
ห† ๐‘โˆ’1 )๐‘ž = ๐‘Œห† ๐‘ž by Proposition 4.2
shown ๐‘‹
ห† ๐‘Œห† = ๐‘Œห† ๐‘ž ๐ฟ๐›ฝ . We deduce ๐ฟโˆ— = ๐ฟ๐›ฝ as ๐‘Œห† > 0.
and this can be written as ๐‘‹
The martingale property of
II.4.2 Reverse Hölder Inequality and Boundedness of ๐ฟ
Let
๐‘Œ,
๐‘ž=
๐‘
๐‘โˆ’1 be the exponent conjugate to
๐‘.
Given a general positive process
we consider the following inequality of reverse Hölder type:
โŽงโˆซ
๐‘‡ [
]
๏ฃด
๏ฃด
๏ฃด
๐ธ (๐‘Œ๐‘  /๐‘Œ๐œ )๐‘ž โ„ฑ๐œ ๐œ‡โˆ˜ (๐‘‘๐‘ ) โ‰ค ๐ถ๐‘ž
โŽจ
๐œ
โˆซ ๐‘‡
๏ฃด
]
[
๏ฃด
๏ฃด
โŽฉ
๐ธ (๐‘Œ๐‘  /๐‘Œ๐œ )๐‘ž โ„ฑ๐œ ๐œ‡โˆ˜ (๐‘‘๐‘ ) โ‰ฅ ๐ถ๐‘ž
๐‘ž < 0,
if
(R๐‘ž (๐‘ƒ ))
๐‘ž โˆˆ (0, 1),
if
๐œ
0 โ‰ค ๐œ โ‰ค ๐‘‡ and some constant ๐ถ๐‘ž > 0 independent
of ๐œ . It is useful to recall that ๐‘ž < 0 corresponds to ๐‘ โˆˆ (0, 1) and vice versa.
๐‘ž
Without intermediate consumption, R๐‘ž (๐‘ƒ ) reduces to ๐ธ[(๐‘Œ๐‘‡ /๐‘Œ๐œ ) โˆฃโ„ฑ๐œ ] โ‰ค
๐ถ๐‘ž (resp. โ‰ฅ). Inequalities of this type are well known. See, e.g., Doléansfor all stopping times
Dade and Meyer [19] for an introduction or Delbaen et al. [16] and the
references therein for some connections to nance. In most applications, the
considered exponent
๐‘ž < 0.
๐‘ž
is greater than one;
R๐‘ž (๐‘ƒ )
then takes the form as for
We recall once more the standing assumptions (2.4) and (2.6).
Proposition 4.5. The following are equivalent:
(i) The process ๐ฟ is uniformly bounded away from zero and innity.
(ii) Inequality R๐‘ž (๐‘ƒ ) holds for the dual minimizer ๐‘Œห† โˆˆ Y .
(iii) Inequality R๐‘ž (๐‘ƒ ) holds for some ๐‘Œ โˆˆ Y .
Proof.
๐ฟ always
๐ฟ โ‰ค ๐‘๐‘œ๐‘›๐‘ ๐‘ก. if ๐‘ < 0.
Under the standing assumption (2.4), a one-sided bound for
holds by Lemma 3.5, namely
๐ฟ โ‰ฅ ๐‘˜1
if
๐‘ โˆˆ (0, 1)
and
(i) is equivalent to (ii): We use (2.4) and then Propositions 4.3 and 4.4
to obtain that
โˆซ๐‘‡
๐œ
]
]
[
[โˆซ๐‘‡
๐ธ (๐‘Œห†๐‘  /๐‘Œห†๐œ )๐‘ž โ„ฑ๐œ ๐œ‡โˆ˜ (๐‘‘๐‘ ) = ๐ธ ๐œ (๐‘Œห†๐‘  /๐‘Œห†๐œ )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐œ โ‰ค
II Opportunity Process
18
๐‘˜1โˆ’๐›ฝ ๐ธ
]
๐ท๐‘ ๐›ฝ (๐‘Œห†๐‘  /๐‘Œห†๐œ )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐œ = ๐‘˜1โˆ’๐›ฝ ๐ฟโˆ—๐œ = ๐‘˜1โˆ’๐›ฝ ๐ฟ๐›ฝ๐œ . Thus when ๐‘ โˆˆ (0, 1)
ห† is equivalent to an upper bound for ๐ฟ. For
and hence ๐‘ž < 0, R๐‘ž (๐‘ƒ ) for ๐‘Œ
๐‘ < 0, we replace ๐‘˜1 by ๐‘˜2 .
(iii) implies (i): Assume ๐‘ โˆˆ (0, 1). Using Propositions 4.4 and 4.3
]
[โˆซ๐‘‡ โˆ—
โˆซ
๐‘ž
1 ๐‘ž ๐›ฝ
1 ๐›ฝ ๐‘‡
โˆ˜
โˆ˜
and (4.2), โˆ’ ๐‘Œ๐‘ก ๐ฟ๐‘ก โ‰ค ๐ธ
๐‘ž
๐‘ก ๐‘ˆ๐‘  (๐‘Œ๐‘  ) ๐œ‡ (๐‘‘๐‘ ) โ„ฑ๐‘ก โ‰ค โˆ’ ๐‘ž ๐‘˜2 ๐‘ก ๐ธ[๐‘Œ๐‘  โˆฃโ„ฑ๐‘ก ] ๐œ‡ (๐‘‘๐‘ ).
[โˆซ๐‘‡
Hence
๐œ
๐ฟ โ‰ค ๐‘˜2 ๐ถ๐‘žโˆ’๐›ฝ .
If
๐‘ < 0,
we obtain
๐ฟ โ‰ฅ ๐‘˜1 ๐ถ๐‘žโˆ’๐›ฝ
in the same way.
If the equivalent conditions of Proposition 4.5 are satised, we say that
R๐‘ž (๐‘ƒ ) holds for the given nancial market model. Although quite frequent
in the literature, this condition is rather restrictive in the sense that it often
fails in explicit models that have stochastic dynamics. For instance, in the
ane models of [40],
๐ฟ
is an exponentially ane function of a typically
unbounded factor process, in which case Proposition 4.5 implies that
fails.
Similarly,
๐ฟ
R๐‘ž (๐‘ƒ )
is an exponentially quadratic function of an Ornstein-
Uhlenbeck process in the model of Kim and Omberg [47].
On the other
hand, exponential Lévy models have constant dynamics and here
๐ฟ
turns
out to be simply a smooth deterministic function.
In a given model, it may be hard to check whether
calling
๐‘ฆ0
M๐‘†
to choose for
R๐‘ž (๐‘ƒ )
holds.
Re-
โŠ† Y , an obvious approach in view of Proposition 4.5(iii) is
๐‘Œ /๐‘ฆ0 the density process of some specic martingale measure.
We illustrate this with an essentially classical example.
Example 4.6.
Assume that
๐‘…
is a special semimartingale with decomposi-
tion
๐‘… = ๐›ผ โˆ™ โŸจ๐‘…๐‘ โŸฉ + ๐‘€ ๐‘… ,
where
and
๐‘…๐‘
๐‘€๐‘…
denotes the continuous local martingale part of
is the local martingale part of
โˆซ
๐œ’๐‘ก :=
๐‘ก
๐‘….
(4.5)
๐‘…, ๐›ผ โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘…๐‘ ),
Suppose that the process
๐›ผ๐‘ โŠค ๐‘‘โŸจ๐‘…๐‘ โŸฉ๐‘  ๐›ผ๐‘  ,
๐‘ก โˆˆ [0, ๐‘‡ ]
0
uniformly bounded.
๐‘ := โ„ฐ(โˆ’๐›ผ โˆ™ ๐‘…๐‘ ) is a martingale by Novikov's
condition and the measure ๐‘„ โ‰ˆ ๐‘ƒ with density ๐‘‘๐‘„/๐‘‘๐‘ƒ = ๐‘๐‘‡ is a local
๐‘
๐‘…
martingale measure for ๐‘† as ๐‘โ„ฐ(๐‘…) = โ„ฐ(โˆ’๐›ผ โˆ™ ๐‘… + ๐‘€ ) by Yor's formula;
(1
)
๐‘ž
๐‘
hence ๐‘ฆ0 ๐‘ โˆˆ Y . Fix ๐‘ž . Using ๐‘ = โ„ฐ(โˆ’๐‘ž๐›ผ โˆ™ ๐‘… ) exp
2 ๐‘ž(๐‘ž โˆ’ 1)๐œ’ , and that
โ„ฐ(โˆ’๐‘ž๐›ผ โˆ™ ๐‘…๐‘ ) is a martingale by Novikov's condition, one readily checks that
๐‘ satises inequality R๐‘ž (๐‘ƒ ).
If ๐‘… is continuous, (4.5) is the structure condition of Schweizer [71] and
under (2.6) ๐‘… is necessarily of this form. Then ๐œ’ is called mean-variance
tradeo process and ๐‘„ is the minimal martingale measure. In Itô process
โˆซ๐‘ก โŠค
models, ๐œ’ takes the form ๐œ’๐‘ก =
0 ๐œƒ๐‘  ๐œƒ๐‘  ๐‘‘๐‘ , where ๐œƒ is the market price of risk
process. Thus ๐œ’ will be bounded whenever ๐œƒ is.
is
Remark 4.7.
Then
The example also gives a sucient condition for (2.5). This
is of interest only for
๐‘ โˆˆ (0, 1)
and we remark that for the case of Itô
II.4 Relation to the Dual Problem
process models with bounded
๐œƒ,
19
the condition corresponds to Karatzas and
Shreve [42, Remark 6.3.9].
Indeed, if there exists
๐‘Œ โˆˆY
satisfying
R๐‘ž (๐‘ƒ ),
then with (4.2) and (2.4)
it follows that the the value of the dual problem (4.1) is nite, and this
suces for (2.5), as in Kramkov and Schachermayer [50].
R๐‘ž (๐‘ƒ )
The rest of the section studies the dependence of
Remark 4.8.
such that
on
๐‘ž.
๐‘Œ satises R๐‘ž (๐‘ƒ ) with a constant ๐ถ๐‘ž . If ๐‘ž1
0 < ๐‘ž < ๐‘ž1 < 1, then R๐‘ž1 (๐‘ƒ ) is satised with
Assume that
๐‘ž < ๐‘ž1 < 0
or
is
(
)1โˆ’๐‘ž1 /๐‘ž
๐ถ๐‘ž1 = ๐œ‡โˆ˜ [0, ๐‘‡ ]
(๐ถ๐‘ž )๐‘ž1 /๐‘ž .
Similarly, if
๐‘ž < 0 < ๐‘ž1 < 1,
we can take
๐ถ๐‘ž1 = (๐ถ๐‘ž )๐‘ž1 /๐‘ž .
This follows from
Jensen's inequality.
There is also a partial converse.
Lemma 4.9. Let 0 < ๐‘ž < ๐‘ž1 < 1 and let ๐‘Œ > 0 be a supermartingale. If ๐‘Œ
satises R๐‘ž1 (๐‘ƒ ), it also satises R๐‘ž (๐‘ƒ ).
In particular, the
{ following dichotomy
} holds: ๐‘Œ satises either all or none
of the inequalities R๐‘ž (๐‘ƒ ), ๐‘ž โˆˆ (0, 1) .
Proof.
โˆซ๐‘‡
โˆ˜
๐‘ž
๐‘ก ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ„ฑ๐‘ก ๐œ‡ (๐‘‘๐‘ ) โ‰ฅ
1โˆ’๐‘ž
])
โˆซ๐‘‡ ( [
1โˆ’๐‘ž
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž1 โ„ฑ๐‘ก 1โˆ’๐‘ž1 ๐œ‡โˆ˜ (๐‘‘๐‘ ). Noting that 1โˆ’๐‘ž
> 1, we apply Jensen's
๐‘ก
1
inequality to the right-hand side and then use R๐‘ž1 (๐‘ƒ ) to deduce the claim
with
From Lemma 4.10 stated below we have
1โˆ’๐‘ž
(
) ๐‘žโˆ’๐‘ž1
๐ถ๐‘ž := ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] 1โˆ’๐‘ž1 (๐ถ๐‘ž1 ) 1โˆ’๐‘ž1 .
[
]
The dichotomy follows by the previous
remark.
For future reference, we state separately the main step of the above proof.
Lemma 4.10. Let ๐‘Œ
>0
be a supermartingale. For xed 0 โ‰ค ๐‘ก โ‰ค ๐‘  โ‰ค ๐‘‡ ,
๐œ™ : (0, 1) โ†’ โ„+ ,
1
( [
]) 1โˆ’๐‘ž
๐‘ž 7โ†’ ๐œ™(๐‘ž) := ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
is a monotone decreasing function
]) ๐‘Œ is a martingale,
(
[ ๐‘ƒ -a.s. If in addition
then lim๐‘žโ†’1โˆ’ ๐œ™(๐‘ž) = exp โˆ’ ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก ) log(๐‘Œ๐‘  /๐‘Œ๐‘ก )โ„ฑ๐‘ก ๐‘ƒ -a.s., where the
conditional expectation has values in โ„ โˆช {+โˆž}.
Proof.
Suppose rst that ๐‘Œ is a martingale; by
๐ธ[๐‘Œ. ] = 1. We dene a probability ๐‘„ โ‰ˆ ๐‘ƒ on โ„ฑ๐‘ 
๐‘Ÿ := (1 โˆ’ ๐‘ž) โˆˆ (0, 1) and Bayes' formula,
scaling we may assume
by
๐‘‘๐‘„/๐‘‘๐‘ƒ := ๐‘Œ๐‘  .
1
(
( [
]) ๐‘Ÿ1
]) 1โˆ’๐‘ž
[
๐œ™(๐‘ž) = ๐‘Œ๐‘ก1โˆ’๐‘ž ๐ธ ๐‘„ ๐‘Œ๐‘ ๐‘žโˆ’1 โ„ฑ๐‘ก
= ๐‘Œ๐‘ก ๐ธ ๐‘„ (1/๐‘Œ๐‘  )๐‘Ÿ โ„ฑ๐‘ก
.
This is increasing in
๐‘Ÿ
by Jensen's inequality, hence decreasing in
๐‘ž.
With
20
II Opportunity Process
Now let ๐‘Œ be a supermartingale. We can decompose it as ๐‘Œ๐‘ข = ๐ต๐‘ข ๐‘€๐‘ข ,
๐‘ข โˆˆ [0, ๐‘ ], where ๐‘€ is a martingale and ๐ต๐‘  = 1. That is, ๐‘€๐‘ก = ๐ธ[๐‘Œ๐‘  โˆฃโ„ฑ๐‘ก ]
๐‘ž/(๐‘žโˆ’1)
and ๐ต๐‘ก = ๐‘Œ๐‘ก /๐ธ[๐‘Œ๐‘  โˆฃโ„ฑ๐‘ก ] โ‰ฅ 1, by the supermartingale property. Hence ๐ต๐‘ก
is decreasing in ๐‘ž โˆˆ (0, 1). Together with the rst part, it follows that
]) 1
๐‘ž/(๐‘žโˆ’1) ( [
๐œ™(๐‘ž) = ๐ต๐‘ก
๐ธ (๐‘€๐‘  /๐‘€๐‘ก )๐‘ž โ„ฑ๐‘ก 1โˆ’๐‘ž is decreasing.
(
)
Assume again that ๐‘Œ is a martingale. The limit lim๐‘žโ†’1โˆ’ log ๐œ™(๐‘ž) can
be calculated as
])
]
( [
[
log ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž log(๐‘Œ๐‘  /๐‘Œ๐‘ก )โ„ฑ๐‘ก
]
[
lim
= lim โˆ’
๐‘žโ†’1โˆ’
๐‘žโ†’1โˆ’
1โˆ’๐‘ž
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
using l'Hôpital's rule and
๐ธ[(๐‘Œ๐‘  /๐‘Œ๐‘ก )โˆฃโ„ฑ๐‘ก ] = 1.
๐‘ƒ -a.s.
The result follows using mono-
tone and bounded convergence in the numerator and dominated convergence
in the denominator.
Remark 4.11.
in] entropic
โˆซ ๐‘‡ ๐‘ž [= 1 corresponds to the
โˆ˜ (๐‘‘๐‘ ) โ‰ค ๐ถ .
โ„ฑ
๐ธ
(๐‘Œ
/๐‘Œ
)
log(๐‘Œ
/๐‘Œ
)
๐œ‡
๐œ
๐‘ 
๐œ
๐‘ 
๐œ
1
๐œ
Lemma 4.10 shows that for a martingale ๐‘Œ > 0, R๐‘ž1 (๐‘ƒ ) with ๐‘ž1 โˆˆ (0, 1)
is weaker than R๐ฟ log ๐ฟ (๐‘ƒ ), which, in turn, is obviously weaker than R๐‘ž0 (๐‘ƒ )
with ๐‘ž0 > 1.
A much deeper argument [19, Proposition 5] shows that if ๐‘Œ is a martinโˆ’1 ๐‘Œ โ‰ค ๐‘Œ โ‰ค ๐‘˜๐‘Œ for some ๐‘˜ > 0,
gale satisfying the condition (S) that ๐‘˜
โˆ’
โˆ’
then ๐‘Œ satises R๐‘ž0 (๐‘ƒ ) for some ๐‘ž0 > 1 if and only if it satises R๐‘ž (๐‘ƒ ) for
some ๐‘ž < 0, and then by Remark 4.8 also R๐‘ž1 (๐‘ƒ ) for all ๐‘ž1 โˆˆ (0, 1).
equality
The limiting case
R๐ฟ log ๐ฟ (๐‘ƒ )
which reads
Coming back to the utility maximization problem, we obtain the follow-
โ‡’
ing dichotomy from Lemma 4.9 and the implication (iii)
(ii) in Proposi-
tion 4.5.
Corollary 4.12. For the given market model, R๐‘ž (๐‘ƒ ) holds either for all or
no values of ๐‘ž โˆˆ (0, 1).
II.5 Applications
In this section we consider only the case
We assume (2.4) and (2.6).
with
intermediate consumption.
However, we remark that all results except for
Proposition 5.4 and Remark 5.5 hold true as soon as there exists an optimal
strategy
(ห†
๐œ‹ , ๐‘ห†) โˆˆ ๐’œ.
We rst show that given the opportunity process, the optimal propensity
๐œ…
ห† can be expressed in feedback form, and therefore any result
๐ฟ leads to a statement about ๐œ…
ห† . This extends results known for special
to consume
about
settings (e.g., Stoikov and Zariphopoulou [72]).
Theorem 5.1. With ๐›ฝ =
๐‘ห†๐‘ก =
( ๐ท )๐›ฝ
๐‘ก
๐ฟ๐‘ก
1
1โˆ’๐‘
ห†๐‘ก
๐‘‹
we have
and hence
๐œ…
ห†๐‘ก =
( ๐ท )๐›ฝ
๐‘ก
๐ฟ๐‘ก
.
(5.1)
II.5 Applications
Proof.
21
This follows from Proposition 4.2 via (4.4) and (2.2).
Remark 5.2.
In Theorem III.3.2 (and Remark III.3.6) of Chapter III we
establish the same formula for
๐œ…
ห†
in the utility maximization problem under
constraints as described in Remark 3.7, under the sole assumption that an
optimal constrained strategy exists.
The special case where the constraints set
C โŠ† โ„๐‘‘
duced from Theorem 5.1 by redening the price process
๐‘†1
โ‰ก1
for
C =
{(๐‘ฅ1 , . . . , ๐‘ฅ๐‘‘ )
โˆˆ
โ„๐‘‘
๐‘ฅ1
:
is linear can be de-
๐‘†.
For instance, set
= 0}.
In the remainder of the section we discuss how certain changes in the
model and the discounting process
๐ท
aect the optimal propensity to con-
sume. This is based on (5.1) and the relation
1 ๐‘
๐‘ ๐‘ฅ0 ๐ฟ๐‘ก
=
ess sup
๐ธ
[โˆซ
๐‘‡
๐‘ก
๐‘โˆˆ๐’œ(0,๐‘ฅ0 1{๐‘‡ } ,๐‘ก)
which is immediate from Proposition 3.1.
]
๐ท๐‘  ๐‘1 ๐‘๐‘๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก ,
(5.2)
In the present non-Markovian
setting the parametrization by the propensity to consume is crucial as one
cannot make statements for xed wealth. There is no immediate way to
infer results about
๐‘ห†,
except of course for the initial value
๐‘ห†0 = ๐œ…
ห† 0 ๐‘ฅ0 .
II.5.1 Variation of the Investment Opportunities
It is classical in economics to compare two identical agents with utility
function
๐‘ˆ,
where only one has access to a stock market. The opportunity
to invest in risky assets gives rise to two contradictory eects. The presence
of risk incites the agent to save cash for the uncertain future; this is the
precautionary savings eect and its strength is related to the absolute prudence P(๐‘ˆ ) = โˆ’๐‘ˆ โ€ฒโ€ฒโ€ฒ /๐‘ˆ โ€ฒโ€ฒ . On the other hand, the agent may prefer to invest
rather than to consume immediately. This substitution eect is related to
โ€ฒโ€ฒ
โ€ฒ
the absolute risk aversion A (๐‘ˆ ) = โˆ’๐‘ˆ /๐‘ˆ .
Classical economic theory (e.g., Gollier [27, Proposition 74]) states that
in a one period model, the presence of a complete nancial market makes
the optimal consumption at time
everywhere on
(0, โˆž),
๐‘ก = 0
smaller if
P(๐‘ˆ ) โ‰ฅ 2A (๐‘ˆ )
holds
and larger if the converse inequality holds. For power
utility, the former condition holds if
๐‘<0
and the latter holds if
๐‘ โˆˆ (0, 1).
We go a step further in the comparison by considering two dierent sets of
constraints, instead of giving no access to the stock market at all (which is
{0}).
C and C โ€ฒ be set-valued mappings of constraints as in Remark 3.7,
C โ€ฒ โŠ† C in the sense that C๐‘กโ€ฒ (๐œ”) โŠ† C๐‘ก (๐œ”) for all (๐‘ก, ๐œ”). Assume that
the constraint
Let
and let
there exist corresponding optimal constrained strategies.
Proposition 5.3. Let ๐œ…ห† and ๐œ…ห†โ€ฒ be the optimal propensities to consume for
the constraints C and C โ€ฒ , respectively. Then C โ€ฒ โŠ† C implies ๐œ…ห† โ‰ค ๐œ…ห† โ€ฒ if ๐‘ > 0
and ๐œ…ห† โ‰ฅ ๐œ…ห† โ€ฒ if ๐‘ < 0. In particular, ๐‘ห†0 โ‰ค ๐‘ห†โ€ฒ0 if ๐‘ > 0 and ๐‘ห†0 โ‰ฅ ๐‘ห†โ€ฒ0 if ๐‘ < 0.
22
II Opportunity Process
Proof.
Let
๐ฟ
and
๐ฟโ€ฒ
be the corresponding opportunity processes; we make
use of Remarks 3.7 and 5.2. Consider relation (5.2) with
โ€ฒ
and the analogue for ๐ฟ with
Cโ€ฒ
๐’œ
. We see that
Cโ€ฒ
๐’œ
โŠ†
instead of
๐’œC implies ๐‘1 ๐ฟโ€ฒ
as the supremum is taken over a larger set in the case of
decreasing function of
๐’œC
C.
By (5.1),
โ‰ค
๐œ…
ห†
๐’œ
1
๐‘ ๐ฟ,
is a
๐ฟ.
Proposition 5.4. The optimal propensity to consume satises
๐œ…
ห†๐‘ก โ‰ค
(๐‘˜2 /๐‘˜1 )๐›ฝ
1+๐‘‡ โˆ’๐‘ก
if ๐‘ โˆˆ (0, 1) and ๐œ…ห† ๐‘ก โ‰ฅ
(๐‘˜2 /๐‘˜1 )๐›ฝ
1+๐‘‡ โˆ’๐‘ก
if ๐‘ < 0.
In particular, we have a model-independent deterministic threshold independent of ๐‘ in the standard case ๐ท โ‰ก 1,
๐œ…
ห†๐‘ก โ‰ค
Proof.
1
1+๐‘‡ โˆ’๐‘ก
if ๐‘ โˆˆ (0, 1) and ๐œ…ห† ๐‘ก โ‰ฅ
1
1+๐‘‡ โˆ’๐‘ก
This follows from Lemma 3.5 and (5.1). The second part can also be
seen as special case of Proposition 5.3 with constraint set
then ๐œ…
ห†โ€ฒ
if ๐‘ < 0.
= (1 + ๐‘‡ โˆ’
๐‘ก)โˆ’1 as in Example 3.6.
C โ€ฒ = {0}
since
(1 + ๐‘‡ โˆ’ ๐‘ก)โˆ’1 coincides with the optimal propensity to
consume for the log-utility function (cf. [25]), which formally corresponds
to ๐‘ = 0. This suggests that the threshold is attained by ๐œ…
ห† (๐‘) in the limit
๐‘ โ†’ 0, a result we prove in Chapter V.
The threshold
Remark 5.5.
๐œ…
ห† opposite to the ones in Proposition 5.4
R๐‘ž (๐‘ƒ ) holds for the given nancial market model. Quan๐ถ๐‘ž > 0 is the constant for R๐‘ž (๐‘ƒ ), then
Uniform bounds for
exist if and only if
titatively, if
๐œ…
ห†๐‘ก โ‰ฅ
( ๐‘˜ )๐›ฝ 1
2
๐‘˜ 1 ๐ถ๐‘ž
if
๐‘ โˆˆ (0, 1)
and
๐œ…
ห†๐‘ก โ‰ค
( ๐‘˜ )๐›ฝ 1
1
๐‘˜2 ๐ถ๐‘ž
if
๐‘ < 0.
This follows from (5.1) and (2.4) by (the proof of ) Proposition 4.5. In view
of Corollary 4.12 we have the following dichotomy:
upper bound either for all values of
๐‘ < 0,
๐œ…
ห†=๐œ…
ห† (๐‘)
has a uniform
or for none of them.
II.5.2 Variation of ๐ท
We now study how
[๐‘ก1 , ๐‘ก2 ).
๐œ…
ห†
To this end, let
is aected if we increase
0 โ‰ค ๐‘ก1 < ๐‘ก2 โ‰ค ๐‘‡
๐ท
on some time interval
be two xed points in time and
๐œ‰
a
bounded càdlàg adapted process which is strictly positive and nonincreasing
on
[๐‘ก1 , ๐‘ก2 ).
In addition to
๐‘ˆ๐‘ก (๐‘ฅ) = ๐ท๐‘ก ๐‘1 ๐‘ฅ๐‘
๐‘ˆ๐‘กโ€ฒ (๐‘ฅ) := ๐ท๐‘กโ€ฒ ๐‘1 ๐‘ฅ๐‘ ,
we consider the utility random eld
(
)
๐ทโ€ฒ := 1 + ๐œ‰1[๐‘ก1 ,๐‘ก2 ) ๐ท.
As an interpretation, recall the modeling of taxation by
๐ท
from Re-
mark 2.2. Then we want to nd out how the agent reacts to a temporary
II.6 Appendix A: Dynamic Programming
change of the tax policy on
tax rate
๐œš := ๐ทโˆ’1/๐‘ โˆ’ 1
[๐‘ก1 , ๐‘ก2 )in
[๐‘ก1 , ๐‘ก2 ),
๐‘ > 0,
the next result
while the contrary holds before the pol-
icy change and there is no eect after
opposite way.
particular whether a reduction of the
stimulates consumption. For
shows this to be true during
23
๐‘ก2 .
An agent with
๐‘<0
reacts in the
Remark 2.2 also suggests other interpretations of the same
result.
Proposition 5.6. Let ๐œ…ห† and
๐‘ˆ and ๐‘ˆ โ€ฒ , respectively. Then
๐œ…
ห†โ€ฒ
be the optimal propensities to consume for
โŽง
๏ฃด
ห† โ€ฒ๐‘ก < ๐œ…
ห†๐‘ก
โŽจ๐œ…
โ€ฒ
๐œ…
ห†๐‘ก > ๐œ…
ห†๐‘ก
๏ฃด
โŽฉ โ€ฒ
๐œ…
ห†๐‘ก = ๐œ…
ห†๐‘ก
Proof.
๐‘ˆ and ๐‘ˆ โ€ฒ . We conโ€ฒ
โ€ฒ
sider (5.2) and compare it with its analogue for ๐ฟ , where ๐ท is replaced by ๐ท .
โ€ฒ
โ€ฒ
As ๐œ‰ > 0, we then see that ๐ฟ๐‘ก > ๐ฟ๐‘ก for ๐‘ก < ๐‘ก1 ; moreover, ๐ฟ๐‘ก = ๐ฟ๐‘ก for ๐‘ก โ‰ฅ ๐‘ก2 .
โ€ฒ
Since ๐œ‰ is nonincreasing, we also see that ๐ฟ๐‘ก < (1+๐œ‰๐‘ก )๐ฟ๐‘ก for ๐‘ก โˆˆ [๐‘ก1 , ๐‘ก2 ). It remains to apply (5.1). For ๐‘ก < ๐‘ก1 , ๐œ…
ห† โ€ฒ = (๐ท๐‘กโ€ฒ /๐ฟโ€ฒ๐‘ก )๐›ฝ = (๐ท๐‘ก /๐ฟโ€ฒ๐‘ก )๐›ฝ < (๐ท๐‘ก /๐ฟ๐‘ก )๐›ฝ = ๐œ…
ห†.
For ๐‘ก โˆˆ [๐‘ก1 , ๐‘ก2 ) we have
( (1 + ๐œ‰ )๐ท )๐›ฝ ( (1 + ๐œ‰ )๐ท )๐›ฝ
๐‘ก
๐‘ก
๐‘ก
๐‘ก
๐œ…
ห† โ€ฒ = (๐ท๐‘กโ€ฒ /๐ฟโ€ฒ๐‘ก )๐›ฝ =
>
=๐œ…
ห†,
๐ฟโ€ฒ๐‘ก
(1 + ๐œ‰๐‘ก )๐ฟ๐‘ก
Let
while for
๐ฟ
and
๐ฟโ€ฒ
if ๐‘ก < ๐‘ก1 ,
if ๐‘ก โˆˆ [๐‘ก1 , ๐‘ก2 ),
if ๐‘ก โ‰ฅ ๐‘ก2 .
be the opportunity processes for
๐‘ก โ‰ฅ ๐‘ก2 , ๐ท๐‘กโ€ฒ = ๐ท๐‘ก
Remark 5.7.
(i) For ๐‘ก2
implies
๐œ…
ห† โ€ฒ๐‘ก = ๐œ…
ห†๐‘ก.
= ๐‘‡ , the statement of Proposition 5.6 remains true
หœ.
๐ท
if the closed interval is chosen in the denition of
(ii) One can see [72, Proposition 12] as a special case of Proposition 5.6.
๐ท = 1[0,๐‘‡ ) ๐พ1 + 1{๐‘‡ } ๐พ2 for two con๐พ1 , ๐พ2 > 0 and obtain monotonicity of the consumption with respect
the ratio ๐พ2 /๐พ1 . This is proved in a Markovian setting by a comparison
In our notation, the authors consider
stants
to
result for PDEs.
II.6 Appendix A: Dynamic Programming
This appendix collects the facts about dynamic programming which are used
in this chapter.
Recall the standing assumption (2.5), the set
from (3.1) and the process
Fact 6.1.
๐ฝ
๐’œ(๐œ‹, ๐‘, ๐‘ก)
from (3.2). We begin with the lattice property.
โˆซ๐‘‡
(๐œ‹, ๐‘) โˆˆ ๐’œ and let ฮ“๐‘ก (หœ
๐‘) := ๐ธ[ ๐‘ก ๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โˆฃโ„ฑ๐‘ก ]. The set
{ฮ“๐‘ก (หœ
๐‘) : ๐‘หœ โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก)} is upward ltering for each ๐‘ก โˆˆ [0, ๐‘‡ ].
๐‘– ๐‘–
1
2
3
Indeed, if (๐œ‹ , ๐‘ ) โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก), ๐‘– = 1, 2, we have ฮ“๐‘ก (๐‘ ) โˆจ ฮ“๐‘ก (๐‘ ) = ฮ“๐‘ก (๐‘ )
3
3
1
1
2
2
1
2
for (๐œ‹ , ๐‘ ) := (๐œ‹ , ๐‘ )1๐ด + (๐œ‹ , ๐‘ )1๐ด๐‘ with ๐ด := {ฮ“๐‘ก (๐‘ ) > ฮ“๐‘ก (๐‘ )}. Clearly
(๐œ‹ 3 , ๐‘3 ) โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก). Regarding Remark 3.7, we note that ๐œ‹ 3 satises the
1
2
constraints if ๐œ‹ and ๐œ‹ do.
Fix
II Opportunity Process
24
Proposition 6.2. Let (๐œ‹, ๐‘) โˆˆ ๐’œ and ๐ผ๐‘ก (๐œ‹, ๐‘) := ๐ฝ๐‘ก (๐œ‹, ๐‘) + 0๐‘ก ๐‘ˆ๐‘  (๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ).
If ๐ธ[ โˆฃ๐ผ๐‘ก (๐œ‹, ๐‘)โˆฃ ] < โˆž for each ๐‘ก, then ๐ผ(๐œ‹, ๐‘) is a supermartingale having a
càdlàg version. It is a martingale if and only if (๐œ‹, ๐‘) is optimal.
โˆซ
Proof.
The technique of proof is well known; see El Karoui and Quenez [46]
or Laurent and Pham [52] for arguments in dierent contexts.
0 โ‰ค ๐‘ก โ‰ค ๐‘ข โ‰ค ๐‘‡ and prove the supermartin๐ผ๐‘ก (๐œ‹, ๐‘) = ess sup๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก) ฮฅ๐‘ก (หœ
๐‘) for the martingale
]
[โˆซ๐‘‡
โˆ˜
ฮฅ๐‘ก (หœ
๐‘) := ๐ธ 0 ๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ ) โ„ฑ๐‘ก . (More precisely, the expectation is well
dened with values in โ„ โˆช {โˆ’โˆž} by (2.5).)
โˆซ๐‘ข
As ฮฅ๐‘ข (หœ
๐‘) = ฮ“๐‘ข (หœ
๐‘) + 0 ๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ), Fact 6.1 implies that there exists
๐‘›
๐‘›
a sequence (๐‘ ) in ๐’œ(๐œ‹, ๐‘, ๐‘ข) such that lim๐‘› ฮฅ๐‘ข (๐‘ ) = ๐ผ๐‘ข (๐œ‹, ๐‘) ๐‘ƒ -a.s., where
the limit is monotone increasing in ๐‘›. We conclude that
We x
(๐œ‹, ๐‘) โˆˆ ๐’œ
as well as
gale property. Note that
๐ธ[๐ผ๐‘ข (๐œ‹, ๐‘)โˆฃโ„ฑ๐‘ก ] = ๐ธ[lim ฮฅ๐‘ข (๐‘๐‘› )โˆฃโ„ฑ๐‘ก ] = lim ๐ธ[ฮฅ๐‘ข (๐‘๐‘› )โˆฃโ„ฑ๐‘ก ]
๐‘›
๐‘›
โ‰ค ess sup๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ข) ๐ธ[ฮฅ๐‘ข (หœ
๐‘)โˆฃโ„ฑ๐‘ก ] = ess sup๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ข) ฮฅ๐‘ก (หœ
๐‘)
โ‰ค ess sup๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก) ฮฅ๐‘ก (หœ
๐‘) = ๐ผ๐‘ก (๐œ‹, ๐‘).
To construct the càdlàg version, denote by
taking the right limits of
with
๐ผ๐‘‡โ€ฒ := ๐ผ๐‘‡ .
Since
๐ผ
๐‘ก 7โ†’ ๐ผ๐‘ก (๐œ‹, ๐‘) =: ๐ผ๐‘ก
the process obtained by
through the rational numbers,
is a supermartingale and the ltration satises the
usual assumptions, these limits exist
โ€ฒ
gale, and ๐ผ๐‘ก
๐ผโ€ฒ
๐‘ƒ -a.s., ๐ผ โ€ฒ
is a (càdlàg) supermartin-
โ‰ค ๐ผ๐‘ก ๐‘ƒ -a.s. (see Dellacherie and Meyer [18, VI.1.2]).
(หœ
๐œ‹ , ๐‘หœ) โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก) we have
But in fact,
equality holds here because for all
ฮฅ๐‘ก (หœ
๐‘) = ๐ธ
๐‘‡
[โˆซ
0
due to
๐ผ๐‘‡ = ๐ผ๐‘‡โ€ฒ ,
]
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐‘‘๐œ‡โˆ˜ โ„ฑ๐‘ก = ๐ธ[๐ผ๐‘‡ (หœ
๐œ‹ , ๐‘หœ)โˆฃโ„ฑ๐‘ก ] = ๐ธ[๐ผ๐‘‡ โˆฃโ„ฑ๐‘ก ] โ‰ค ๐ผ๐‘กโ€ฒ
and hence also
is a càdlàg version of
๐ผ๐‘กโ€ฒ โ‰ฅ ess sup๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก) ฮฅ๐‘ก (หœ
๐‘) = ๐ผ๐‘ก .
Therefore
๐ผโ€ฒ
๐ผ.
(๐œ‹, ๐‘) be optimal. Then ๐ผ0 (๐œ‹, ๐‘) =
๐ผ(๐œ‹, ๐‘) is a martingale. Conversely, this relation states that (๐œ‹, ๐‘) is optimal, by denition of ๐ผ(๐œ‹, ๐‘).
Turning to the martingale property, let
ฮฅ0 (๐œ‹, ๐‘) = ๐ธ[๐ผ๐‘‡ (๐œ‹, ๐‘)],
so the supermartingale
The following property was used in the body of the text.
Fact 6.3.
time
(๐œ‹, ๐‘), (๐œ‹ โ€ฒ , ๐‘โ€ฒ ) โˆˆ ๐’œ with corresponding
(๐œ‹ โ€ฒโ€ฒ , ๐‘โ€ฒโ€ฒ ) โˆˆ ๐’œ(๐œ‹ โ€ฒ , ๐‘โ€ฒ , ๐‘ก). Then
Consider
๐‘ก โˆˆ [0, ๐‘‡ ]
and
๐‘1[0,๐‘ก] +
process is
๐‘‹1[0,๐‘ก] +
๐‘‹๐‘ก , ๐‘‹๐‘กโ€ฒ
at
๐‘‹๐‘ก โ€ฒโ€ฒ
๐‘ 1(๐‘ก,๐‘‡ ] โˆˆ ๐’œ(๐œ‹, ๐‘, ๐‘ก).
๐‘‹๐‘กโ€ฒ
๐œ‹1[0,๐‘ก] + ๐œ‹ โ€ฒโ€ฒ 1(๐‘ก,๐‘‡ ] ,
> 0 by (2.1).
Indeed, for the trading strategy
๐‘‹๐‘ก โ€ฒโ€ฒ
๐‘‹ 1(๐‘ก,๐‘‡ ]
๐‘‹๐‘กโ€ฒ
wealths
the corresponding wealth
II.7 Appendix B: Martingale Property of the Optimal Processes
25
II.7 Appendix B: Martingale Property of the Optimal Processes
The purpose of this appendix is to provide a statement which follows from [43]
and is known to its authors, but which we could not nd in the literature.
For the case without intermediate consumption, the following assertion is
contained in [49, Theorem 2.2].
Lemma 7.1. Assume
๐‘Œ โˆˆ Y D , then
(2.4)
and
. Let (๐œ‹, ๐‘) โˆˆ ๐’œ, ๐‘‹ = ๐‘‹(๐œ‹, ๐‘) and
(2.6)
๐‘ก
โˆซ
๐‘๐‘  ๐‘Œ๐‘  ๐œ‡(๐‘‘๐‘ ),
๐‘๐‘ก := ๐‘‹๐‘ก ๐‘Œ๐‘ก +
๐‘ก โˆˆ [0, ๐‘‡ ]
0
ห† ๐‘ห†, ๐‘Œห† ) are the optimal processes solvis a supermartingale. If (๐‘‹, ๐‘, ๐‘Œ ) = (๐‘‹,
ing the primal and the dual problem, respectively, then ๐‘ is a martingale.
Proof.
It follows from [43, Theorem 3.10(vi)] that
๐ธ[๐‘๐‘‡ ] = ๐ธ[๐‘0 ]
for the
optimal processes, so it suces to prove the rst part.
๐‘Œ โˆˆ Y M , i.e., ๐‘Œ /๐‘Œ0 โˆซis the density process
of a
โˆซ
M
โˆ—
measure ๐‘„ โ‰ˆ ๐‘ƒ . As Y
โŠ† Y , the process ๐‘‹ + ๐‘๐‘ข ๐œ‡(๐‘‘๐‘ข) = ๐‘ฅ0 + ๐‘‹โˆ’ ๐œ‹ ๐‘‘๐‘…
โˆซ๐‘ 
โˆซ๐‘ก
๐‘„
is a ๐‘„-supermartingale, that is, ๐ธ [๐‘‹๐‘ก +
0 ๐‘๐‘ข ๐œ‡(๐‘‘๐‘ข)โˆฃโ„ฑ๐‘  ] โ‰ค ๐‘‹๐‘  + 0 ๐‘๐‘ข ๐œ‡(๐‘‘๐‘ข)
for ๐‘  โ‰ค ๐‘ก. We obtain the claim by Bayes' rule,
โˆซ ๐‘ก
]
[
๐ธ ๐‘‹๐‘ก ๐‘Œ๐‘ก +
๐‘๐‘ข ๐‘Œ๐‘ข ๐œ‡(๐‘‘๐‘ข)โ„ฑ๐‘  โ‰ค ๐‘‹๐‘  ๐‘Œ๐‘  .
(i)
Assume rst that
๐‘ 
(ii) Let
๐‘Œ โˆˆYD
be arbitrary. By [43, Corollary 2.11], there is a sequence
๐‘Œ โˆˆ Y M which Fatou-converges to ๐‘Œ . Consider the supermartingale ๐‘Œ โ€ฒ :=
lim inf ๐‘› ๐‘Œ ๐‘› . By ยšitkovi¢ [75, Lemma 8], ๐‘Œ๐‘กโ€ฒ = ๐‘Œ๐‘ก ๐‘ƒ -a.s. for all ๐‘ก in a (dense)
๐‘›
subset
ฮ› โŠ† [0, ๐‘‡ ]
which contains
๐‘‡
and whose complement is countable. It
๐‘ is a supermartingale on ฮ›;
๐‘  โ‰ค ๐‘ก in ฮ›,
โˆซ ๐‘ก
โˆซ ๐‘ก
]
]
[
[
โ€ฒ
๐‘๐‘ข ๐‘Œ๐‘ข ๐œ‡(๐‘‘๐‘ข)โ„ฑ๐‘  = ๐ธ ๐‘‹๐‘ก ๐‘Œ๐‘ก +
๐‘๐‘ข ๐‘Œ๐‘ขโ€ฒ ๐œ‡(๐‘‘๐‘ข)โ„ฑ๐‘ 
๐ธ ๐‘‹๐‘ก ๐‘Œ๐‘ก +
๐‘ 
๐‘ 
โˆซ ๐‘ก
]
[
๐‘›
๐‘›
โ‰ค lim inf ๐ธ ๐‘‹๐‘ก ๐‘Œ๐‘ก +
๐‘๐‘ข ๐‘Œ๐‘ข ๐œ‡(๐‘‘๐‘ข)โ„ฑ๐‘ 
follows from Fatou's lemma and step (i) that
indeed, for
๐‘›
๐‘ 
โ‰ค lim inf ๐‘‹๐‘  ๐‘Œ๐‘ ๐‘› = ๐‘‹๐‘  ๐‘Œ๐‘ 
๐‘›
[0, ๐‘‡ ] by taking right limits in ฮ› and obtain a rightโ€ฒ
continuous supermartingale ๐‘ on [0, ๐‘‡ ], by right-continuity of the ltration.
โ€ฒ
But ๐‘ is indistinguishable from ๐‘ because ๐‘ is also right-continuous. Hence
๐‘ is a supermartingale as claimed.
We can extend
๐‘โˆฃฮ›
๐‘ƒ -a.s.
to
26
II Opportunity Process
Chapter III
Bellman Equation
In this chapter, which corresponds to the article [59], we consider a general
semimartingale model with closed portfolio constraints.
We focus on the
local representation of the optimization problem, which is formalized by the
Bellman equation.
III.1 Introduction
This chapter presents the local dynamic programming for power utility maximization in a general constrained semimartingale framework. We have seen
that the homogeneity of these utility functions leads to a factorization of
the value process into a power of the current wealth and the opportunity
process
๐ฟ.
In our setting, the Bellman equation describes the drift rate of
๐ฟ
and claries the local structure of our problem. Finding an optimal strategy
boils down to maximizing a random function
state
๐œ”
๐‘ก.
๐ฟ
๐ฟ.
๐‘ฆ 7โ†’ ๐‘”(๐œ”, ๐‘ก, ๐‘ฆ)
on
โ„๐‘‘
for every
and date
This function is given in terms of the semimartingale
characteristics of
as well as the asset returns, and its maximum yields
the drift rate of
The role of the opportunity process is to augment the
information contained in the return characteristics in order to have a local
sucient statistic for the global optimization problem.
We present three main results. First, we show that if there exists an opti-
the opportunity process ๐ฟ
solves the Bellman equation and we provide a local description of the optimal
mal strategy for the utility maximization problem,
strategies. We state the Bellman equation in two forms, as an identity for the
drift rate of
๐ฟ.
๐ฟ and as a backward stochastic dierential equation (BSDE) for
Second, we characterize the opportunity process as the
minimal solution
of this equation. Finally, given some solution and an associated strategy, one
can ask whether the strategy is optimal and the solution is the opportunity
process. We present two dierent approaches which lead to two
theorems
verication
not comparable in strength unless the constraints are convex.
The present dynamic programming approach should be seen as comple-
28
III Bellman Equation
mentary to convex duality, which remains the only method to obtain
tence
exis-
of optimal strategies in general models; see Kramkov and Schacher-
mayer [49], Karatzas and ยšitkovi¢ [43], Karatzas and Kardaras [41]. In some
cases the Bellman equation can be solved directly, e.g., in the setting of
Example 5.8 with continuous asset prices or in the Lévy process setting of
Chapter IV. In addition to existence, one then typically obtains additional
properties of the optimal strategies.
This chapter is organized as follows. The next section species the optimization problem in detail, recalls the opportunity process and the martingale optimality principle, and xes the notation for the characteristics.
We also introduce set-valued processes describing the budget condition and
state the assumptions on the portfolio constraints. Section III.3 derives the
Bellman equation, rst as a drift condition and then as a BSDE. It becomes
more explicit as we specialize to the case of continuous asset prices. The definition of a solution of the Bellman equation is given in Section III.4, where
we show the minimality of the opportunity process. Section III.5 deals with
the verication problem, which is converse to the derivation of the Bellman
equation since it requires the passage from the local maximization to the
global optimization problem.
We present an approach via the value pro-
cess and a second approach via a deator, which corresponds to the dual
problem in a suitable setting.
Appendix A is linked to Section III.3 and
contains the measurable selections for the construction of the Bellman equation. It is complemented by Appendix B, where we construct an alternative
parametrization of the market model by representative portfolios.
III.2 Preliminaries
๐‘ฅ, ๐‘ฆ โˆˆ โ„ are reals, ๐‘ฅ+ = max{๐‘ฅ, 0} and
๐‘ฅ โˆง ๐‘ฆ = min{๐‘ฅ, ๐‘ฆ}. We set 1/0 := โˆž where necessary. If ๐‘ง โˆˆ โ„๐‘‘ is a
๐‘‘-dimensional vector, ๐‘ง ๐‘– is its ๐‘–th coordinate, ๐‘ง โŠค its transpose, and โˆฃ๐‘งโˆฃ =
(๐‘ง โŠค ๐‘ง)1/2 the Euclidean norm. If ๐‘‹ is an โ„๐‘‘ -valued semimartingale and
๐œ‹ is an โ„๐‘‘ -valued predictable integrand, the vector stochasticโˆซ integral is a
scalar semimartingale with initial value zero and denoted by
๐œ‹ ๐‘‘๐‘‹ or by
๐œ‹ โˆ™ ๐‘‹ . The quadratic variation of ๐‘‹ is the ๐‘‘ × ๐‘‘-matrix [๐‘‹] := [๐‘‹, ๐‘‹]
and if ๐‘Œ is a scalar semimartingale, [๐‘‹, ๐‘Œ ] is the ๐‘‘-vector which is given
๐‘–
๐‘–
by [๐‘‹, ๐‘Œ ] := [๐‘‹ , ๐‘Œ ]. Relations between measurable functions hold almost
The following notation is used. If
everywhere unless otherwise mentioned. Our reference for any unexplained
notion from stochastic calculus is Jacod and Shiryaev [34].
III.2.1 The Optimization Problem
๐‘‡ โˆˆ (0, โˆž) and a stochastic basis (ฮฉ, โ„ฑ, ๐”ฝ, ๐‘ƒ ),
๐”ฝ = (โ„ฑ๐‘ก )๐‘กโˆˆ[0,๐‘‡ ] satises the usual assumptions of rightcompleteness as well as โ„ฑ0 = {โˆ…, ฮฉ} ๐‘ƒ -a.s. We consider an
We x the time horizon
where the ltration
continuity and
III.2 Preliminaries
29
โ„๐‘‘ -valued càdlàg semimartingale ๐‘… with ๐‘…0 = 0 representing the returns of ๐‘‘
risky assets. Their discounted prices are given by the stochastic exponential
๐‘† = โ„ฐ(๐‘…) = (โ„ฐ(๐‘…1 ), . . . , โ„ฐ(๐‘…๐‘‘ )).
Our agent also has a bank account at his
disposal; it does not pay interest.
The agent is endowed with a deterministic initial capital
trading strategy
๐œ‹๐‘–
is a predictable
๐‘…-integrable โ„๐‘‘ -valued
๐‘ฅ0 > 0. A
๐œ‹ , where
process
indicates the fraction of wealth (or the portfolio proportion) invested
๐‘–th risky asset.
โˆซ ๐‘‡ A consumption strategy is a nonnegative optional
๐‘ such that 0 ๐‘๐‘ก ๐‘‘๐‘ก < โˆž ๐‘ƒ -a.s. We want to consider two cases.
Either consumption occurs only at the terminal time ๐‘‡ (utility from terminal
in the
process
wealth only); or there is intermediate consumption plus a bulk consumption
at the time horizon. To unify the notation, we introduce the measure
[0, ๐‘‡ ]
๐œ‡
on
by
{
0
๐œ‡(๐‘‘๐‘ก) :=
๐‘‘๐‘ก
in the case without intermediate consumption,
in the case with intermediate consumption.
๐œ‡โˆ˜ := ๐œ‡+๐›ฟ{๐‘‡ } , where ๐›ฟ{๐‘‡ } is the unit Dirac measure at ๐‘‡ . The wealth
process ๐‘‹(๐œ‹, ๐‘) corresponding to a pair (๐œ‹, ๐‘) is dened by the equation
โˆซ ๐‘ก
โˆซ ๐‘ก
๐‘‹๐‘ก (๐œ‹, ๐‘) = ๐‘ฅ0 +
๐‘‹๐‘ โˆ’ (๐œ‹, ๐‘)๐œ‹๐‘  ๐‘‘๐‘…๐‘  โˆ’
๐‘๐‘  ๐œ‡(๐‘‘๐‘ ), 0 โ‰ค ๐‘ก โ‰ค ๐‘‡.
Let also
0
0
We dene the set of trading and consumption pairs
{
๐’œ0 (๐‘ฅ0 ) := (๐œ‹, ๐‘) : ๐‘‹(๐œ‹, ๐‘) > 0, ๐‘‹โˆ’ (๐œ‹, ๐‘) > 0
and
}
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘) .
These are the strategies that satisfy the budget constraint. The convention
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘) means that all the remaining wealth is consumed at time
๐‘‡ . We consider also exogenous constraints imposed on the agent. For each
(๐œ”, ๐‘ก) โˆˆ ฮฉ × [0, ๐‘‡ ] we are given a set C๐‘ก (๐œ”) โŠ† โ„๐‘‘ which contains the origin.
The set of (constrained)
admissible
strategies is
{
๐’œ(๐‘ฅ0 ) := (๐œ‹, ๐‘) โˆˆ ๐’œ0 (๐‘ฅ0 ) : ๐œ‹๐‘ก (๐œ”) โˆˆ C๐‘ก (๐œ”) for
it is nonempty as
C
0 โˆˆ C๐‘ก (๐œ”).
all
}
(๐œ”, ๐‘ก) ;
Further assumptions on the set-valued mapping
We x the initial capital ๐‘ฅ0 and
๐’œ(๐‘ฅ0 ). We write ๐‘ โˆˆ ๐’œ and call ๐‘ admissible if there
(๐œ‹, ๐‘) โˆˆ ๐’œ; an analogous convention is used for similar
will be introduced in Section III.2.4.
usually write
exists
๐œ‹
๐’œ
for
such that
expressions.
We will often parametrize the consumption strategies as a fraction of
wealth. Let
(๐œ‹, ๐‘) โˆˆ ๐’œ
and
๐‘‹ = ๐‘‹(๐œ‹, ๐‘).
๐œ… :=
is called the
Then
๐‘
๐‘‹
propensity to consume corresponding to (๐œ‹, ๐‘).
to-one correspondence between the pairs
This yields a one-
(๐œ‹, ๐‘) โˆˆ ๐’œ and the pairs (๐œ‹, ๐œ…) such
30
III Bellman Equation
that
๐œ‹โˆˆ๐’œ
๐‘ƒ -a.s.
and
โˆซ๐‘‡
๐œ… is a nonnegative optional process satisfying 0 ๐œ…๐‘  ๐‘‘๐‘  < โˆž
๐œ…๐‘‡ = 1 (see Remark II.2.1 for details). We shall abuse the notaand
tion and identify a consumption strategy with the corresponding propensity
(๐œ‹, ๐œ…) โˆˆ ๐’œ. Note that
(
)
๐‘‹(๐œ‹, ๐œ…) = ๐‘ฅ0 โ„ฐ ๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡ .
to consume, e.g., we write
This simplies verifying that some pair
implies
๐‘‹โˆ’ (๐œ‹, ๐œ…) > 0
(๐œ‹, ๐œ…)
is admissible as
๐‘‹(๐œ‹, ๐œ…) > 0
(cf. [34, II.8a]).
The preferences of the agent are modeled by a time-additive random
Let ๐ท be a càdlàg, adapted, strictly positive
]
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ ) < โˆž and x ๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1). We
utility function as follows.
process such that
๐ธ
[โˆซ๐‘‡
0
dene the power utility random eld
๐‘ˆ๐‘ก (๐‘ฅ) := ๐ท๐‘ก ๐‘1 ๐‘ฅ๐‘ ,
This is the general form of a
๐‘ฅ โˆˆ (0, โˆž), ๐‘ก โˆˆ [0, ๐‘‡ ].
๐‘-homogeneous
utility random eld such that
a constant consumption yields nite expected utility.
Interpretations and
๐ท are discussed in Chapter II.
๐‘ฅ 7โ†’ ๐‘ˆ๐‘ก (๐‘ฅ),
applications for the process
the convex conjugate of
We denote by
{
}
๐‘ˆ๐‘กโˆ— (๐‘ฆ) = sup ๐‘ˆ๐‘ก (๐‘ฅ) โˆ’ ๐‘ฅ๐‘ฆ = โˆ’ 1๐‘ž ๐‘ฆ ๐‘ž ๐ท๐‘ก๐›ฝ ;
๐‘ˆโˆ—
(2.1)
๐‘ฅ>0
๐‘
๐‘ž := ๐‘โˆ’1
โˆˆ (โˆ’โˆž, 0) โˆช (0, 1) is the exponent conjugate to ๐‘
1
constant ๐›ฝ :=
1โˆ’๐‘ > 0 is the relative risk tolerance of ๐‘ˆ . Note
here
and the
that we
exclude the well-studied logarithmic utility (e.g., Goll and Kallsen [25]) which
corresponds to
The
๐ธ
[โˆซ๐‘‡
๐‘ = 0.
expected utility
]
๐ธ[๐‘ˆ๐‘‡ (๐‘๐‘‡ )]
utility maximization problem is said to be
๐‘ข(๐‘ฅ0 ) := sup ๐ธ
[โˆซ
Note that this condition is void if
(๐œ‹, ๐‘) โˆˆ ๐’œ(๐‘ฅ0 )
is called
๐‘‡
or
nite
if
]
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) < โˆž.
(2.2)
0
๐‘โˆˆ๐’œ(๐‘ฅ0 )
strategy
๐‘ โˆˆ ๐’œ is
โˆซ๐‘‡
๐ธ[ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ๐‘‡ (๐‘๐‘‡ )]. The
corresponding to a consumption strategy
โˆ˜
0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡ (๐‘‘๐‘ก) , i.e., either
๐‘<0
optimal
if
๐‘ˆ < 0. If (2.2) holds,
]
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) = ๐‘ข(๐‘ฅ0 ).
as then
๐ธ
[โˆซ๐‘‡
0
a
Finally, we introduce the following sets; they are of minor importance
and used only in the case
๐‘ < 0:
{
}
โˆซ๐‘‡
๐’œ๐‘“ := (๐œ‹, ๐‘) โˆˆ ๐’œ : 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) > โˆ’โˆž ,
{
[โˆซ ๐‘‡
]
}
๐’œ๐‘“ ๐ธ := (๐œ‹, ๐‘) โˆˆ ๐’œ : ๐ธ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) > โˆ’โˆž .
Anticipating that (2.2) will be in force, the indices stand for nite and
nite expectation. Clearly
๐’œ๐‘“ ๐ธ โŠ† ๐’œ๐‘“ โŠ† ๐’œ,
and equality holds if
๐‘ โˆˆ (0, 1).
III.2 Preliminaries
31
III.2.2 Opportunity Process
We recall the opportunity process. We assume (2.2) in this section, which
ensures that the following process is nite. By Proposition II.3.1 and Remark II.3.7 there exists a unique càdlàg semimartingale
process, such that
๐ฟ๐‘ก ๐‘1
(
[
)๐‘
๐‘‹๐‘ก (๐œ‹, ๐‘) = ess sup ๐ธ
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
โˆซ
๐‘ก
๐‘‡
๐ฟ, called opportunity
]
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
(2.3)
{
}
(๐œ‹, ๐‘) โˆˆ ๐’œ, where ๐’œ(๐œ‹, ๐‘, ๐‘ก) := (หœ
๐œ‹ , ๐‘หœ) โˆˆ ๐’œ : (หœ
๐œ‹ , ๐‘หœ) = (๐œ‹, ๐‘) on [0, ๐‘ก] .
1 ๐‘
note that ๐ฟ๐‘‡ = ๐ท๐‘‡ and that ๐‘ข(๐‘ฅ0 ) = ๐ฟ0 ๐‘ฅ0 is the value function
๐‘
for any
We
from (2.2). The following is contained in Lemma II.3.5.
Lemma 2.1. ๐ฟ is a special semimartingale for all ๐‘. If ๐‘ โˆˆ (0, 1), then
๐ฟ, ๐ฟโˆ’ > 0. If ๐‘ < 0, the same holds provided that an optimal strategy exists.
Proposition 2.2
. Let (๐œ‹, ๐‘) โˆˆ ๐’œ๐‘“ ๐ธ . Then the process
(Proposition II.3.4)
๐ฟ๐‘ก ๐‘1
(
)๐‘
๐‘‹๐‘ก (๐œ‹, ๐‘) +
โˆซ
๐‘ก
๐‘ˆ๐‘  (๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ),
๐‘ก โˆˆ [0, ๐‘‡ ]
0
is a supermartingale; it is a martingale if and only if (๐œ‹, ๐‘) is optimal.
This is the martingale optimality principle.
โˆซ๐‘‡
๐ธ[ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก)],
value of this process equals
(๐œ‹, ๐‘) โˆˆ ๐’œ โˆ– ๐’œ๐‘“ ๐ธ .
The expected terminal
hence the assertion fails for
III.2.3 Semimartingale Characteristics
In the remainder of this section we introduce tools which are necessary to
describe the optimization problem locally. The use of semimartingale characteristics and set-valued processes follows [25] and [41], which consider logarithmic utility and convex constraints. That problem diers from ours in
that it is myopic, i.e., the characteristics of
๐‘…
are sucient to describe the
local problem and so there is no opportunity process.
We refer to [34] for background regarding semimartingale characteristics
and random measures. Let
๐œ‡๐‘…
be the integer-valued random measure associ-
โ„Ž : โ„๐‘‘ โ†’ โ„๐‘‘ be a cut-o function, i.e., โ„Ž is
๐‘…
๐‘… ๐‘…
bounded and โ„Ž(๐‘ฅ) = ๐‘ฅ in a neighborhood of ๐‘ฅ = 0. Let (๐ต , ๐ถ , ๐œˆ ) be the
predictable characteristics of ๐‘… relative to โ„Ž. The canonical representation
of ๐‘… (cf. [34, II.2.35]) is
ated with the jumps of
๐‘…
and let
๐‘… = ๐ต ๐‘… + ๐‘…๐‘ + โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) + (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘… .
The nite variation process
jumps of
๐‘….
(๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘…
(2.4)
contains essentially the large
The rest is the canonical decomposition of the semimartingale
III Bellman Equation
32
¯ = ๐‘… โˆ’ (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘… ,
๐‘…
which has bounded jumps:
๐ต ๐‘… = ๐ต ๐‘… (โ„Ž)
is
๐‘
predictable of nite variation, ๐‘… is a continuous local martingale, and nally
โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) is a purely discontinuous local martingale.
As
๐ฟ
is a special semimartingale (Lemma 2.1), it has a canonical de-
๐ฟ = ๐ฟ 0 + ๐ด๐ฟ + ๐‘€ ๐ฟ .
composition
Here
of nite variation and also called the
and
๐ฟ
๐ด๐ฟ
0 = ๐‘€0 = 0 .
๐ฟ0
drift
is constant,
๐ด๐ฟ
is predictable
๐ฟ is a local martingale,
of ๐ฟ, ๐‘€
Analogous notation will be used for other special semi-
(๐ด๐ฟ , ๐ถ ๐ฟ , ๐œˆ ๐ฟ )
โ€ฒ
Writing ๐‘ฅ for
martingales. It is then possible to consider the characteristics
of
๐ฟ
with respect to the identity instead of a cut-o function.
the identity on
โ„,
the canonical representation is
๐ฟ = ๐ฟ0 + ๐ด๐ฟ + ๐ฟ๐‘ + ๐‘ฅโ€ฒ โˆ— (๐œ‡๐ฟ โˆ’ ๐œˆ ๐ฟ );
see [34, II.2.38]. It will be convenient to use the joint characteristics of the
โ„๐‘‘ × โ„-valued process (๐‘…, ๐ฟ). We denote a generic point in โ„๐‘‘ × โ„ by (๐‘ฅ, ๐‘ฅโ€ฒ )
๐‘…,๐ฟ , ๐ถ ๐‘…,๐ฟ , ๐œˆ ๐‘…,๐ฟ ) be the characteristics of (๐‘…, ๐ฟ) with respect to the
and let (๐ต
โ€ฒ
โ€ฒ
function (๐‘ฅ, ๐‘ฅ ) 7โ†’ (โ„Ž(๐‘ฅ), ๐‘ฅ ). More precisely, we choose good versions of
the characteristics so that they satisfy the properties given in [34, II.2.9]. For
(๐‘‘ + 1)-dimensional process (๐‘…, ๐ฟ) we have the canonical representation
)
)
(
( ) ( ) ( ๐‘…) ( ๐‘) (
๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)
โ„Ž(๐‘ฅ)
๐‘…
๐ต
0
๐‘…
๐‘…,๐ฟ
๐‘…,๐ฟ
โˆ—๐œ‡๐‘…,๐ฟ .
โˆ—(๐œ‡ โˆ’๐œˆ )+
+
+
+
=
0
๐‘ฅโ€ฒ
๐ฟ๐‘
๐ด๐ฟ
๐ฟ0
๐ฟ
the
(๐‘๐‘…,๐ฟ , ๐‘๐‘…,๐ฟ , ๐น ๐‘…,๐ฟ ; ๐ด)
We denote by
the
dierential
characteristics with
respect to a predictable locally integrable increasing process
๐ด๐‘ก := ๐‘ก +
โˆ‘
Var(๐ต ๐‘…๐ฟ,๐‘– )๐‘ก +
๐‘–
โˆ‘
๐‘–,๐‘—
โˆ™ ๐ด = ๐ถ ๐‘…,๐ฟ , and ๐น ๐‘…,๐ฟ
๐‘๐‘…,๐ฟ โˆ™ ๐ด = ๐ต ๐‘…,๐ฟ , ๐‘๐‘…,๐ฟ (
)
๐‘๐‘…
๐‘๐‘…๐ฟ
= (๐‘๐‘… , ๐‘Ž๐ฟ )โŠค and ๐‘๐‘…,๐ฟ = (๐‘๐‘…๐ฟ
, i.e., ๐‘๐‘…๐ฟ
โŠค
๐ฟ
)
๐‘
(๐‘๐‘…๐ฟ )
โˆ™
๐ด=
e.g.,
(
)
Var(๐ถ ๐‘…๐ฟ,๐‘–๐‘— )๐‘ก + โˆฃ(๐‘ฅ, ๐‘ฅโ€ฒ )โˆฃ2 โˆง 1 โˆ— ๐œˆ๐‘ก๐‘…,๐ฟ .
Then
๐‘๐‘…,๐ฟ
๐ด,
๐ด = ๐œˆ ๐‘…,๐ฟ .
is a ๐‘‘-vector
โˆ™
We write
satisfying
โŸจ๐‘…๐‘ , ๐ฟ๐‘ โŸฉ. We will often use that
โˆซ
(โˆฃ๐‘ฅโˆฃ2 + โˆฃ๐‘ฅโ€ฒ โˆฃ2 ) โˆง (1 + โˆฃ๐‘ฅโ€ฒ โˆฃ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )) < โˆž
(2.5)
โ„๐‘‘ ×โ„
because
๐ฟ
and a cut-o function
โ„Žฬ„.
๐‘Œ
drift rate
of
๐‘Œ
be any scalar
relative to
๐ด
We call
๐‘Œ
๐‘Ž := ๐‘ +
the
๐‘Œ
(๐‘๐‘Œ , ๐‘๐‘Œ , ๐น ๐‘Œ )
is a special semimartingale (cf. [34, II.2.29]). Let
semimartingale with dierential characteristics
โˆซ
(
)
๐‘ฅ โˆ’ โ„Žฬ„(๐‘ฅ) ๐น ๐‘Œ (๐‘‘๐‘ฅ)
whenever the integral is well dened with values in
[โˆ’โˆž, โˆž], even if it is not nite. Note that ๐‘Ž๐‘Œ does not depend on the
choice of โ„Žฬ„. If ๐‘Œ is special, the drift rate is nite and even ๐ด-integrable (and
๐ฟ
๐ฟ
๐ฟ
vice versa). As an example, ๐‘Ž is the drift rate of ๐ฟ and ๐‘Ž โˆ™ ๐ด = ๐ด yields
the drift.
III.2 Preliminaries
33
Remark 2.3.
Then its
drift rate
fact that
Assume ๐‘Œ is a nonpositive scalar semimartingale.
๐‘Ž๐‘Œ is well dened with values in [โˆ’โˆž, โˆž). Indeed, the
๐‘Œ = ๐‘Œโˆ’ + ฮ”๐‘Œ โ‰ค 0 implies that ๐‘ฅ โ‰ค โˆ’๐‘Œโˆ’ ๐น ๐‘Œ (๐‘‘๐‘ฅ)-a.e.
If
๐‘Œ
is a scalar semimartingale with drift rate
semimartingale with nonpositive drift rate
๐‘Ž๐‘Œ โˆˆ [โˆ’โˆž, 0],
we call
๐‘Œ
a
๐‘Œ
. Here ๐‘Ž need not be nite, as in
the case of a compound Poisson process with negative, non-integrable jumps.
We refer to Kallsen [39] for the concept of
๐œŽ -localization.
Recalling that
โ„ฑ0
is trivial, we conclude the following, e.g., from [41, Appendix 3].
Lemma 2.4. Let ๐‘Œ be a semimartingale with nonpositive drift rate.
(i) ๐‘Œ is a ๐œŽ-supermartingale โ‡” ๐‘Ž๐‘Œ is nite โ‡” ๐‘Œ is ๐œŽ-locally of class (D).
(ii) ๐‘Œ is a local supermartingale โ‡” ๐‘Ž๐‘Œ โˆˆ ๐ฟ(๐ด) โ‡” ๐‘Œ is locally of class (D).
(iii) If ๐‘Œ is uniformly bounded from below, it is a supermartingale.
III.2.4 Constraints and Degeneracies
We introduce some set-valued processes that will be used in the sequel, that
is, for each
(๐œ”, ๐‘ก)
they describe a subset of
โ„๐‘‘ .
We refer to Rockafellar [64]
and Aliprantis and Border [1, ยŸ18] for background.
We start by expressing the budget constraint in this fashion. The process
{
}
{
}
C๐‘ก0 (๐œ”) := ๐‘ฆ โˆˆ โ„๐‘‘ : ๐น๐‘ก๐‘… (๐œ”) ๐‘ฅ โˆˆ โ„๐‘‘ : ๐‘ฆ โŠค ๐‘ฅ < โˆ’1 = 0
was called the
natural constraints
in [41].
Clearly
C0
is closed, convex,
and contains the origin. Moreover, one can check (see [41, ยŸ3.3]) that it is
predictable
(C 0 )โˆ’1 (๐บ)
in the sense that for each closed
= {(๐œ”, ๐‘ก) : C๐‘ก (๐œ”) โˆฉ ๐บ โˆ•= โˆ…}
๐บ โŠ† โ„๐‘‘ ,
the lower inverse image
is predictable. (Here one can replace
closed by compact or by open; see [64, 1A].) A statement such as C
0
closed means that C๐‘ก (๐œ”) is closed for all
omit the arguments
(๐œ”, ๐‘ก).
(๐œ”, ๐‘ก);
0 is
moreover, we will often
We also consider the slightly smaller set-valued
process
{
}
{
}
C 0,โˆ— := ๐‘ฆ โˆˆ โ„๐‘‘ : ๐น ๐‘… ๐‘ฅ โˆˆ โ„๐‘‘ : ๐‘ฆ โŠค ๐‘ฅ โ‰ค โˆ’1 = 0 .
These processes relate to the budget constraint as follows.
Lemma 2.5. A process ๐œ‹ โˆˆ ๐ฟ(๐‘…) satises โ„ฐ(๐œ‹ โˆ™ ๐‘…) โ‰ฅ 0 (> 0) up to evanescence if and only if ๐œ‹ โˆˆ C 0 (C 0,โˆ— ) ๐‘ƒ โŠ— ๐ด-a.e.
Proof.
โ„ฐ(๐œ‹ โˆ™ ๐‘…) > 0 if and only if 1 + ๐œ‹ โŠค ฮ”๐‘… > 0 ([34, II.8a]).
Writing ๐‘‰ (๐‘ฅ) = 1{๐‘ฅ: 1+๐œ‹ โŠค ๐‘ฅโ‰ค0} (๐‘ฅ), we have that (๐‘ƒ โŠ— ๐ด){๐œ‹ โˆˆ
/ C 0,โˆ— } =
[
]
โˆ‘
๐ธ[๐‘‰ (๐‘ฅ) โˆ— ๐œˆ๐‘‡๐‘… ] = ๐ธ[๐‘‰ (๐‘ฅ) โˆ— ๐œ‡๐‘…
๐‘ โ‰ค๐‘‡ 1{๐‘ฅ: 1+๐œ‹๐‘ โŠค ฮ”๐‘…๐‘  โ‰ค0} . For the equiva๐‘‡] = ๐ธ
0
lence with C , interchange strict and non-strict inequality signs.
Recall that
III Bellman Equation
34
The process
C 0,โˆ—
{(1
C 0,โˆ—
is not closed in general (nor relatively open). Clearly
โŠ† C 0 , and in fact
+ ๐‘›โˆ’1 )๐‘ฆ}๐‘›โ‰ฅ1 is in
is predictable; cf. [1,
C 0 is the closure of C 0,โˆ— : for ๐‘ฆ โˆˆ C๐‘ก0 (๐œ”), the sequence
C๐‘ก0,โˆ— (๐œ”) and converges to ๐‘ฆ . This implies that C 0,โˆ—
0,โˆ—
18.3]. We will not be able to work directly with C
because closedness is essential for the measurable selection arguments that
will be used.
We turn to the exogenous portfolio constraints, i.e., the set-valued process
C
containing the origin. We consider the following conditions:
(C1) C
(C2) C
(C3) If
is predictable.
is closed.
๐‘ โˆˆ (0, 1): There exists a (0, 1)-valued process ๐œ‚ such that
๐‘ฆ โˆˆ (C โˆฉ C 0 ) โˆ– C 0,โˆ— =โ‡’ ๐œ‚๐‘ฆ โˆˆ C for all ๐œ‚ โˆˆ (๐œ‚, 1), ๐‘ƒ โŠ— ๐ด-a.e.
Condition (C3) is clearly satised if
case of a continuous process
๐‘…,
C โˆฉ C 0 โŠ† C 0,โˆ— ,
which includes the
and it is always satised if
with respect to the origin or even convex. If
๐‘ < 0,
C
is star-shaped
(C3) should be read as
always being satised. We motivate (C3) by
Example 2.6.
We assume that there is no intermediate consumption and
๐‘ฅ0 = 1. Consider the one-period binomial model of a nancial market, i.e.,
๐‘† = โ„ฐ(๐‘…) is a scalar process which is constant up to time ๐‘‡ , where it has
a single jump, say, ๐‘ƒ [ฮ”๐‘…๐‘‡ = โˆ’1] = ๐‘0 and ๐‘ƒ [ฮ”๐‘…๐‘‡ = ๐พ] = 1 โˆ’ ๐‘0 , where
๐พ > 0 is a constant and ๐‘0 โˆˆ (0, 1). The ltration is generated by ๐‘… and
we consider C โ‰ก {0} โˆช {1}. Then ๐ธ[๐‘ˆ (๐‘‹๐‘‡ (๐œ‹))] = ๐‘ˆ (1) if ๐œ‹๐‘‡ = 0 and
๐ธ[๐‘ˆ (๐‘‹๐‘‡ (๐œ‹))] = ๐‘0 ๐‘ˆ (0) + (1 โˆ’ ๐‘0 )๐‘ˆ (1 + ๐พ) if ๐œ‹๐‘‡ = 1. If ๐‘ˆ (0) > โˆ’โˆž, and if
๐พ is large enough, ๐œ‹๐‘‡ = 1 performs better and its terminal wealth vanishes
with probability ๐‘0 > 0. Of course, this cannot happen if ๐‘ˆ (0) = โˆ’โˆž,
i.e., ๐‘ < 0. The constants can also be chosen such that both strategies are
optimal, so there is
no uniqueness.
We have included only positive wealth processes in our denition of
๐’œ;
only these match our multiplicative setting. Under (C3), the Inada condition
๐‘ˆ โ€ฒ (0) = โˆž
ensures that vanishing wealth is not optimal.
The nal set-valued process is related to linear dependencies of the assets.
As in [41], the predictable process of
null-investments
is
{
}
N := ๐‘ฆ โˆˆ โ„๐‘‘ : ๐‘ฆ โŠค ๐‘๐‘… = 0, ๐‘ฆ โŠค ๐‘๐‘… = 0, ๐น ๐‘… {๐‘ฅ : ๐‘ฆ โŠค ๐‘ฅ โˆ•= 0} = 0 .
โ„๐‘‘ , hence closed, and provide the pointwise
description of the null-space of ๐ป 7โ†’ ๐ป โˆ™ ๐‘…. That is, ๐ป โˆˆ ๐ฟ(๐‘…) satises
๐ป โˆ™ ๐‘… โ‰ก 0 if and only if ๐ป โˆˆ N ๐‘ƒ โŠ— ๐ด-a.e. An investment with values in
N has no eect on the wealth process.
Its values are linear subspaces of
III.3 The Bellman Equation
35
III.3 The Bellman Equation
We have now introduced the necessary notation to formulate our rst main
result. Two special cases of our Bellman equation can be found in the pioneering work of Mania and Tevzadze [54] and Hu
et al. [33].
These articles
consider models with continuous asset prices and we shall indicate the connections as we specialize to that case in Section III.3.3. A related equation
also arises in the study of mean-variance hedging by ยƒerný and Kallsen [11]
in the context of locally square-integrable semimartingales, although they do
not use dynamic programming explicitly. Due to the quadratic setting, that
equation is more explicit than ours and the mathematical treatment is quite
dierent. Czichowsky and Schweizer [13] study a cone-constrained version of
the related Markowitz problem and there the equation is no longer explicit.
The Bellman equation highlights the local structure of our utility maximization problem. In addition, it has two main benets. First, it can be
used as an abstract tool to derive properties of the optimal strategies and
the opportunity process. Second, one can try to solve the equation directly
in a given model and to deduce the optimal strategies. This is the point of
view taken in Section III.5 and obviously requires the precise form of the
equation.
The following assumptions are in force for the entire Section III.3.
Assumptions 3.1.
The utility maximization problem is nite, there exists
an optimal strategy
(ห†
๐œ‹ , ๐‘ห†) โˆˆ ๐’œ,
and
C
satises (C1)-(C3).
III.3.1 Bellman Equation in Joint Characteristics
Our rst main result is the Bellman equation stated as a description of
the drift rate of the opportunity process. We recall the conjugate function
๐‘ˆ๐‘กโˆ— (๐‘ฆ) = โˆ’ 1๐‘ž ๐‘ฆ ๐‘ž ๐ท๐‘ก๐›ฝ .
Theorem 3.2. The drift rate ๐‘Ž๐ฟ of the opportunity process satises
๐‘‘๐œ‡
โˆ’๐‘โˆ’1 ๐‘Ž๐ฟ = ๐‘ˆ โˆ— (๐ฟโˆ’ ) ๐‘‘๐ด
+ max ๐‘”(๐‘ฆ),
๐‘ฆโˆˆC โˆฉC 0
(3.1)
where ๐‘” is the predictable random function
๐‘”(๐‘ฆ) := ๐ฟโˆ’ ๐‘ฆ
โˆซ
+
โŠค
(
๐‘…
๐‘ +
๐‘๐‘…๐ฟ
๐ฟโˆ’
+
(๐‘โˆ’1) ๐‘…
2 ๐‘ ๐‘ฆ
)
โˆซ
๐‘ฅโ€ฒ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
+
โ„๐‘‘ ×โ„
{
}
(๐ฟโˆ’ + ๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )).
โ„๐‘‘ ×โ„
(3.2)
The unique (๐‘ƒ โŠ— ๐œ‡โˆ˜ -a.e.) optimal propensity to consume is
๐œ…
ห†=
(๐ท)
๐ฟ
1
1โˆ’๐‘
.
(3.3)
III Bellman Equation
36
Any optimal trading strategy ๐œ‹โˆ— satises
๐œ‹ โˆ— โˆˆ arg max ๐‘”
(3.4)
C โˆฉC 0
and the corresponding optimal wealth process and consumption are given by
(
)
๐‘‹ โˆ— = ๐‘ฅ0 โ„ฐ ๐œ‹ โˆ— โˆ™ ๐‘… โˆ’ ๐œ…
ห†โˆ™ ๐œ‡ ;
๐‘โˆ— = ๐‘‹ โˆ— ๐œ…
ห†.
Note that the maximization in (3.1) can be understood as a local version
of the optimization problem.
Indeed, recalling (2.1), the right hand side
of (3.1) is the maximum of a single function over certain points
โ„+ ×โ„๐‘‘ that correspond to the admissible controls
(๐œ…, ๐œ‹).
(๐‘˜, ๐‘ฆ) โˆˆ
Moreover, optimal
controls are related to maximizers of this function, a characteristic feature
of any dynamic programming equation. The maximum of
due to the jumps of
๐‘…;
๐‘”
is not explicit
this simplies in the continuous case considered in
Section III.3.3 below. Some mathematical comments are also in order.
Remark 3.3.
(i) The random function
๐‘”
is well dened on
C0
in the
extended sense (see Lemma 6.2) and it does not depend on the choice
of the cut-o function
(ii) For
๐‘<0
โ„Ž
by [34, II.2.25].
we have a more precise statement: Given
โˆ— ห† ) is optimal
as in (3.3), (๐œ‹ , ๐œ…
๐‘”.
(๐ฟ, ๐œ‹ โˆ— , ๐œ…
ห† ).
and maximizes
triplet
(iii) For
if and only if
๐œ‹ โˆ— โˆˆ ๐ฟ(๐‘…) and ๐œ…
ห†
C โˆฉ C0
๐œ‹ โˆ— takes values in
This will follow from Corollary 5.4 applied to the
๐‘ โˆˆ (0, 1), partial results in this direction follow from Section III.5.
C by the next item.
The question is trivial for convex
(iv) If
C
is convex,
arg maxC โˆฉC 0 ๐‘”
N
of any two elements lies in
is unique in the sense that the dierence
(see Lemma 6.3).
We split the proof of Theorem 3.2 into several steps; the plan is as follows.
Let
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ
that
๐‘‹ = ๐‘‹(๐œ‹, ๐œ…). We recall from
โˆซ
๐‘(๐œ‹, ๐œ…) := ๐ฟ ๐‘1 ๐‘‹ ๐‘ + ๐‘ˆ๐‘  (๐œ…๐‘  ๐‘‹๐‘  ) ๐œ‡(๐‘‘๐‘ )
and denote
Proposition 2.2
(๐œ‹, ๐œ…) is optimal. Hence
(๐œ‹, ๐œ…); the maximum
is a supermartingale, and a martingale if and only if
we shall calculate its drift rate and then maximize over
will be attained at any optimal strategy. This is fairly straightforward and
essentially the content of Lemma 3.7 below.
๐‘‘
maximize over a subset of โ„ for each
(๐œ”, ๐‘ก)
In the Bellman equation, we
and not over a set of strategies.
This nal step is a measurable selection problem and its solution will be the
second part of the proof.
Lemma 3.4. Let (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ . The drift rate of ๐‘(๐œ‹, ๐œ…) is
(
)
๐‘‘๐œ‡
๐‘Ž๐‘(๐œ‹,๐œ…) = ๐‘‹(๐œ‹, ๐œ…)๐‘โˆ’ ๐‘โˆ’1 ๐‘Ž๐ฟ + ๐‘“ (๐œ…) ๐‘‘๐ด
+ ๐‘”(๐œ‹) โˆˆ [โˆ’โˆž, โˆž),
where ๐‘“๐‘ก (๐‘˜) := ๐‘ˆ๐‘ก (๐‘˜) โˆ’ ๐ฟ๐‘กโˆ’ ๐‘˜ and ๐‘” is given by
and ๐‘Ž๐‘(๐œ‹,๐œ…) โˆˆ (โˆ’โˆž, 0] for (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ .
. Moreover, ๐‘Ž๐‘(ห†๐œ‹,ห†๐œ…) = 0,
(3.2)
III.3 The Bellman Equation
Proof.
We can assume that the initial capital is
then in particular
formula, we have
๐‘ := ๐‘(๐œ‹, ๐œ…) is nite. We also
๐‘‹ ๐‘ = โ„ฐ(๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡)๐‘ = โ„ฐ(๐‘Œ )
๐‘Œ = ๐‘(๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡) +
๐‘(๐‘โˆ’1) โŠค ๐‘…
๐œ‹ ๐‘ ๐œ‹
2
โˆ™
37
๐‘ฅ0 = 1. Let (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ,
set ๐‘‹ := ๐‘‹(๐œ‹, ๐œ…). By Itô's
with
{
}
๐ด + (1 + ๐œ‹ โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ’ ๐‘๐œ‹ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘… .
๐‘ and using ๐‘‹๐‘  = ๐‘‹๐‘ โˆ’ ๐œ‡(๐‘‘๐‘ )-a.e.
โˆ’๐‘ โˆ™
๐‘‹โˆ’
๐‘ = ๐‘โˆ’1 (๐ฟโˆ’๐ฟ0 +๐ฟโˆ’ โˆ™ ๐‘Œ +[๐ฟ, ๐‘Œ ])+๐‘ˆ (๐œ…) โˆ™ ๐œ‡.
Integrating by parts in the denition of
(path-by-path), we have
Here
[๐ฟ, ๐‘Œ ] = [๐ฟ๐‘ , ๐‘Œ ๐‘ ] +
โˆ‘
ฮ”๐ฟฮ”๐‘Œ
{
}
= ๐‘๐œ‹ โŠค ๐‘๐‘…๐ฟ โˆ™ ๐ด + ๐‘๐‘ฅโ€ฒ ๐œ‹ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘…,๐ฟ + ๐‘ฅโ€ฒ (1 + ๐œ‹ โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ’ ๐‘๐œ‹ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘…,๐ฟ .
โˆ’๐‘ โˆ™
๐‘‹โˆ’
๐‘
Thus
equals
โŠค ๐‘… โˆ™
โŠค ๐‘…๐ฟ โˆ™
๐ด
๐‘โˆ’1 (๐ฟ โˆ’ ๐ฟ0 ) + ๐ฟโˆ’ ๐œ‹ โˆ™ ๐‘… + ๐‘“ (๐œ…) โˆ™ ๐œ‡ + ๐ฟโˆ’ (๐‘โˆ’1)
2 ๐œ‹ ๐‘ ๐œ‹ ๐ด+๐œ‹ ๐‘
{
}
โ€ฒ โŠค
๐‘…,๐ฟ
โ€ฒ
โˆ’1
โŠค ๐‘
โˆ’1
โŠค
๐‘…,๐ฟ
+๐‘ฅ๐œ‹ ๐‘ฅโˆ—๐œ‡
+ (๐ฟโˆ’ + ๐‘ฅ ) ๐‘ (1 + ๐œ‹ ๐‘ฅ) โˆ’ ๐‘ โˆ’ ๐œ‹ ๐‘ฅ โˆ— ๐œ‡ .
Writing
๐‘ฅ = โ„Ž(๐‘ฅ) + ๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)
and
¯ = ๐‘… โˆ’ (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘…
๐‘…
as in (2.4),
โˆ’๐‘ โˆ™
๐‘‹โˆ’
๐‘=
โˆ’1
๐‘
(3.5)
¯+
(๐ฟ โˆ’ ๐ฟ0 ) + ๐ฟโˆ’ ๐œ‹ โˆ™ ๐‘…
โ€ฒ โŠค
+ ๐‘ฅ ๐œ‹ โ„Ž(๐‘ฅ) โˆ— ๐œ‡
Since
๐œ‹
๐‘…,๐ฟ
๐‘…๐ฟ
๐‘“ (๐œ…) โˆ™ ๐œ‡ + ๐ฟโˆ’ ๐œ‹ โŠค ๐‘๐ฟโˆ’
{ โˆ’1
โ€ฒ
โŠค ๐‘
(
+ (๐ฟโˆ’ + ๐‘ฅ ) ๐‘
+
(๐‘โˆ’1) ๐‘… )
2 ๐‘ ๐œ‹
โˆ’1
โŠค
(1 + ๐œ‹ ๐‘ฅ) โˆ’ ๐‘
๐ด
}
โˆ’ ๐œ‹ โ„Ž(๐‘ฅ) โˆ— ๐œ‡๐‘…,๐ฟ .
โˆ™
need not be locally bounded, we use from now on a predictable cut-
Then the compensator
๐œ‹ โŠค โ„Ž(๐‘ฅ) is bounded, e.g., โ„Ž(๐‘ฅ) = ๐‘ฅ1{โˆฃ๐‘ฅโˆฃโ‰ค1}โˆฉ{โˆฃ๐œ‹โŠค ๐‘ฅโˆฃโ‰ค1} .
โ€ฒ โŠค
๐‘…,๐ฟ exists, since ๐ฟ is special.
of ๐‘ฅ ๐œ‹ โ„Ž(๐‘ฅ) โˆ— ๐œ‡
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ .
Then the compensator of the last integral in the
o function
Let
โ„Ž
such that
right hand side of (3.5) also exists; indeed, all other terms in that equality
are special, since
๐‘
is a supermartingale.
The drift rate can now be read
from (3.5) and (2.4), and it is nonpositive by the supermartingale property.
The drift rate vanishes for the optimal
(ห†
๐œ‹, ๐œ…
ห†)
by the martingale condition
from Proposition 2.2.
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ โˆ– ๐’œ๐‘“ ๐ธ . Note that necessarily ๐‘ < 0 (otherwise
๐‘ is well
=
Thus ๐‘ โ‰ค 0, so by Remark 2.3 the drift rate ๐‘Ž
dened with values in [โˆ’โˆž, โˆž)alternatively, this can also be read from
๐‘
the integrals in (3.5) via (2.5). Using directly the denition of ๐‘Ž , we nd
๐‘
the same formula for ๐‘Ž is as above.
Now consider
๐’œ๐‘“
๐’œ๐‘“ ๐ธ ).
We do not have the supermartingale property for
๐‘(๐œ‹,๐œ…)
is not evident that ๐‘Ž
โ‰ค0
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ โˆ– ๐’œ๐‘“ ๐ธ , so it
in that case. However, we have the following
Lemma 3.5. Let (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ . Then ๐‘Ž๐‘ (๐œ‹, ๐œ…) โˆˆ [0, โˆž] implies ๐‘Ž๐‘ (๐œ‹, ๐œ…) = 0.
38
III Bellman Equation
Proof.
Denote
๐‘ = ๐‘(๐œ‹, ๐œ…).
For
๐‘ > 0
claim is immediate from Lemma 3.4.
Let
we have
๐‘ < 0.
๐’œ๐‘“ = ๐’œ๐‘“ ๐ธ and the
Then ๐‘ โ‰ค 0 and by
๐‘ โˆˆ [0, โˆž] implies that ๐‘ is a submartingale . Hence
Lemma 2.4(iii), ๐‘Ž
[โˆซ๐‘‡
]
๐ธ[๐‘๐‘‡ ] = ๐ธ 0 ๐‘ˆ๐‘ก (๐œ…๐‘ก ๐‘‹๐‘ก (๐œ‹, ๐œ…)) ๐œ‡โˆ˜ (๐‘‘๐‘ก) > โˆ’โˆž, that is, (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ . Now
๐‘
Lemma 3.4 yields ๐‘Ž (๐œ‹, ๐œ…) โ‰ค 0.
We observe in Lemma 3.4 that the drift rate splits into separate functions
involving
๐œ…
and
๐œ‹,
respectively. For this reason, we can single out the
Proof of the consumption formula
.
(๐œ‹, ๐œ…) โˆˆ ๐’œ. Note the follow(๐œ‹, ๐œ…โˆ— ) โˆˆ ๐’œ for any nonnegaโˆซ
๐‘‡ โˆ—
โˆ—
โˆ—
tive optional process ๐œ… such that
0 ๐œ…๐‘  ๐œ‡(๐‘‘๐‘ ) < โˆž and ๐œ…๐‘‡ = 1. Indeed,
๐‘‹(๐œ‹, ๐œ…) = ๐‘ฅ0 โ„ฐ(๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡) is positive by assumption. As ๐œ‡ is continuous,
๐‘‹(๐œ‹, ๐œ…โˆ— ) = ๐‘ฅ0 โ„ฐ(๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ…โˆ— โˆ™ ๐œ‡) is also positive.
In particular, let (ห†
๐œ‹, ๐œ…
ห† ) be optimal, ๐›ฝ = (1 โˆ’ ๐‘)โˆ’1 and ๐œ…โˆ— = (๐ท/๐ฟ)๐›ฝ ; then
(ห†
๐œ‹ , ๐œ…โˆ— ) โˆˆ ๐’œ. In fact the paths of ๐‘ˆ (๐œ…โˆ— ๐‘‹(ห†
๐œ‹ , ๐œ…โˆ— )) = ๐‘โˆ’1 ๐ท๐›ฝ๐‘+1 ๐‘‹(ห†
๐œ‹ , ๐œ…โˆ— )๐‘ ๐ฟโˆ’๐›ฝ๐‘
are bounded ๐‘ƒ -a.s. (because the processes are càdlàg; ๐ฟ, ๐ฟโˆ’ > 0 and ๐›ฝ๐‘+1 =
๐›ฝ > 0) so that (ห†
๐œ‹ , ๐œ…โˆ— ) โˆˆ ๐’œ๐‘“ .
โˆ—
๐›ฝ
Note that ๐‘ƒ โŠ— ๐œ‡-a.e., we have ๐œ… = (๐ท/๐ฟโˆ’ ) = arg max๐‘˜โ‰ฅ0 ๐‘“ (๐‘˜), hence
๐‘“ (๐œ…โˆ— ) โ‰ฅ ๐‘“ (ห†
๐œ…). Suppose (๐‘ƒ โŠ— ๐œ‡){๐‘“ (๐œ…โˆ— ) > ๐‘“ (ห†
๐œ…)} > 0, then the formula from
๐‘(ห†
๐œ‹ ,ห†
๐œ…) = 0 imply ๐‘Ž๐‘(ห†
๐œ‹ ,๐œ…โˆ— ) โ‰ฅ 0 and (๐‘ƒ โŠ—๐ด){๐‘Ž๐‘(ห†
๐œ‹ ,๐œ…โˆ— ) > 0} > 0,
Lemma 3.4 and ๐‘Ž
a contradiction to Lemma 3.5. It follows that ๐œ…
ห† = ๐œ…โˆ— ๐‘ƒ โŠ— ๐œ‡-a.e. since ๐‘“ has
(3.3)
Let
ing feature of our parametrization: we have
a unique maximum.
Remark 3.6.
The previous proof does not use the assumptions (C1)-(C3).
Lemma 3.7. Let ๐œ‹ be a predictable process with values in C
โˆฉ C 0,โˆ— .
Then
{
}
(๐‘ƒ โŠ— ๐ด) ๐‘”(ห†
๐œ‹ ) < ๐‘”(๐œ‹) = 0.
Proof.
We argue by contradiction and assume
By redening
๐œ‹,
we may assume that
๐œ‹ = ๐œ‹
ห†
(๐‘ƒ โŠ— ๐ด){๐‘”(ห†
๐œ‹ ) < ๐‘”(๐œ‹)} > 0.
on the complement of this
predictable set. Then
๐‘”(ห†
๐œ‹ ) โ‰ค ๐‘”(๐œ‹)
and
(๐‘ƒ โŠ— ๐ด){๐‘”(ห†
๐œ‹ ) < ๐‘”(๐œ‹)} > 0.
(3.6)
๐œ‹ is ๐œŽ -bounded, we can nd a constant ๐ถ > 0 such that the process
๐œ‹
หœ := ๐œ‹1โˆฃ๐œ‹โˆฃโ‰ค๐ถ + ๐œ‹
ห† 1โˆฃ๐œ‹โˆฃ>๐ถ again satises (3.6); that is, we may assume that ๐œ‹
0,โˆ— , this implies (๐œ‹, ๐œ…
is ๐‘…-integrable. Since ๐œ‹ โˆˆ C โˆฉ C
ห† ) โˆˆ ๐’œ (as observed
above, the consumption ๐œ…
ห† plays no role here). The contradiction follows as
As
in the previous proof.
In view of Lemma 3.7, the main task will be to construct a
maximizing sequence for
๐‘”.
measurable
III.3 The Bellman Equation
Lemma 3.8. Under Assumptions 3.1, there exists a sequence
dictable C โˆฉ C 0,โˆ— -valued processes such that
lim sup ๐‘”(๐œ‹ ๐‘› ) = sup ๐‘” ๐‘ƒ โŠ— ๐ด-a.e.
(๐œ‹ ๐‘› )
39
of pre-
C โˆฉC 0
๐‘›
We defer the proof of this lemma to Appendix III.6, together with the
study of the properties of
Proof of Theorem 3.2.
๐œ‹ ๐‘› yields
๐‘Ž๐‘(ห†๐œ‹,ห†๐œ…) =
๐œ‹ =
0 =
๐‘ƒ โŠ— ๐œ‡-a.e.
๐‘”.
Let
The theorem can then be proved as follows.
๐œ‹๐‘›
be as in Lemma 3.8. Then Lemma 3.7 with
๐‘”(ห†
๐œ‹ ) = supC โˆฉC 0 ๐‘” , which
๐‘‘๐œ‡
๐‘โˆ’1 ๐‘Ž๐ฟ + ๐‘“ (ห†
๐œ…) ๐‘‘๐ด
+ ๐‘”(ห†
๐œ‹ ).
is (3.4).
By Lemma 3.4 we have
This is (3.1) as
๐‘“ (ห†
๐œ…) = ๐‘ˆ โˆ— (๐ฟโˆ’ )
due to (3.3).
III.3.2 Bellman Equation as BSDE
In this section we express the Bellman equation as a BSDE. The unique
orthogonal decomposition of the local martingale
๐‘€๐ฟ
with respect to
๐‘…
(cf. [34, III.4.24]) leads to the representation
๐ฟ = ๐ฟ0 + ๐ด๐ฟ + ๐œ‘๐ฟ โˆ™ ๐‘…๐‘ + ๐‘Š ๐ฟ โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) + ๐‘ ๐ฟ ,
(3.7)
๐œ‘๐ฟ โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘…๐‘ ), ๐‘Š ๐ฟ โˆˆ ๐บ๐‘™๐‘œ๐‘ (๐œ‡๐‘… ), and ๐‘ ๐ฟ is
๐ฟ
๐‘
๐‘
๐‘ƒ
๐ฟ หœ = 0. The
a local martingale such that โŸจ(๐‘ ) , ๐‘… โŸฉ = 0 and ๐‘€ ๐‘… (ฮ”๐‘ โˆฃ๐’ซ)
๐œ‡
๐ฟ
๐‘…
last statement means that ๐ธ[(๐‘‰ ฮ”๐‘ )โˆ—๐œ‡๐‘‡ ] = 0 for any suciently integrable
where, using the notation of [34],
predictable function
๐‘‰ = ๐‘‰ (๐œ”, ๐‘ก, ๐‘ฅ). We also introduce
โˆซ
๐ฟ
ห†
๐‘Š๐‘ก :=
๐‘Š ๐ฟ (๐‘ก, ๐‘ฅ) ๐œˆ ๐‘… ({๐‘ก} × ๐‘‘๐‘ฅ),
โ„๐‘‘
)
๐ฟ โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) = ๐‘Š ๐ฟ (ฮ”๐‘…)1
ห† ๐ฟ by denition of the
then ฮ” ๐‘Š
{ฮ”๐‘…โˆ•=0} โˆ’ ๐‘Š
๐ฟ
๐‘…
๐‘…
purely discontinuous local martingale ๐‘Š โˆ— (๐œ‡ โˆ’ ๐œˆ ) and we can write
(
ห† ๐ฟ + ฮ”๐‘ ๐ฟ .
ฮ”๐ฟ = ฮ”๐ด๐ฟ + ๐‘Š ๐ฟ (ฮ”๐‘…)1{ฮ”๐‘…โˆ•=0} โˆ’ ๐‘Š
We recall that Assumptions 3.1 are in force. Now (3.1) can be restated as
follows, the random function
๐‘”
being the same as before but in new notation.
Corollary 3.9. The opportunity process ๐ฟ and the processes dened by
satisfy the BSDE
(3.7)
๐ฟ = ๐ฟ0 โˆ’ ๐‘๐‘ˆ โˆ— (๐ฟโˆ’ ) โˆ™ ๐œ‡ โˆ’ ๐‘ max ๐‘”(๐‘ฆ) โˆ™ ๐ด + ๐œ‘๐ฟ โˆ™ ๐‘…๐‘ + ๐‘Š ๐ฟ โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) + ๐‘ ๐ฟ
๐‘ฆโˆˆC โˆฉC 0
with terminal condition ๐ฟ๐‘‡ = ๐ท๐‘‡ , where ๐‘” is given by
(3.8)
๐‘”(๐‘ฆ) :=
(
) โˆซ (
(
)
(๐‘โˆ’1) )
โŠค ๐‘…
๐‘… ๐œ‘๐ฟ
ห† ๐ฟ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ)
๐ฟโˆ’ ๐‘ฆ ๐‘ + ๐‘ ๐ฟโˆ’ + 2 ๐‘ฆ +
ฮ”๐ด๐ฟ + ๐‘Š ๐ฟ (๐‘ฅ) โˆ’ ๐‘Š
โ„๐‘‘
โˆซ
(
){
}
ห† ๐ฟ ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ).
+
๐ฟโˆ’ + ฮ”๐ด๐ฟ + ๐‘Š ๐ฟ (๐‘ฅ) โˆ’ ๐‘Š
โ„๐‘‘
III Bellman Equation
40
We observe that the orthogonal part
๐‘๐ฟ
plays a minor role here. In a
suitable setting, it is linked to the dual problem; see Remark 5.18.
It is possible (but notationally more cumbersome) to prove a version of
๐‘”
Lemma 3.4 using
as in Corollary 3.9 and the decomposition (3.7), thus
involving only the characteristics of
of
(๐‘…, ๐ฟ).
๐‘…
instead of the joint characteristics
Using this approach, we see that the increasing process
BSDE can be chosen based on
๐‘… and without reference to ๐ฟ.
๐ด
in the
This is desirable
if we want to consider other solutions of the equation, as in Section III.4.
๐ด can be chosen to be continuous if and only if ๐‘… is
โˆ’1 ๐ด๐ฟ = โˆ’๐‘“ (ห†
quasi left continuous (cf. [34, II.2.9]). Since ๐‘
๐œ…) โˆ™ ๐œ‡ โˆ’ ๐‘”(ห†
๐œ‹ ) โˆ™ ๐ด,
๐ฟ
Var(๐ด ) is absolutely continuous with respect to ๐ด + ๐œ‡, and we conclude:
One consequence is that
Remark 3.10.
If
๐‘…
is quasi left continuous,
๐ด๐ฟ
is continuous.
๐‘… is quasi left continuous, ๐œˆ ๐‘… ({๐‘ก} × โ„๐‘‘ ) = 0 for all ๐‘ก by [34, II.1.19],
ห† ๐ฟ = 0 and we have the simpler formula
hence ๐‘Š
(
) โˆซ
(
(๐‘โˆ’1) )
๐‘… ๐œ‘๐ฟ
โŠค ๐‘…
๐‘”(๐‘ฆ) = ๐ฟโˆ’ ๐‘ฆ ๐‘ + ๐‘ ๐ฟโˆ’ + 2 ๐‘ฆ +
๐‘Š ๐ฟ (๐‘ฅ)๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ)
๐‘‘
โ„
โˆซ
(
){
}
+
๐ฟโˆ’ + ๐‘Š ๐ฟ (๐‘ฅ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ).
If
โ„๐‘‘
III.3.3 The Case of Continuous Prices
In this section we specialize the previous results to the case where
๐‘…
is a
continuous semimartingale and mild additional conditions are satised. As
usual in this setting, the martingale part of
than
๐‘…๐‘ .
๐‘…
will be denoted by
๐‘€
rather
In addition to Assumptions 3.1, the following conditions are in
force for the present Section III.3.3.
Assumptions 3.11.
(i)
(ii)
๐‘…
is continuous,
๐‘…=๐‘€+
โˆซ
๐‘‘โŸจ๐‘€ โŸฉ๐œ†
for some
(iii) the orthogonal projection of
Note that
C 0,โˆ— = โ„๐‘‘
๐œ† โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ )
C
onto
N
structure condition ),
(
โŠฅ is closed.
due to (i), in particular (C3) is void. When
๐‘…
is
continuous, it necessarily satises (ii) when a no-arbitrage property holds;
see Schweizer [71]. By (i) and (ii) we can write the dierential characteristics
โˆ‘
๐ด๐‘ก := ๐‘ก + ๐‘‘๐‘–=1 โŸจ๐‘€ ๐‘– โŸฉ๐‘ก . It will be convenient to
๐‘… = ๐œŽ๐œŽ โŠค , where ๐œŽ is a predictable matrix-valued process; hence
factorize ๐‘
๐œŽ๐œŽ โŠค ๐‘‘๐ด = ๐‘‘โŸจ๐‘€ โŸฉ. Then (ii) implies N = ker ๐œŽ โŠค because ๐œŽ๐œŽ โŠค ๐‘ฆ = 0 implies
(๐œŽ โŠค ๐‘ฆ)โŠค (๐œŽ โŠค ๐‘ฆ) = 0. Since ๐œŽ โŠค : ker(๐œŽ โŠค )โŠฅ โ†’ ๐œŽ โŠค โ„๐‘‘ is a homeomorphism, we
of
๐‘…
with respect to, e.g.,
see that (iii) is equivalent to
๐œŽโŠคC
is closed.
III.3 The Bellman Equation
This condition depends on the semimartingale
closedness of
C
itself if
๐œŽ
has full rank.
๐‘….
41
It is equivalent to the
For certain constraint sets (e.g.,
closed polyhedral or compact) the condition is satised for
all
matrices
๐œŽ,
but not so, e.g., for non-polyhedral cone constraints. We mention that violation of (iii) leads to nonexistence of optimal strategies in simple examples
(cf. Example IV.3.5) and we refer to Czichowsky and Schweizer [14] for background.
Under (i), (3.7) is the more usual Kunita-Watanabe decomposition
๐ฟ = ๐ฟ0 + ๐ด๐ฟ + ๐œ‘๐ฟ โˆ™ ๐‘€ + ๐‘ ๐ฟ ,
where
๐œ‘๐ฟ โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ )
and
see Ansel and Stricker [2,
๐‘๐ฟ
cas
is a local martingale such that
[๐‘€, ๐‘ ๐ฟ ] = 0;
โˆ… โˆ•= ๐พ โŠ† โ„๐‘‘ is a closed set,
๐‘‘๐พ (๐‘ฅ) = min{โˆฃ๐‘ฅ โˆ’ ๐‘ฆโˆฃ : ๐‘ฆ โˆˆ ๐พ},
3]. If
we denote
๐‘‘2๐พ is
๐พ
the squared distance. We also dene the (set-valued) projection ฮ 
which
๐‘‘
maps ๐‘ฅ โˆˆ โ„ to the points in ๐พ with minimal distance to ๐‘ฅ,
the Euclidean distance to
๐พ
by
and
{
}
ฮ ๐พ (๐‘ฅ) = ๐‘ฆ โˆˆ ๐พ : โˆฃ๐‘ฅ โˆ’ ๐‘ฆโˆฃ = ๐‘‘๐พ (๐‘ฅ) โˆ•= โˆ….
If
๐พ
is convex,
ฮ ๐พ
is the usual (single-valued) Euclidean projection. In the
present continuous setting, the random function
๐‘”
(
๐œ‘๐ฟ
๐‘”(๐‘ฆ) = ๐ฟโˆ’ ๐‘ฆ โŠค ๐œŽ๐œŽ โŠค ๐œ† +
+
๐ฟโˆ’
simplies considerably:
๐‘โˆ’1
2 ๐‘ฆ
)
(3.9)
and so the Bellman BSDE becomes more explicit.
Corollary 3.12. Any optimal trading strategy ๐œ‹โˆ— satises
๐œŽ โŠค ๐œ‹ โˆ— โˆˆ ฮ ๐œŽ
โŠคC
(
{
๐œ‘๐ฟ )}
๐œŽ โŠค (1 โˆ’ ๐‘)โˆ’1 ๐œ† +
.
๐ฟโˆ’
The opportunity process satises the BSDE
๐ฟ = ๐ฟ0 โˆ’ ๐‘๐‘ˆ โˆ— (๐ฟโˆ’ ) โˆ™ ๐œ‡ + ๐น (๐ฟโˆ’ , ๐œ‘๐ฟ ) โˆ™ ๐ด + ๐œ‘๐ฟ โˆ™ ๐‘€ + ๐‘ ๐ฟ ;
๐ฟ๐‘‡ = ๐ท๐‘‡ ,
where
๐น (๐ฟโˆ’ , ๐œ‘๐ฟ ) =
{
(
)
(
1
๐œ‘๐ฟ )
2
โŠค
โˆ’1
๐ฟโˆ’ ๐‘(1 โˆ’ ๐‘)๐‘‘๐œŽโŠค C ๐œŽ (1 โˆ’ ๐‘)
๐œ†+
+
2
๐ฟโˆ’
๐‘ โŠค
๐‘โˆ’1 ๐œŽ
(
}
๐œ‘๐ฟ )2
๐œ†+
.
๐ฟโˆ’
2
๐‘
If C is a convex cone, ๐น (๐ฟโˆ’ , ๐œ‘๐ฟ ) = 2(๐‘โˆ’1)
๐ฟโˆ’ ฮ ๐œŽ C ๐œŽ โŠค ๐œ† + ๐ฟ๐œ‘โˆ’ . If
(
(
โˆซ
๐ฟ )โŠค
๐ฟ)
๐‘
C = โ„๐‘‘ , then ๐น (๐ฟโˆ’ , ๐œ‘๐ฟ ) โˆ™ ๐ด = 2(๐‘โˆ’1)
๐ฟโˆ’ ๐œ† + ๐ฟ๐œ‘โˆ’ ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + ๐ฟ๐œ‘โˆ’ and
(
๐ฟ)
the unique (mod. N ) optimal trading strategy is ๐œ‹โˆ— = (1 โˆ’ ๐‘)โˆ’1 ๐œ† + ๐ฟ๐œ‘โˆ’ .
โŠค
{
(
๐ฟ
)}
III Bellman Equation
42
Proof.
๐›ฝ = (1โˆ’๐‘)โˆ’1 .
Let
by completing the square in (3.9),
โˆ—
๐‘”(๐œ‹ ) =
1
2 ๐ฟโˆ’
(
๐ฟ )}
โŠค {
๐œŽ โŠค (arg maxC ๐‘”) = ฮ ๐œŽ C ๐œŽ โŠค ๐›ฝ ๐œ†+ ๐ฟ๐œ‘โˆ’
โˆ—
moreover, for any ๐œ‹ โˆˆ arg maxC ๐‘” ,
We obtain
{ (
)}
(
(
๐œ‘ ๐ฟ )โŠค โŠค (
๐œ‘๐ฟ )
๐œ‘๐ฟ )
โˆ’1 2
โŠค
๐›ฝ ๐œ†+
๐œŽ๐œŽ ๐œ† +
.
โˆ’ ๐›ฝ ๐‘‘๐œŽโŠค C ๐œŽ ๐›ฝ ๐œ† +
๐ฟโˆ’
๐ฟโˆ’
๐ฟโˆ’
ฮ  := ฮ ๐œŽ C is singlevalued, positively homogeneous, and ฮ ๐‘ฅ is orthogonal to ๐‘ฅ โˆ’ ฮ ๐‘ฅ for any
(
๐ฟ)
๐‘ฅ โˆˆ โ„๐‘‘ . Writing ฮจ := ๐œŽ โŠค ๐œ† + ๐ฟ๐œ‘โˆ’ we get ๐‘”(๐œ‹ โˆ— ) = ๐ฟโˆ’ ๐›ฝ(ฮ ฮจ)โŠค (ฮจ โˆ’ 12 ฮ ฮจ) =
(
(
๐ฟโˆ’ 21 ๐›ฝ ฮ ฮจ)โŠค ฮ ฮจ). Finally, ฮ ฮจ = ฮจ if C = โ„๐‘‘ . The result follows from
In the case where
C,
and hence
๐œŽโŠคC ,
โŠค
is a convex cone,
Corollary 3.9.
Of course the consumption formula (3.3) and Remark 3.3 still apply.
We remark that the BSDE for the unconstrained case
๐œ‡ = 0, ๐ท = 1)
C = โ„๐‘‘
(and with
was previously obtained in [54] in a similar spirit. A variant
of the constrained BSDE for an Itô process model (and
๐œ‡ = 0, ๐ท = 1)
appears in [33], where a converse approach is taken: the equation is derived
only formally and then existence results for BSDEs are employed together
with a verication argument.
We shall extend that result in Section III.5
(Example 5.8) when we study verication.
๐ฟ is
log(๐ฟ)
If
for
continuous, the BSDE of Corollary 3.12 simplies if it is stated
rather than
๐ฟ,
but in general the given form is more convenient
as the jumps are hidden in
Remark 3.13.
๐‘ ๐ฟ.
(i) Continuity of
๐‘… does not imply that ๐ฟ is continuous.
For
instance, in the Itô process model of Barndor-Nielsen and Shephard [3] with
Lévy driven coecients, the opportunity process is not continuous. See, e.g.,
Theorem 3.3 and the subsequent remark in Kallsen and Muhle-Karbe [40]. If
๐‘… satises the structure condition and the ltration ๐”ฝ is continuous, it clearly
follows that ๐ฟ is continuous. Here ๐”ฝ is called continuous if all ๐”ฝ-martingales
are continuous, as, e.g., for the Brownian ltration. In general, ๐ฟ is related
to the predictable characteristics of the asset returns rather than their levels.
As an example, Lévy models have jumps but constant characteristics; here
๐ฟ
turns out to be a smooth function (see Chapter IV).
(ii)
In the present setting we see that
๐น
has quadratic growth in
๐œ‘๐ฟ ,
so that the Bellman equation is a quadratic BSDE (see also Example 5.8).
In general,
๐น
does not satisfy the bounds which are usually assumed in
the theory of such BSDEs.
Together with existence results for the utility
maximization problem (see the citations from the introduction), the Bellman
equation yields various examples of BSDEs with the opportunity process as
a solution. This includes terminal conditions
unbounded (see also Remark II.2.4).
๐ท๐‘‡
which are integrable and
III.4 Minimality of the Opportunity Process
43
III.4 Minimality of the Opportunity Process
This section considers the Bellman equation as such, having possibly many
solutions, and we characterize the opportunity process as the minimal solution. As mentioned above, it seems more natural to use the BSDE formulation for this purpose (but see Remark 4.4). We rst have to clarify what we
mean by a solution of the BSDE. We consider
๐‘…
and
๐ด
as given. Since the
nite variation part in the BSDE is predictable, a solution will certainly be
a special semimartingale. If
โ„“
is any special semimartingale, there exists a
unique orthogonal decomposition [34, III.4.24]
โ„“ = โ„“0 + ๐ดโ„“ + ๐œ‘โ„“ โˆ™ ๐‘…๐‘ + ๐‘Š โ„“ โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) + ๐‘ โ„“ ,
(4.1)
using the same notation as in (3.7). These processes are unique in the sense
that the integrals are uniquely determined, and so it suces to consider the
left hand side of the BSDE for the notion of a solution. (In BSDE theory, a
solution would be, at least, a quadruple.) We dene the random function
as in Corollary 3.9, with
โˆซ
๐ฟ
replaced by
โ„“.
Since
โ„“
๐‘”โ„“
is special, we have
(โˆฃ๐‘ฅโˆฃ2 + โˆฃ๐‘ฅโ€ฒ โˆฃ2 ) โˆง (1 + โˆฃ๐‘ฅโ€ฒ โˆฃ) ๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )) < โˆž
(4.2)
โ„๐‘‘ ×โ„
and the arguments from Lemma 6.2 show that
โ„ โˆช {sign(๐‘)โˆž}. Hence we
(3.8) with ๐ฟ replaced by โ„“, i.e.,
values in
BSDE
๐‘”โ„“
is well dened on
C0
โ„“ = โ„“0 โˆ’๐‘๐‘ˆ โˆ— (โ„“โˆ’ ) โˆ™ ๐œ‡โˆ’๐‘ max ๐‘” โ„“ (๐‘ฆ) โˆ™ ๐ด+๐œ‘โ„“ โˆ™ ๐‘…๐‘ +๐‘Š โ„“ โˆ—(๐œ‡๐‘… โˆ’๐œˆ ๐‘… )+๐‘ โ„“
๐‘ฆโˆˆC โˆฉC 0
with terminal condition
with
can consider (formally at rst) the
(4.3)
โ„“๐‘‡ = ๐ท ๐‘‡ .
Denition 4.1. A càdlàg special semimartingale โ„“ is called a solution of the
Bellman equation (4.3) if
โˆ™ โ„“, โ„“โˆ’ > 0,
โˆ™
C โˆฉ C 0,โˆ— -valued
< โˆž,
there exists a
supC โˆฉC 0
โˆ™ โ„“
๐‘”โ„“
๐œ‹
ห‡ โˆˆ ๐ฟ(๐‘…)
such that
โ„“๐‘‡ = ๐ท๐‘‡ .
โˆ’1
๐‘) . We call
๐‘” โ„“ (ห‡
๐œ‹) =
and the processes from (4.1) satisfy (4.3) with
strategy associated
If the process
๐œ‹
ห‡
(๐ท/โ„“)๐›ฝ , where
๐œ…
ห‡ :=
with โ„“,
Moreover, we dene
โ„“>0
process
๐›ฝ = (1 โˆ’
and for brevity, we also call
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
(ห‡
๐œ‹, ๐œ…
ห‡)
the
a solution.
is not unique, we choose and x one. The assumption
excludes pathological cases where
โ„“
jumps to zero and becomes posi-
tive immediately afterwards and thereby ensures that
precisely, the following holds.
๐œ…
ห‡
is admissible. More
44
III Bellman Equation
Remark 4.2.
(i)
(ii)
(ห‡
๐œ‹, ๐œ…
ห‡) โˆˆ
Let
supC โˆฉC 0 ๐‘” โ„“
(iii) If
๐‘โˆˆ
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
is a predictable,
(0, 1), ๐‘” โ„“ is nite on
(iv) The condition
๐‘<0
be a solution of the Bellman equation.
๐’œ๐‘“ ๐ธ .
โ„“>0
๐ด-integrable
Cโˆฉ
process.
C 0.
is automatically satised if (a)
๐‘ โˆˆ (0, 1)
or if (b)
and there is no intermediate consumption and Assumptions 3.1
are satised.
Proof.
โˆซ๐‘‡
๐œ…
ห‡ ๐‘  ๐œ‡(๐‘‘๐‘ ) < โˆž ๐‘ƒ -a.s. since the paths of โ„“ are bounded
โˆซ๐‘‡
away from zero. Moreover,
๐œ…๐‘ก ๐‘‹๐‘ก (ห‡
๐œ‹, ๐œ…
ห‡ )) ๐œ‡(๐‘‘๐‘ก) < โˆž as in the proof
0 ๐‘ˆ๐‘ก (ห‡
of (3.3) (stated after Lemma 3.5). This shows (ห‡
๐œ‹, ๐œ…
ห‡ ) โˆˆ ๐’œ๐‘“ . The fact that
(ห‡
๐œ‹, ๐œ…
ห‡ ) โˆˆ ๐’œ๐‘“ ๐ธ is contained in the proof of Lemma 4.9 below.
โ„“
โ„“
โ„“ ๐œ‹ ). Hence sup
โ„“ โˆ™ ๐ด is
(ii) We have 0 = ๐‘” (0) โ‰ค supC โˆฉC 0 ๐‘” = ๐‘” (ห‡
C โˆฉC 0 ๐‘”
(i) We have
0
well dened, and it is nite because otherwise (4.3) could not hold.
(iii)
๐‘”โ„“
<โˆž
Note that
๐‘>0
implies
๐‘” โ„“ > โˆ’โˆž
by its denition and (4.2), while
by assumption.
๐‘ > 0, (4.3) states that ๐ดโ„“
โ„“โˆ’ > 0 implies โ„“ โ‰ฅ 0,
โ„“๐‘‡ = ๐ท๐‘‡ > 0, the minimum
principle for nonnegative supermartingales shows โ„“ > 0. Under (b) the
assertion is a consequence of Theorem 4.5 below (which shows โ„“ โ‰ฅ ๐ฟ > 0)
upon noting that the condition โ„“ > 0 is not used in its proof when there is
(iv) If
โ„“
is decreasing. As
is a supermartingale by Lemma 2.4.
Since
no intermediate consumption.
It may seem debatable to make existence of the maximizer
๐œ‹
ห‡
part of the
denition of a solution. However, associating a control with the solution is
crucial for the following theory. Some justication is given by the following
result for the continuous case (where
C 0,โˆ— = โ„๐‘‘ ).
Proposition 4.3. Let โ„“ be any càdlàg special semimartingale such that
โ„“, โ„“โˆ’ > 0. Under Assumptions 3.11, (C1) and (C2), there exists a C โˆฉ C 0,โˆ— valued predictable process ๐œ‹ห‡ such that ๐‘”โ„“ (ห‡๐œ‹) = supC โˆฉC 0 ๐‘”โ„“ < โˆž, and any
such process is ๐‘…-integrable.
Proof.
As
๐‘”โ„“
is analogous to (3.9), it is continuous and its supremum over
โ„๐‘‘ is nite. By continuity of ๐‘… and the structure condition,
โˆซ๐‘‡ โŠค
โˆซ๐‘‡ โŠค 2
only if
0 ๐œ‹ ๐‘‘โŸจ๐‘€ โŸฉ๐œ‹ = 0 โˆฃ๐œŽ ๐œ‹โˆฃ ๐‘‘๐ด < โˆž ๐‘ƒ -a.s.
๐œ‹ โˆˆ ๐ฟ(๐‘…) if and
C is compact, then Lemma 6.4 yields a measurable
โŠค
๐œ‹ for arg maxC ๐‘” . As in the proof of Corollary 3.12, ๐œŽ โŠค ๐œ‹ โˆˆ ฮ ๐œŽ C ๐œŽ โŠค ๐œ“
(
)
โˆซ
โ„“
๐‘‡
๐œ‘
โŠค 2
for ๐œ“ := ๐›ฝ ๐œ† +
โ„“โˆ’ , which satises 0 โˆฃ๐œŽ ๐œ“โˆฃ ๐‘‘๐ด < โˆž by denition of ๐œ† and
๐œ‘โ„“ . We note that โˆฃ๐œŽ โŠค ๐œ‹โˆฃ โ‰ค โˆฃ๐œŽ โŠค ๐œ“โˆฃ+โˆฃ๐œŽ โŠค ๐œ‹โˆ’๐œŽ โŠค ๐œ“โˆฃ โ‰ค 2โˆฃ๐œŽ โŠค ๐œ“โˆฃ due to the denition
of the projection and 0 โˆˆ C .
In the general case we approximate C by a sequence of compact con๐‘› := C โˆฉ {๐‘ฅ โˆˆ โ„๐‘‘ : โˆฃ๐‘ฅโˆฃ โ‰ค ๐‘›}, each of which yields a selector ๐œ‹ ๐‘›
straints C
โŠค ๐‘›
โŠค
โŠค ๐‘›
for arg maxC ๐‘› ๐‘” . By the above, โˆฃ๐œŽ ๐œ‹ โˆฃ โ‰ค 2โˆฃ๐œŽ ๐œ“โˆฃ, so the sequence (๐œŽ ๐œ‹ )๐‘›
Assume rst that
selector
III.4 Minimality of the Opportunity Process
is bounded for xed
(๐œ”, ๐‘ก).
45
A random index argument as in the proof of
Lemma 6.4 yields a selector
๐œ—
for a cluster point of this sequence. We have
๐œ— โˆˆ ๐œŽ โŠค C by closedness of this set and we nd a selector ๐œ‹
ห‡ for the preimage
((๐œŽ โŠค )โˆ’1 ๐œ—)โˆฉC using [64, 1Q]. We have ๐œ‹
ห‡ โˆˆ arg maxC ๐‘” as the sets C ๐‘› increase
โˆซ๐‘‡ โŠค 2
โˆซ๐‘‡
to C , and
ห‡ โˆฃ ๐‘‘๐ด โ‰ค 2 0 โˆฃ๐œŽ โŠค ๐œ“โˆฃ2 ๐‘‘๐ด < โˆž shows ๐œ‹
ห‡ โˆˆ ๐ฟ(๐‘…).
0 โˆฃ๐œŽ ๐œ‹
Another example for the construction of
๐œ‹
ห‡
is given in Chapter IV.
general, two ingredients are needed: Existence of a maximizer for xed
In
(๐œ”, ๐‘ก)
will typically require a compactness condition in the form of a no-arbitrage
assumption (in the previous proof, this is the structure condition). Moreover,
a measurable selection is required; here the techniques from the appendices
may be useful.
Remark 4.4.
The BSDE formulation of the Bellman equation has the ad-
vantage that we can choose
all solutions.
๐ด
based on
๐‘…
and speak about the class of
However, we do not want to write proofs in this cumber-
some notation. Once we x a solution
โ„“
(and maybe
๐ฟ,
and nitely many
other semimartingales), we can choose a new reference process
๐ดหœ = ๐ด + ๐ดโ€ฒ
โ€ฒ
(where ๐ด is increasing), with respect to which our semimartingales admit
dierential characteristics; in particular we can use the joint characteristics
หœ . As we change ๐ด, all drift rates change in that they
(๐‘๐‘…,โ„“ , ๐‘๐‘…,โ„“ , ๐น ๐‘…,โ„“ ; ๐ด)
หœ
are multiplied by ๐‘‘๐ด/๐‘‘๐ด
, so any (in)equalities between them are preserved.
With this in mind, we shall use the joint characteristics of (๐‘…, โ„“) in the sequel without further comment and treat the two formulations of the Bellman
equation as equivalent.
Our denition of a solution of the Bellman equation is loose in terms of
integrability assumptions.
Even in the continuous case, it is unclear how
many solutions exist. The next result shows that we can always identify
by taking the smallest one, i.e.,
๐ฟโ‰คโ„“
for any solution
๐ฟ
โ„“.
Theorem 4.5. Under Assumptions 3.1, the opportunity process ๐ฟ is characterized as the minimal solution of the Bellman equation.
Remark 4.6. As a consequence, the Bellman equation has a bounded solution if and only if the opportunity process is bounded (and similarly for
other integrability properties). In conjunction with Section II.4.2 this yields
examples of quadratic BSDEs which have bounded terminal value (for
๐ท๐‘‡
bounded), but no bounded solution.
The proof of Theorem 4.5 is based on the following result; it is the fundamental property of any Bellman equation.
Proposition 4.7. Let
any (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ,
๐‘(๐œ‹, ๐œ…) :=
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
โ„“ ๐‘1
(
be a solution of the Bellman equation. For
)๐‘
๐‘‹(๐œ‹, ๐œ…) +
โˆซ
(
)
๐‘ˆ๐‘  ๐œ…๐‘  ๐‘‹๐‘  (๐œ‹, ๐œ…) ๐œ‡(๐‘‘๐‘ )
(4.4)
III Bellman Equation
46
is a semimartingale with nonpositive drift rate. Moreover, ๐‘(ห‡๐œ‹, ๐œ…ห‡ ) is a local
martingale.
Proof.
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ .
Let
Note that
hence has a well dened drift rate
calculated as in Lemma 3.4: If
that lemma but with
๐ฟ
๐‘“โ„“
๐‘ := ๐‘(๐œ‹, ๐œ…)
๐‘Ž๐‘
satises
sign(๐‘)๐‘ โ‰ฅ 0,
by Remark 2.3. The drift rate can be
is dened similarly to the function
replaced by
โ„“,
๐‘“
in
then
{
}
๐‘‘๐œ‡
๐‘Ž๐‘ = ๐‘‹(๐œ‹, ๐œ…)๐‘โˆ’ ๐‘โˆ’1 ๐‘Žโ„“ + ๐‘“ โ„“ (๐œ…) ๐‘‘๐ด
+ ๐‘” โ„“ (๐œ‹)
{(
) ๐‘‘๐œ‡
}
= ๐‘‹(๐œ‹, ๐œ…)๐‘โˆ’ ๐‘“ โ„“ (๐œ…) โˆ’ ๐‘“ โ„“ (ห‡
๐œ…) ๐‘‘๐ด
+ ๐‘” โ„“ (๐œ‹) โˆ’ ๐‘” โ„“ (ห‡
๐œ‹) .
This is nonpositive because
case
(๐œ‹, ๐œ…) := (ห‡
๐œ‹, ๐œ…
ห‡ ) we have
sign(๐‘)๐‘ โ‰ฅ 0.
๐œ…
ห‡ and ๐œ‹
ห‡ maximize ๐‘“ โ„“ and ๐‘” โ„“ . For the special
๐‘
๐‘Ž = 0 and so ๐‘ is a ๐œŽ -martingale, thus a local
martingale as
Remark 4.8.
In Proposition 4.7, semimartingale with nonpositive drift
rate can be replaced by ๐œŽ -supermartingale if
๐‘”โ„“
is nite on
C โˆฉ C 0.
Theorem 4.5 follows from the next lemma (which is actually stronger).
We recall that for
๐‘<0
the opportunity process
๐ฟ
can be dened without
further assumptions.
Lemma 4.9. Let โ„“ be a solution of the Bellman equation. If ๐‘ < 0, then
For ๐‘ โˆˆ (0, 1), the same holds if (2.2) is satised and there exists an
optimal strategy.
๐ฟ โ‰ค โ„“.
Proof. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡ ) be a solution and dene ๐‘(๐œ‹, ๐œ…) as in (4.4).
Case ๐‘ < 0: We choose (๐œ‹, ๐œ…) := (ห‡๐œ‹, ๐œ…ห‡ ). As ๐‘(ห‡๐œ‹, ๐œ…ห‡ ) is a negative
cal martingale by Proposition 4.7, it is a submartingale.
๐ธ[๐‘๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ )] > โˆ’โˆž,
mark 4.2(i). Recall that
ห‡๐‘
โ„“๐‘ก ๐‘1 ๐‘‹
๐‘ก
โ„“๐‘‡ = ๐ท๐‘‡ = ๐ฟ๐‘‡ ,
โˆซ
+
0
In particular,
๐ฟ๐‘‡ = ๐ท๐‘‡ , this is the statement that the
i.e., (ห‡
๐œ‹, ๐œ…
ห‡ ) โˆˆ ๐’œ๐‘“ ๐ธ this completes the proof of Reห‡ := ๐‘‹(ห‡
ห‡,
๐œ‡โˆ˜ = ๐œ‡ + ๐›ฟ{๐‘‡ } . With ๐‘‹
๐œ‹, ๐œ…
ห‡ ) and ๐‘ห‡ := ๐œ…
ห‡๐‘‹
and using
expected utility is nite,
and using
lo-
we deduce
๐‘ก
]
[
๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ) = ๐‘๐‘ก (ห‡
๐œ‹, ๐œ…
ห‡ ) โ‰ค ๐ธ ๐‘๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ )โ„ฑ๐‘ก
] โˆซ ๐‘ก
[โˆซ ๐‘‡
โˆ˜
โ‰ค ess sup๐‘หœโˆˆ๐’œ(ห‡๐œ‹,ห‡๐‘,๐‘ก) ๐ธ
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก +
๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ )
๐‘ก
0
โˆซ ๐‘ก
1 ห‡๐‘
= ๐ฟ๐‘ก ๐‘ ๐‘‹๐‘ก +
๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ),
0
Case ๐‘ โˆˆ (0, 1):
Then
We choose
1 ห‡๐‘
๐‘ ๐‘‹๐‘ก
< 0, we have โ„“๐‘ก โ‰ฅ ๐ฟ๐‘ก .
(๐œ‹, ๐œ…) := (ห†
๐œ‹, ๐œ…
ห† ) to be an optimal strategy.
where the last equality holds by (2.3). As
๐‘(ห†
๐œ‹, ๐œ…
ห† ) โ‰ฅ 0 is a supermartingale by Proposition 4.7 and Lemma 2.4(iii),
III.5 Verication
47
and we obtain
ห†๐‘
โ„“๐‘ก ๐‘1 ๐‘‹
๐‘ก
โˆซ
+
0
๐‘ก
]
[
๐‘ˆ๐‘  (ห†
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ) = ๐‘๐‘ก (ห†
๐œ‹, ๐œ…
ห† ) โ‰ฅ ๐ธ ๐‘๐‘‡ (ห†
๐œ‹, ๐œ…
ห† )โ„ฑ๐‘ก
โˆซ ๐‘ก
]
[โˆซ ๐‘‡
โˆ˜
1 ห†๐‘
๐‘ˆ๐‘  (ห†
๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก = ๐ฟ๐‘ก ๐‘ ๐‘‹๐‘ก +
=๐ธ
๐‘ˆ๐‘  (ห†
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ )
0
by the optimality of
that
(ห†
๐œ‹, ๐œ…
ห†)
0
(ห†
๐œ‹, ๐œ…
ห†)
and (2.3). More precisely, we have used the fact
is also conditionally optimal (see Remark II.3.3). As
we conclude
1 ห†๐‘
๐‘ ๐‘‹๐‘ก
> 0,
โ„“๐‘ก โ‰ฅ ๐ฟ ๐‘ก .
III.5 Verication
Suppose that we have found a solution of the Bellman equation; then we want
to know whether it is the opportunity process and whether the associated
strategy is optimal. In applications, it might not be clear
a priori that an op-
timal strategy exists or even that the utility maximization problem is nite.
Therefore, we stress that in this section these properties are not assumed.
Also, we do not need the assumptions on
C
made in Section III.2.4they
are not necessary because we start with a given solution.
Generally speaking, verication involves the candidate for an optimal
control,
(ห‡
๐œ‹, ๐œ…
ห‡)
in our case, and all the competing ones.
It is often very
dicult to check a condition involving all these controls, so it is desirable to
have a verication theorem whose assumptions involve only
(ห‡
๐œ‹, ๐œ…
ห‡ ).
We present two verication approaches. The rst one is via the value process and is classical for general dynamic programming: it uses little structure
of the given problem. For
๐‘ โˆˆ (0, 1),
it yields the desired result. However,
in a general setting, this is not the case for
๐‘ < 0.
The second approach
uses the concavity of the utility function. To fully exploit this and make the
verication conditions necessary, we will assume that
C
is convex. In this
case, we shall obtain the desired verication theorem for all values of
๐‘.
III.5.1 Verication via the Value Process
The basis of this approach is the following simple result; we state it separately
for better comparison with Lemma 5.10 below. In the entire section,
is dened by (4.4) whenever
โ„“
๐‘(๐œ‹, ๐œ…)
is given.
Lemma 5.1. Let โ„“ be any positive càdlàg semimartingale with โ„“๐‘‡ = ๐ท๐‘‡ and
let (ห‡๐œ‹, ๐œ…ห‡ ) โˆˆ ๐’œ. Assume that for all (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ , the process ๐‘(๐œ‹, ๐œ…) is a
supermartingale. Then ๐‘(ห‡๐œ‹, ๐œ…ห‡ ) is a martingale if and only if (2.2) holds and
(ห‡
๐œ‹, ๐œ…
ห‡ ) is optimal and โ„“ = ๐ฟ.
Proof. โ‡’:
that
Recall that
๐‘0 (๐œ‹, ๐œ…) = โ„“0 ๐‘1 ๐‘ฅ๐‘0
does not depend on
โˆซ๐‘‡
๐ธ[๐‘๐‘‡ (๐œ‹, ๐œ…)] = ๐ธ[ 0 ๐‘ˆ๐‘ก (๐œ…๐‘ก (๐‘‹๐‘ก (๐œ‹, ๐œ…))) ๐œ‡โˆ˜ (๐‘‘๐‘ก)]
(๐œ‹, ๐œ…)
and
is the expected utility cor-
48
III Bellman Equation
ห‡ := ๐‘‹(ห‡
(๐œ‹, ๐œ…). With ๐‘‹
๐œ‹, ๐œ…
ห‡ ), the (super)martingale condiโˆซ๐‘‡
โˆซ๐‘‡
โˆ˜
ห‡
tion implies that ๐ธ[
๐œ…๐‘ก ๐‘‹๐‘ก ) ๐œ‡ (๐‘‘๐‘ก)] โ‰ฅ ๐ธ[ 0 ๐‘ˆ๐‘ก (๐œ…๐‘ก ๐‘‹๐‘ก (๐œ‹, ๐œ…)) ๐œ‡โˆ˜ (๐‘‘๐‘ก)] for
0 ๐‘ˆ๐‘ก (ห‡
๐‘“ ๐ธ . Since for (๐œ‹, ๐œ…) โˆˆ ๐’œ โˆ– ๐’œ๐‘“ ๐ธ the expected utility is โˆ’โˆž, this
all (๐œ‹, ๐œ…) โˆˆ ๐’œ
shows that (ห‡
๐œ‹, ๐œ…
ห‡ ) is optimal with ๐ธ[๐‘๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ )] = ๐‘0 (ห‡
๐œ‹, ๐œ…
ห‡ ) = โ„“0 ๐‘1 ๐‘ฅ๐‘0 < โˆž. In
particular, the opportunity process ๐ฟ is well dened. By Proposition 2.2,
โˆซ
ห‡ ๐‘ + ๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡(๐‘‘๐‘ ) is a martingale, and as its terminal value equals
๐ฟ ๐‘1 ๐‘‹
ห‡ > 0.
๐‘๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ ), we deduce โ„“ = ๐ฟ by comparison with (4.4), using ๐‘‹
responding to
The converse is contained in Proposition 2.2.
We can now state our rst verication theorem.
Theorem 5.2. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solution of the Bellman equation.
(i) If ๐‘ โˆˆ (0, 1), the following are equivalent:
(a) ๐‘(ห‡๐œ‹, ๐œ…ห‡ ) is of class (D),
(b) ๐‘(ห‡๐œ‹, ๐œ…ห‡ ) is a martingale,
(c) (2.2) holds and (ห‡๐œ‹, ๐œ…ห‡ ) is optimal and โ„“ = ๐ฟ.
(ii) If ๐‘ < 0, the following are equivalent:
(a) ๐‘(๐œ‹, ๐œ…) is of class (D) for all (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ ,
(b) ๐‘(๐œ‹, ๐œ…) is a supermartingale for all (๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ ,
(c) (ห‡๐œ‹, ๐œ…ห‡ ) is optimal and โ„“ = ๐ฟ.
Proof.
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ , ๐‘(๐œ‹, ๐œ…) is positive and ๐‘Ž๐‘(๐œ‹,๐œ…) โ‰ค 0 by
Proposition 4.7, hence ๐‘(๐œ‹, ๐œ…) is a supermartingale according to Lemma 2.4.
By Proposition 4.7, ๐‘(ห‡
๐œ‹, ๐œ…
ห‡ ) is a local martingale, so it is a martingale if and
When
๐‘>0
and
only if it is of class (D). Lemma 5.1 implies the result.
If
๐‘ < 0, ๐‘(๐œ‹, ๐œ…)
is negative.
Thus the local martingale
๐‘(ห‡
๐œ‹, ๐œ…
ห‡)
is a
submartingale, and a martingale if and only if it is also a supermartingale.
Note that a class (D) semimartingale with nonpositive drift rate is a supermartingale. Conversely, any negative supermartingale
to the bounds
that if
โ„“ = ๐ฟ,
0 โ‰ฅ ๐‘ โ‰ฅ ๐ธ[๐‘๐‘‡ โˆฃ๐”ฝ].
๐‘<0
is of class (D) due
then Proposition 2.2 yields (b).
Theorem 5.2 is as good as it gets for
for
๐‘
Lemma 5.1 implies the result after noting
๐‘ > 0, but as announced, the result
is not satisfactory. In particular settings, this can be improved.
Remark 5.3
< 0). (i) Assume we know a priori that if there
(ห†
๐œ‹, ๐œ…
ห† ) โˆˆ ๐’œ, then
{
}
(ห†
๐œ‹, ๐œ…
ห† ) โˆˆ ๐’œ(๐ท) := (๐œ‹, ๐œ…) โˆˆ ๐’œ : ๐‘‹(๐œ‹, ๐œ…)๐‘ is of class (D) .
(๐‘
is an
optimal strategy
๐’œ(๐ท) . If
when ๐‘ < 0),
In this case we can reduce our optimization problem to the class
furthermore
โ„“
is bounded (which is not a strong assumption
the class (D) condition in Theorem 5.2(ii) is automatically satised for any
(๐œ‹, ๐œ…) โˆˆ ๐’œ(๐ท) .
The verication then reduces to checking that
(ห‡
๐œ‹, ๐œ…
ห‡ ) โˆˆ ๐’œ(๐ท) .
III.5 Verication
(ii)
49
How can we establish the condition needed for (i)? One possibility
is to show that
๐ฟ
is uniformly bounded away from zero; then the condition
follows (see the argument in the next proof ).
Of course,
๐ฟ
is not known
when we try to apply this. However, Section II.4.2 gives veriable conditions
for
๐ฟ
to be (bounded and) bounded away from zero.
C = โ„๐‘‘ ,
They are stated for
๐‘‘
๐ฟโ„ is the
๐‘‘
โ„๐‘‘
opportunity process corresponding to C = โ„ , the actual ๐ฟ satises ๐ฟ โ‰ฅ ๐ฟ
the unconstrained case
but can be used nevertheless: if
because the supremum in (2.3) is taken over a smaller set in the constrained
case.
In the situation where
โ„“
and
๐ฟโˆ’1
are bounded, we can also use the fol-
lowing result. Note also its use in Remark 3.3(ii) and recall that
1/0 := โˆž.
Corollary 5.4. Let ๐‘ < 0 and let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solution of the Bellman
equation. Let ๐ฟ be the opportunity process and assume that โ„“/๐ฟ is uniformly
bounded. Then (ห‡๐œ‹, ๐œ…ห‡ ) is optimal and โ„“ = ๐ฟ.
Proof.
(
๐ฟ ๐‘1
(๐œ‹, ๐œ…) โˆˆ ๐’œ๐‘“ ๐ธ
)๐‘ โˆซ
๐‘‹(๐œ‹, ๐œ…) + ๐‘ˆ๐‘  (๐œ…๐‘  ๐‘‹๐‘  ) ๐œ‡(๐‘‘๐‘ ) is
Fix arbitrary
and let
๐‘‹ = ๐‘‹(๐œ‹, ๐œ…).
The process
a negative supermartingale by Propo-
โˆซ
๐‘ˆ๐‘  (๐œ…๐‘  ๐‘‹๐‘  ) ๐œ‡(๐‘‘๐‘ ) is decreasing and its
๐‘“ ๐ธ ), ๐ฟ 1 ๐‘‹ ๐‘ is also of class (D).
terminal value is integrable (denition of ๐’œ
๐‘
1 ๐‘
The assumption yields that โ„“ ๐‘‹ is of class (D), and then so is ๐‘(๐œ‹, ๐œ…).
๐‘
sition 2.2, hence of class (D). Since
As bounded solutions are of special interest in BSDE theory, let us note
the following consequence.
Corollary 5.5. Let ๐‘ < 0. Under Assumptions 3.1 the following are equivalent:
(i) ๐ฟ is bounded and bounded away from zero.
(ii) There exists a unique bounded solution of the Bellman equation, and
this solution is bounded away from zero.
One can note that in the setting of Section II.4.2, these conditions are
further equivalent to a reverse Hölder inequality for the market model.
We give an illustration of Theorem 5.2 also for the case
far, we have considered only the given exponent
many situations, there will exist some
the exponent
๐‘0
instead of
๐‘,
๐‘0 โˆˆ (๐‘, 1)
๐‘
๐‘ โˆˆ (0, 1).
Thus
and assumed (2.2).
In
such that, if we consider
the utility maximization problem is still nite.
Note that by Jensen's inequality this is a stronger assumption. We dene
for
๐‘ž0 โ‰ฅ 1
the class of semimartingales
โ„“
bounded in
๐ฟ๐‘ž0 (๐‘ƒ ),
B(๐‘ž0 ) := {โ„“ : sup๐œ โˆฅโ„“๐œ โˆฅ๐ฟ๐‘ž0 (๐‘ƒ ) < โˆž},
where the supremum ranges over all stopping times
๐œ.
50
III Bellman Equation
Corollary 5.6. Let ๐‘ โˆˆ (0, 1) and let there be a constant ๐‘˜1 > 0 such that
๐ท โ‰ฅ ๐‘˜1 . Assume that the utility maximization problem is nite for some
๐‘0 โˆˆ (๐‘, 1) and let ๐‘ž0 โ‰ฅ 1 be such that ๐‘ž0 > ๐‘0 /(๐‘0 โˆ’ ๐‘). If (โ„“, ๐œ‹
ห‡, ๐œ…
ห‡ ) is a
solution of the Bellman equation (for ๐‘) with โ„“ โˆˆ B(๐‘ž0 ), then โ„“ = ๐ฟ and
(ห‡
๐œ‹, ๐œ…
ห‡ ) is optimal.
Proof.
Let
๐‘‹(ห‡
๐œ‹, ๐œ…
ห‡ ).
โ„“ โˆˆ B(๐‘ž0 )
be a solution,
(ห‡
๐œ‹, ๐œ…
ห‡)
the associated strategy, and
ห‡ =
๐‘‹
By Theorem 5.2 and an argument as in the previous proof, it suces
to show that
ห‡๐‘
โ„“๐‘‹
is of class (D). Let
For every stopping time
๐œ,
๐›ฟ>1
be such that
๐›ฟ/๐‘ž0 + ๐›ฟ๐‘/๐‘0 = 1.
Hölder's inequality yields
ห‡ ๐‘ )๐›ฟ ] = ๐ธ[(โ„“๐‘ž0 )๐›ฟ/๐‘ž0 (๐‘‹
ห‡ ๐‘0 )๐›ฟ๐‘/๐‘0 ] โ‰ค ๐ธ[โ„“๐‘ž0 ]๐›ฟ/๐‘ž0 ๐ธ[๐‘‹
ห‡ ๐œ๐‘0 ]๐›ฟ๐‘/๐‘0 .
๐ธ[(โ„“๐œ ๐‘‹
๐œ
๐œ
๐œ
๐œ
ห‡ ๐œ๐‘
๐œ ; then {โ„“๐œ ๐‘‹
๐›ฟ
is bounded in ๐ฟ (๐‘ƒ ) and hence uniformly integrable.
We show that this is bounded uniformly in
: ๐œ
stopping time}
Indeed,
๐ธ[โ„“๐‘ž๐œ0 ]
is
bounded by assumption. The set of wealth processes corresponding to admis-
ห‡ ๐œ๐‘0 ] โ‰ค ๐‘ข(๐‘0 ) (๐‘ฅ0 ),
๐ธ[๐ท๐‘‡ ๐‘10 ๐‘‹
problem with exponent ๐‘0 .
sible strategies is stable under stopping. Therefore
the value function for the utility maximization
The result follows as
Remark 5.7.
๐ท๐‘‡ โ‰ฅ ๐‘˜1 .
In Example II.4.6 we give a condition which implies that the
utility maximization problem is nite for
such a
all ๐‘0 โˆˆ (0, 1).
๐‘0 โˆˆ (๐‘, 1), one can show that ๐ฟ โˆˆ B(๐‘0 /๐‘) if ๐ท
Conversely, given
is uniformly bounded
from above (see Corollary V.4.2).
Example 5.8.
We apply our results in an Itô model with bounded mean
variance tradeo process together with an existence result for BSDEs. For
the case of utility from terminal wealth only, we retrieve (a minor generalization of ) the pioneering result of [33, ยŸ3]; the case with intermediate consumption is new. Let
(๐‘š
โ‰ฅ ๐‘‘)
๐‘Š
and assume that
๐‘š-dimensional
generated by ๐‘Š .
be an
๐”ฝ
is
standard Brownian motion
We consider
๐‘‘๐‘…๐‘ก = ๐‘๐‘ก ๐‘‘๐‘ก + ๐œŽ๐‘ก ๐‘‘๐‘Š๐‘ก ,
โ„๐‘‘×๐‘š -valued with
everywhere full rank; moreover, we consider constraints C satisfying (C1)
and (C2). We are in the situation of Assumptions III.3.3 with ๐‘‘๐‘€ = ๐œŽ ๐‘‘๐‘Š
โŠค โˆ’1 ๐‘. The process ๐œƒ := ๐œŽ โŠค ๐œ† is called market price of risk. We
and ๐œ† = (๐œŽ๐œŽ )
assume that there are constants ๐‘˜๐‘– > 0 such that
โˆซ ๐‘‡
0 < ๐‘˜1 โ‰ค ๐ท โ‰ค ๐‘˜2 and
โˆฃ๐œƒ๐‘  โˆฃ2 ๐‘‘๐‘  โ‰ค ๐‘˜3 .
where
๐‘
is predictable
โ„๐‘‘ -valued
and
๐œŽ
is predictable
0
The latter condition is called
bounded mean-variance tradeo.
Note that
๐‘‘๐‘„/๐‘‘๐‘ƒ = โ„ฐ(โˆ’๐œ† โˆ™ ๐‘€ )๐‘‡ = โ„ฐ(โˆ’๐œƒ โˆ™ ๐‘Š )๐‘‡ denes a local martingale measure
for โ„ฐ(๐‘…). By Section II.4.2 the utility maximization problem is nite for all
III.5 Verication
๐‘
and the opportunity process
๐ฟ
51
is bounded and bounded away from zero.
It is continuous due to Remark 3.13(i).
๐‘Œ := log(๐ฟ) rather
๐‘Œ = ๐ด๐‘Œ + ๐œ‘๐‘Œ โˆ™ ๐‘€ + ๐‘ ๐‘Œ is the Kunita-Watanabe
โŠค ๐‘Œ
โŠฅ
โŠฅ โˆ™ ๐‘Š = ๐‘๐‘Œ
decomposition, we write ๐‘ := ๐œŽ ๐œ‘ and choose ๐‘ such that ๐‘
As suggested above, we write the Bellman BSDE for
๐ฟ
than
in this setting. If
by Brownian representation. The orthogonality of the decomposition implies
๐œŽโŠค๐‘ โŠฅ = 0
consumption and
(with
๐‘ โŠค ๐‘ โŠฅ = 0. We write ๐›ฟ = 1 if there is intermediate
๐›ฟ = 0 otherwise. Then Itô's formula and Corollary 3.12
and that
๐ด๐‘ก := ๐‘ก)
yield the BSDE
๐‘‘๐‘Œ = ๐‘“ (๐‘Œ, ๐‘, ๐‘ โŠฅ ) ๐‘‘๐‘ก + (๐‘ + ๐‘ โŠฅ ) ๐‘‘๐‘Š ;
๐‘Œ๐‘‡ = log(๐ท๐‘‡ )
(5.1)
with
(
)
๐‘“ (๐‘Œ, ๐‘, ๐‘ โŠฅ ) = 12 ๐‘(1 โˆ’ ๐‘) ๐‘‘2๐œŽโŠค C ๐›ฝ(๐œƒ + ๐‘) + 2๐‘ž โˆฃ๐œƒ + ๐‘โˆฃ2
(
)
+ ๐›ฟ(๐‘ โˆ’ 1)๐ท๐›ฝ exp (๐‘ž โˆ’ 1)๐‘Œ โˆ’ 12 (โˆฃ๐‘โˆฃ2 + โˆฃ๐‘ โŠฅ โˆฃ2 ).
๐›ฝ = (1 โˆ’ ๐‘)โˆ’1
๐‘ž = ๐‘/(๐‘ โˆ’ 1); the dependence on (๐œ”, ๐‘ก) is suppressed
2
in the notation. Using the orthogonality relations and ๐‘(1 โˆ’ ๐‘)๐›ฝ = โˆ’๐‘ž , one
โŠฅ
โŠฅ
โŠฅ
หœ
หœ
can check that ๐‘“ (๐‘Œ, ๐‘, ๐‘ ) = ๐‘“ (๐‘Œ, ๐‘ +๐‘ , 0) =: ๐‘“ (๐‘Œ, ๐‘), where ๐‘ := ๐‘ +๐‘ .
2
2
As 0 โˆˆ C , we have ๐‘‘ โŠค (๐‘ฅ) โ‰ค โˆฃ๐‘ฅโˆฃ . Hence there exist a constant ๐ถ > 0 and
๐œŽ C
an increasing continuous function ๐œ™ such that
(
)
โˆฃ๐‘“ (๐‘ฆ, ๐‘งหœ)โˆฃ โ‰ค ๐ถ โˆฃ๐œƒโˆฃ2 + ๐œ™(๐‘ฆ) + โˆฃหœ
๐‘ง โˆฃ2 .
Here
and
The following monotonicity property handles the exponential nonlinearity
caused by the consumption: As
๐‘โˆ’1<0
and
๐‘ž โˆ’ 1 < 0,
[
]
โˆ’๐‘ฆ ๐‘“ (๐‘ฆ, ๐‘งหœ) โˆ’ ๐‘“ (0, ๐‘งหœ) โ‰ค 0.
Thus we have Briand and Hu's [9, Condition (A.1)] after noting that they
call
โˆ’๐‘“
what we call
and [9, Lemma 2] states the existence of a bounded
โ„“ := exp(๐‘Œ ) is the op(ห‡
๐œ‹, ๐œ…
ห‡ ) by ๐œ…
ห‡ := (๐ท/โ„“)๐›ฝ
and Proposition 4.3; then we have a solution (โ„“, ๐œ‹
ห‡, ๐œ…
ห‡ ) of the Bellman equation in the sense of Denition 4.1. For ๐‘ < 0 (๐‘ โˆˆ (0, 1)), Corollary 5.4
(Corollary 5.6) yields โ„“ = ๐ฟ and the optimality of (ห‡
๐œ‹, ๐œ…
ห‡ ). In fact, the same
verication argument applies if we replace ๐œ‹
ห‡ by any other predictable C โˆ—
โŠค โˆ—
๐œŽ โŠค C {๐›ฝ(๐œƒ + ๐‘)}; recall from Proposition 4.3
valued ๐œ‹ such that ๐œŽ ๐œ‹ โˆˆ ฮ 
โˆ—
that ๐œ‹ โˆˆ ๐ฟ(๐‘…) automatically. To conclude: we have that
solution
๐‘Œ
๐‘“,
to the BSDE (5.1).
Let us check that
portunity process. We dene an associated strategy
๐ฟ = exp(๐‘Œ )
is the opportunity process
and the set of optimal strategies equals the set of all
โˆ™ ๐œ…
ห†=
โˆ™ ๐œ‹โˆ—
(๐ท/๐ฟ)๐›ฝ
(๐œ‹ โˆ— , ๐œ…
ห†)
such that
๐œ‡โˆ˜ -a.e.
is predictable,
C -valued
and
๐œŽ โŠค ๐œ‹ โˆ— โˆˆ ฮ ๐œŽ
โŠคC
{๐›ฝ(๐œƒ + ๐‘)} ๐‘ƒ โŠ— ๐‘‘๐‘ก-a.e.
52
III Bellman Equation
One can remark that the previous arguments show
๐‘Œ
๐‘Œ โ€ฒ = log(๐ฟ)
whenever
โ€ฒ is a solution of the BSDE (5.1) which is uniformly bounded from above.
Hence we have proved uniqueness for (5.1) in this class of solutions, which
is not immediate from BSDE theory.
to [33], we did not use the theory of
One can also note that, in contrast
๐ต๐‘€ ๐‘‚
martingales in this example.
We close this section with a formula intended for future applications.
Remark 5.9.
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡ ) be a solution of the Bellman equation. Sometimes
๐‘(ห‡
๐œ‹, ๐œ…
ห‡ ) is of class (D).
โŠค
Let โ„Ž be a predictable cut-o function such that ๐œ‹
ห‡ โ„Ž(๐‘ฅ) is bounded, e.g.,
โ„Ž(๐‘ฅ) = ๐‘ฅ1{โˆฃ๐‘ฅโˆฃโ‰ค1}โˆฉ{โˆฃห‡๐œ‹โŠค ๐‘ฅโˆฃโ‰ค1} , and dene ฮจ to be the local martingale
Let
exponential formulas can be used to verify that
โˆ™ ๐‘€ โ„“ + ๐‘ห‡
โ„“โˆ’1
๐œ‹ โˆ™ ๐‘…๐‘ + ๐‘ห‡
๐œ‹ โŠค โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) + ๐‘(๐‘ฅโ€ฒ /โ„“โˆ’ )ห‡
๐œ‹ โŠค โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘…,โ„“ โˆ’ ๐œˆ ๐‘…,โ„“ )
โˆ’
{
}
+ (1 + ๐‘ฅโ€ฒ /โ„“โˆ’ ) (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ’ ๐‘ห‡
๐œ‹ โŠค โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘…,โ„“ โˆ’ ๐œˆ ๐‘…,โ„“ ).
Then
โ„ฐ(ฮจ) > 0,
Proof.
Let
and if
โ„ฐ(ฮจ)
๐‘ = ๐‘(ห‡
๐œ‹, ๐œ…
ห‡ ).
is of class (D), then
๐‘(ห‡
๐œ‹, ๐œ…
ห‡)
By a calculation as in the proof of Lemma 3.4 and
the local martingale condition from Proposition 4.7,
Hence
๐‘ = ๐‘0 โ„ฐ(ฮจ)
is also of class (D).
ห‡ ๐‘ )โˆ’1 โˆ™ ๐‘ = โ„“โˆ’ โˆ™ ฮจ.
( ๐‘1 ๐‘‹
โˆ’
in the case without intermediate consumption.
For
๐‘ is of
ห‡ ๐‘ is. Writing the denition of ๐œ…
โ„“ ๐‘1 ๐‘‹
ห‡ as ๐œ…
ห‡ ๐‘โˆ’1 = โ„“โˆ’ /๐ท
โˆซ
ห‡๐‘ = ๐‘ โˆ’ ๐œ…
ห‡ ๐‘ ๐‘‘๐œ‡ = (โ„“โˆ’ 1 ๐‘‹
ห‡๐‘ โˆ™
๐œ‡-a.e., we have โ„“ ๐‘1 ๐‘‹
ห‡ โ„“โˆ’ ๐‘1 ๐‘‹
ห‡ โˆ™ ๐œ‡), hence
๐‘ โˆ’ ) (ฮจ โˆ’ ๐œ…
1 ห‡๐‘
ห‡ โˆ™ ๐œ‡) = ๐‘0 โ„ฐ(ฮจ) exp(โˆ’ห‡
๐œ… โˆ™ ๐œ‡). It remains to note that
โ„“ ๐‘ ๐‘‹ = ๐‘0 โ„ฐ(ฮจ โˆ’ ๐œ…
โˆ™
exp(โˆ’ห‡
๐œ… ๐œ‡) โ‰ค 1.
the general case, we have seen in the proof of Corollary 5.4 that
class (D) whenever
III.5.2 Verication via Deator
The goal of this section is a verication theorem which involves only the candidate for the optimal strategy and holds for general semimartingale models.
Our plan is as follows. Let
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
C
and assume for the moment that
has a maximum at
๐œ‹
ห‡,
be a solution of the Bellman equation
is convex. As the concave function
the directional derivatives at
๐œ‹
ห‡
๐‘”โ„“
in all directions should
be nonpositive (if they can be dened). A calculation will show that, at the
level of processes, this yields a supermartingale property which is well known
from duality theory and allows for verication.
In the case of non-convex
constraints, the directional derivatives need not be dened in any sense. Nevertheless, the formally corresponding quantities yield the expected result. To
make the rst order conditions necessary, we later specialize to convex
C.
As
in the previous section, we rst state a basic result; it is essentially classical.
Lemma 5.10. Let โ„“ be any positive càdlàg semimartingale with โ„“๐‘‡ = ๐ท๐‘‡ .
Suppose there exists (ห‡๐œ‹, ๐œ…ห‡ ) โˆˆ ๐’œ with ๐œ…ห‡ = (๐ท/โ„“)๐›ฝ and let ๐‘‹ห‡ := ๐‘‹(ห‡๐œ‹, ๐œ…ห‡ ).
Assume ๐‘Œ := โ„“๐‘‹ห‡ ๐‘โˆ’1 has the property that for all (๐œ‹, ๐œ…) โˆˆ ๐’œ,
โˆซ
ฮ“(๐œ‹, ๐œ…) := ๐‘‹(๐œ‹, ๐œ…)๐‘Œ +
๐œ…๐‘  ๐‘‹๐‘  (๐œ‹, ๐œ…)๐‘Œ๐‘  ๐œ‡(๐‘‘๐‘ )
III.5 Verication
is a supermartingale. Then ฮ“(ห‡๐œ‹, ๐œ…ห‡ ) is a martingale if and only if
and (ห‡๐œ‹, ๐œ…ห‡ ) is optimal and โ„“ = ๐ฟ.
Proof. โ‡’:
(2.2)
53
holds
ห‡ . Note
(๐œ‹, ๐œ…) โˆˆ ๐’œ and denote ๐‘ = ๐œ…๐‘‹(๐œ‹, ๐œ…) and ๐‘ห‡ = ๐œ…
ห‡๐‘‹
๐‘โˆ’1
๐‘โˆ’1
๐‘โˆ’1
ห‡
ห‡
the partial derivative โˆ‚๐‘ˆ (ห‡
๐‘) = ๐ทห‡
๐œ… ๐‘‹
= โ„“๐‘‹
= ๐‘Œ . Concavity of ๐‘ˆ
implies ๐‘ˆ (๐‘) โˆ’ ๐‘ˆ (ห‡
๐‘) โ‰ค โˆ‚๐‘ˆ (ห‡
๐‘)(๐‘ โˆ’ ๐‘ห‡) = ๐‘Œ (๐‘ โˆ’ ๐‘ห‡), hence
]
]
[โˆซ ๐‘‡
]
[โˆซ ๐‘‡
[โˆซ ๐‘‡
โˆ˜
โˆ˜
๐‘Œ๐‘  (๐‘๐‘  โˆ’ ๐‘ห‡๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )
๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ ) โ‰ค ๐ธ
๐‘ˆ๐‘  (๐‘๐‘  ) ๐œ‡ (๐‘‘๐‘ ) โˆ’ ๐ธ
๐ธ
Let
0
0
0
= ๐ธ[ฮ“๐‘‡ (๐œ‹, ๐œ…)] โˆ’ ๐ธ[ฮ“๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ )].
Let
ฮ“(ห‡
๐œ‹, ๐œ…
ห‡)
be a martingale; then
ฮ“0 (๐œ‹, ๐œ…) = ฮ“0 (ห‡
๐œ‹, ๐œ…
ห‡)
and the supermartin-
(๐œ‹, ๐œ…) was arbitrary,
[โˆซ๐‘‡
]
(ห‡
๐œ‹, ๐œ…
ห‡ ) is optimal with expected utility ๐ธ 0 ๐‘ˆ๐‘  (ห‡
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ ) = ๐ธ[ ๐‘1 ฮ“๐‘‡ (ห‡
๐œ‹, ๐œ…
ห‡ )] =
1
1 ๐‘
๐œ‹, ๐œ…
ห‡ ) = ๐‘ ๐‘ฅ0 โ„“0 < โˆž. The rest is as in the proof of Lemma 5.1.
๐‘ ฮ“0 (ห‡
gale property imply that the last line is nonpositive. As
The process
๐‘Œ
is a supermartingale deator in the language of [41]. We
refer to Chapter II for the connection of the opportunity process with convex
duality, which in fact suggests Lemma 5.10. Note that unlike
the previous section,
ฮ“(๐œ‹, ๐œ…)
is positive for all values of
๐‘(๐œ‹, ๐œ…)
from
๐‘.
Our next goal is to link the supermartingale property to local rst order
conditions. Let
๐‘ฆ, ๐‘ฆห‡ โˆˆ C โˆฉC 0 (we will plug in ๐œ‹
ห‡ for ๐‘ฆห‡). The formal directional
๐‘ฆห‡ in the direction of ๐‘ฆ is (๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค โˆ‡๐‘” โ„“ (ห‡
๐‘ฆ ) = ๐บโ„“ (๐‘ฆ, ๐‘ฆห‡), where,
โ„“
derivative of ๐‘” at
by formal dierentiation under the integral sign (cf. (3.2)),
๐บโ„“ (๐‘ฆ, ๐‘ฆห‡) :=
(5.2)
(
๐‘…โ„“
) โˆซ
๐‘…
+ (๐‘ โˆ’ 1)๐‘ ๐‘ฆห‡ +
โ„“โˆ’ (๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค ๐‘๐‘… + ๐‘โ„“โˆ’
(๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค ๐‘ฅโ€ฒ โ„Ž(๐‘ฅ) ๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
โ„๐‘‘ ×โ„
โˆซ
{
}
+
(โ„“โˆ’ + ๐‘ฅโ€ฒ ) (1 + ๐‘ฆห‡โŠค ๐‘ฅ)๐‘โˆ’1 (๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค ๐‘ฅ โˆ’ (๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )).
โ„๐‘‘ ×โ„
We take this expression as the denition of
๐บโ„“ (๐‘ฆ, ๐‘ฆห‡) whenever the last integral
is well dened (the rst one is nite by (4.2)). The dierentiation cannot be
justied in general, but see the subsequent section.
Lemma 5.11. Let ๐‘ฆ โˆˆ C 0 and ๐‘ฆห‡ โˆˆ C 0,โˆ— โˆฉ {๐‘”โ„“ > โˆ’โˆž}. Then ๐บโ„“ (๐‘ฆ, ๐‘ฆห‡) is well
dened with values in (โˆ’โˆž, โˆž] and ๐บโ„“ (โ‹…, ๐‘ฆห‡) is lower semicontinuous on C 0 .
Proof. Writing (๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค ๐‘ฅ = 1 + ๐‘ฆโŠค ๐‘ฅ โˆ’ (1 + ๐‘ฆห‡โŠค ๐‘ฅ), we can express ๐บโ„“ (๐‘ฆ, ๐‘ฆห‡) as
โ„“โˆ’ (๐‘ฆ โˆ’ ๐‘ฆห‡)
โŠค
(
๐‘…
๐‘ +
๐‘๐‘…โ„“
โ„“โˆ’
) โˆซ
+ (๐‘ โˆ’ 1)๐‘ ๐‘ฆห‡ +
๐‘…
(๐‘ฆ โˆ’ ๐‘ฆห‡)โŠค ๐‘ฅโ€ฒ โ„Ž(๐‘ฅ) ๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
โ„๐‘‘ ×โ„
}
1 + ๐‘ฆโŠค๐‘ฅ
โŠค
โˆ’
1
โˆ’
(๐‘ฆ
+
(๐‘
โˆ’
1)ห‡
๐‘ฆ
)
โ„Ž(๐‘ฅ)
๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
(1 + ๐‘ฆห‡โŠค ๐‘ฅ)1โˆ’๐‘
โ„๐‘‘ ×โ„
โˆซ
{
}
โˆ’
(โ„“โˆ’ + ๐‘ฅโ€ฒ ) (1 + ๐‘ฆห‡โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ’ ๐‘ห‡
๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,โ„“ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )).
โˆซ
+
โ„๐‘‘ ×โ„
(โ„“โˆ’ + ๐‘ฅโ€ฒ )
{
III Bellman Equation
54
๐‘ฆ by (4.2). The last integral
๐‘” โ„“ (ห‡
๐‘ฆ ), cf. (3.2), and it is nite if ๐‘” โ„“ (ห‡
๐‘ฆ ) > โˆ’โˆž
The rst integral is nite and continuous in
above occurs in the denition of
and equals
+โˆž
otherwise. Finally, consider the second integral above and
โŠค
1+๐‘ฆ ๐‘ฅ
๐œ“ = ๐œ“(๐‘ฆ, ๐‘ฆห‡, ๐‘ฅ, ๐‘ฅโ€ฒ ). The Taylor expansion (1+ห‡
=
๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
)โŠค
(๐‘โˆ’1) (
โŠค
โŠค
3
1โˆซ + (๐‘ฆ + (๐‘ โˆ’ 1)ห‡
๐‘ฆ ) ๐‘ฅ + 2 2๐‘ฆ + (๐‘ โˆ’ 2)ห‡
๐‘ฆ ๐‘ฅ ๐‘ฅ ๐‘ฆห‡ + ๐‘œ(โˆฃ๐‘ฅโˆฃ ) shows that
๐‘…,โ„“
is well dened and nite. It also shows that given a
{โˆฃ๐‘ฅโˆฃ+โˆฃ๐‘ฅโ€ฒ โˆฃโ‰ค1} ๐œ“ ๐‘‘๐น
โˆซ
๐‘‘
๐‘…,โ„“ is continuous
compact ๐พ โŠ‚ โ„ , there is ๐œ€ > 0 such that
{โˆฃ๐‘ฅโˆฃ+โˆฃ๐‘ฅโ€ฒ โˆฃโ‰ค๐œ€} ๐œ“ ๐‘‘๐น
in ๐‘ฆ โˆˆ ๐พ (and also in ๐‘ฆ
ห‡ โˆˆ ๐พ ). The details are as in Lemma 6.2. Moreover,
0
โ€ฒ
for ๐‘ฆ โˆˆ C we have the lower bound ๐œ“ โ‰ฅ (โ„“โˆ’ +๐‘ฅ ){โˆ’1โˆ’(๐‘ฆ +(๐‘โˆ’1)ห‡
๐‘ฆ )โŠค โ„Ž(๐‘ฅ)},
๐‘…,โ„“
โ€ฒ
which is ๐น
-integrable on {โˆฃ๐‘ฅโˆฃ + โˆฃ๐‘ฅ โˆฃ > ๐œ€} for any ๐œ€ > 0, again by (4.2).
call its integrand
The result now follows by Fatou's lemma.
We can now connect the local rst order conditions for
๐‘”โ„“
and the global
supermartingale property: it turns out that the formal derivative
mines the sign of the drift rate of
ฮ“
๐บโ„“
deter-
(cf. (5.3) below), which leads to the
following proposition. Here and in the sequel, we denote
ห‡ = ๐‘‹(ห‡
๐‘‹
๐œ‹, ๐œ…
ห‡ ).
Proposition 5.12. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solutionโˆซ of the Bellman equation and
ห‡ ๐‘โˆ’1 ๐‘‹(๐œ‹, ๐œ…) + ๐œ…๐‘  โ„“๐‘  ๐‘‹
ห‡ ๐‘ ๐‘โˆ’1 ๐‘‹๐‘  (๐œ‹, ๐œ…) ๐œ‡(๐‘‘๐‘ ) is a
(๐œ‹, ๐œ…) โˆˆ ๐’œ. Then ฮ“(๐œ‹, ๐œ…) := โ„“๐‘‹
supermartingale (local martingale) if and only if ๐บโ„“ (๐œ‹, ๐œ‹ห‡ ) โ‰ค 0 (= 0).
Proof.
¯ = ๐‘… โˆ’ (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘… as in (2.4). We abbreviate ๐œ‹
๐‘…
¯ :=
(๐‘ โˆ’ 1)ห‡
๐œ‹ + ๐œ‹ and similarly ๐œ…
¯ := (๐‘ โˆ’ 1)ห‡
๐œ… + ๐œ…. We defer to Lemma 8.1 a
( ๐‘โˆ’1
)โˆ’1 ( ๐‘โˆ’1
)
ห‡
ห‡
โˆ™ โ„“๐‘‹
calculation showing that ๐‘‹
๐‘‹(๐œ‹, ๐œ…) equals
โˆ’ ๐‘‹โˆ’ (๐œ‹, ๐œ…)
Dene
(
)โŠค
¯ โˆ’ โ„“โˆ’ ๐œ…
โ„“ โˆ’ โ„“0 + โ„“โˆ’ ๐œ‹
¯โˆ™๐‘…
¯ โˆ™ ๐œ‡ + โ„“โˆ’ (๐‘ โˆ’ 1) ๐‘โˆ’2
ห‡ + ๐œ‹ ๐‘๐‘… ๐œ‹
ห‡ โˆ™๐ด+๐œ‹
¯ โŠค ๐‘๐‘…โ„“ โˆ™ ๐ด
2 ๐œ‹
{
}
+๐œ‹
¯ โŠค๐‘ฅโ€ฒ โ„Ž(๐‘ฅ) โˆ— ๐œ‡๐‘…,โ„“ + (โ„“โˆ’ + ๐‘ฅโ€ฒ ) (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ) โˆ’ 1 โˆ’ ๐œ‹
¯ โŠค โ„Ž(๐‘ฅ) โˆ— ๐œ‡๐‘…,โ„“ .
โ„Ž such
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡ ) is a
Here we use a predictable cut-o function
e.g.,
โ„Ž(๐‘ฅ) = ๐‘ฅ1{โˆฃ๐‘ฅโˆฃโ‰ค1}โˆฉ{โˆฃ¯๐œ‹โŠค ๐‘ฅโˆฃโ‰ค1} .
Since
that
๐œ‹
¯ โŠค โ„Ž(๐‘ฅ)
is bounded,
solution, the drift of
โ„“
is
๐ดโ„“ = โˆ’๐‘๐‘ˆ โˆ— (โ„“โˆ’ ) โˆ™ ๐œ‡ โˆ’ ๐‘๐‘” โ„“ (ห‡
๐œ‹ ) โˆ™ ๐ด = (๐‘ โˆ’ 1)โ„“โˆ’ ๐œ…
ห‡ โˆ™ ๐œ‡ โˆ’ ๐‘๐‘” โ„“ (ห‡
๐œ‹ ) โˆ™ ๐ด.
By Remark 2.3,
(โˆ’โˆž, โˆž].
ฮ“ := ฮ“(๐œ‹, ๐œ…)
has a well dened drift rate
๐‘Žฮ“
with values in
From the two formulas above and (2.4) we deduce
ห‡ ๐‘โˆ’1 ๐‘‹(๐œ‹, ๐œ…)โˆ’ ๐บโ„“ (๐œ‹, ๐œ‹
๐‘Žฮ“ = ๐‘‹
ห‡ ).
โˆ’
ห‡ ๐‘โˆ’1 ๐‘‹(๐œ‹, ๐œ…)โˆ’ > 0 by admissibility. If ฮ“ is a supermartingale,
๐‘‹
โˆ’
๐‘Žฮ“ โ‰ค 0, and the converse holds by Lemma 2.4 in view of ฮ“ โ‰ฅ 0.
Here
(5.3)
then
We obtain our second verication theorem from Proposition 5.12 and
Lemma 5.10.
III.5 Verication
55
Theorem 5.13. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solution of the Bellman equation. Assume
that ๐‘ƒ โŠ— ๐ด-a.e., ๐บโ„“ (๐‘ฆ, ๐œ‹ห‡ ) โˆˆ [โˆ’โˆž, 0] for all ๐‘ฆ โˆˆ C โˆฉ C 0,โˆ— . Then
โˆซ
ห‡๐‘ +
ฮ“(ห‡
๐œ‹, ๐œ…
ห‡ ) := โ„“๐‘‹
ห‡ ๐‘ ๐œ‡(๐‘‘๐‘ )
๐œ…
ห‡ ๐‘  โ„“๐‘  ๐‘‹
๐‘ 
is a local martingale. It is a martingale if and only if
is optimal and โ„“ = ๐ฟ is the opportunity process.
If
C
(2.2)
holds and (ห‡๐œ‹, ๐œ…ห‡ )
is not convex, one can imagine situations where the directional
derivative of
๐‘”โ„“
at the maximum is positivei.e., the assumption on
๐บโ„“ (๐‘ฆ, ๐œ‹
ห‡)
is sucient but not necessary. This changes in the subsequent section.
The Convex-Constrained Case
We assume in this section that
C
C โˆฉ C 0 is also convex. Our
โ„“
condition on ๐บ in Theorem 5.13 is
is convex; then
aim is to show that the nonnegativity
automatically satised in this case. We start with an elementary but crucial
observation about dierentiation under the integral sign.
Lemma 5.14. Consider two distinct points ๐‘ฆ0 and ๐‘ฆห‡ in โ„๐‘‘ and let ๐ถ =
{๐œ‚๐‘ฆ0 + (1 โˆ’ ๐œ‚)ห‡
๐‘ฆ : 0 โ‰ค ๐œ‚ โ‰ค 1}. Let ๐œŒ be a function on ฮฃ × ๐ถ , where ฮฃ is
some
Borel
space
with measure ๐œˆ , such that ๐‘ฅ 7โ†’ ๐œŒ(๐‘ฅ, ๐‘ฆ) is ๐œˆ -measurable,
โˆซ +
๐œŒ (๐‘ฅ, โ‹…) ๐œˆ(๐‘‘๐‘ฅ) < โˆž on ๐ถ , and ๐‘ฆ 7โ†’ ๐œŒ(๐‘ฅ, ๐‘ฆ) is concave. In particular, the
directional derivative
(
)
๐œŒ ๐‘ฅ, ๐‘ฆห‡ + ๐œ€(๐‘ฆ โˆ’ ๐‘ฆห‡) โˆ’ ๐œŒ(๐‘ฅ, ๐‘ฆห‡)
๐ท๐‘ฆห‡,๐‘ฆ ๐œŒ(๐‘ฅ, โ‹…) := lim
๐œ€โ†’0+
๐œ€
exists in (โˆ’โˆž, โˆž] for all ๐‘ฆโˆซ โˆˆ ๐ถ . Let ๐›ผ be another concave function on ๐ถ .
Dene ๐›พ(๐‘ฆ) := ๐›ผ(๐‘ฆ) + ๐œŒ(๐‘ฅ, ๐‘ฆ) ๐œˆ(๐‘‘๐‘ฅ) and assume that ๐›พ(๐‘ฆ0 ) > โˆ’โˆž and
that ๐›พ(ห‡๐‘ฆ) = max๐ถ ๐›พ < โˆž. Then for all ๐‘ฆ โˆˆ ๐ถ ,
โˆซ
๐ท๐‘ฆห‡,๐‘ฆ ๐›พ = ๐ท๐‘ฆห‡,๐‘ฆ ๐›ผ +
๐ท๐‘ฆห‡,๐‘ฆ ๐œŒ(๐‘ฅ, โ‹…) ๐œˆ(๐‘‘๐‘ฅ) โˆˆ (โˆ’โˆž, 0]
(5.4)
and in particular ๐ท๐‘ฆห‡,๐‘ฆ ๐œŒ(๐‘ฅ, โ‹…) < โˆž ๐œˆ(๐‘‘๐‘ฅ)-a.e.
Proof.
๐›พ โˆซ> โˆ’โˆž on ๐ถ . Let ๐‘ฃ =
๐‘ฆ)
=
+ ๐œŒ(๐‘ฅ,ห‡๐‘ฆ+๐œ€๐‘ฃ)โˆ’๐œŒ(๐‘ฅ,ห‡
๐œˆ(๐‘‘๐‘ฅ).
(๐‘ฆ โˆ’ ๐‘ฆห‡) and ๐œ€ >
๐œ€
By concavity, these quotients increase monotonically as ๐œ€ โ†“ 0, in particular
their limits exist. The left hand side is nonpositive as ๐‘ฆ
ห‡ is a maximum and
Note that
๐›พ
is concave, hence we also have
๐‘ฆ)
0, then ๐›พ(ห‡๐‘ฆ+๐œ€๐‘ฃ)โˆ’๐›พ(ห‡
๐œ€
๐›ผ(ห‡
๐‘ฆ +๐œ€๐‘ฃ)โˆ’๐›ผ(ห‡
๐‘ฆ)
๐œ€
monotone convergence yields (5.4).
For completeness, let us mention that if
where the left hand side of (5.4) is
โˆ’โˆž
๐›พ(๐‘ฆ0 ) = โˆ’โˆž, there are examples
but the right hand side is nite;
we shall deal with this case separately. We deduce the following version of
Theorem 5.13; as discussed, it involves only the control
(ห‡
๐œ‹, ๐œ…
ห‡ ).
56
III Bellman Equation
Theorem 5.15. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solution of the Bellman
equation and asโˆซ
sume that C is convex. Then ฮ“(ห‡๐œ‹, ๐œ…ห‡ ) := โ„“๐‘‹ห‡ ๐‘ + ๐œ…ห‡ ๐‘  โ„“๐‘  ๐‘‹ห‡ ๐‘ ๐‘ ๐œ‡(๐‘‘๐‘ ) is a local
martingale. It is a martingale if and only if (2.2) holds and (ห‡๐œ‹, ๐œ…ห‡ ) is optimal
and โ„“ = ๐ฟ.
Proof.
๐บโ„“ (๐‘ฆ, ๐œ‹
ห‡ ) โˆˆ [โˆ’โˆž, 0] for
0,โˆ—
โ„“
๐‘ฆ โˆˆ C โˆฉ C . Recall that ๐œ‹
ห‡ is a maximizer for ๐‘” and that ๐บโ„“ was dened
โ„“
by dierentiation under the integral sign. Lemma 5.14 yields ๐บ (๐‘ฆ, ๐œ‹
ห‡) โ‰ค 0
โ„“
โ„“
whenever ๐‘ฆ โˆˆ {๐‘” > โˆ’โˆž}. This ends the proof for ๐‘ โˆˆ (0, 1) as ๐‘” is then
โ„“
nite. If ๐‘ < 0, the denition of ๐‘” and Remark 6.7 show that the set
โˆช
โ„“
{๐‘” > โˆ’โˆž} contains the set ๐œ‚โˆˆ[0,1) ๐œ‚(C โˆฉ C 0 ) which, in turn, is clearly
0,โˆ— . Hence {๐‘” โ„“ > โˆ’โˆž} is dense in C โˆฉ C 0,โˆ— and we obtain
dense in C โˆฉ C
๐บโ„“ (๐‘ฆ, ๐œ‹
ห‡ ) โˆˆ [โˆ’โˆž, 0] for all ๐‘ฆ โˆˆ C โˆฉ C 0,โˆ— using the lower semicontinuity from
To apply Theorem 5.13, we have to check that
Lemma 5.11.
Remark 5.16.
(i)
ฮ“(ห‡
๐œ‹, ๐œ…
ห‡ ) = ๐‘๐‘(ห‡
๐œ‹, ๐œ…
ห‡ ) if ๐‘ is
can be used also for ฮ“(ห‡
๐œ‹, ๐œ…
ห‡ ).
We note that
in (4.4). In particular, Remark 5.9
dened as
(ii) Muhle-Karbe [58] considers certain one-dimensional (unconstrained)
ane models and introduces a sucient optimality condition in the form of
an algebraic inequality (see [58, Theorem 4.20(3)]). This condition can be
seen as a special case of the statement that
๐บ๐ฟ (๐‘ฆ, ๐œ‹
ห‡ ) โˆˆ [โˆ’โˆž, 0]
for
๐‘ฆ โˆˆ C 0,โˆ— ;
in particular, we have shown its necessity.
Of course, all our verication results can be seen as a uniqueness result
for the Bellman equation. As an example, Theorem 5.15 yields:
Corollary 5.17. If C is convex, there is at most one solution of the Bellman
equation in the class of solutions (โ„“, ๐œ‹ห‡ , ๐œ…ห‡ ) such that ฮ“(ห‡๐œ‹, ๐œ…ห‡ ) is of class (D).
Similarly, one can give corollaries for the other results. We close with a
comment concerning convex duality.
Remark 5.18.
(i)
A major insight in [49] was that the dual domain for
utility maximization (here with
C = โ„๐‘‘ )
should be a set of supermartin-
gales rather than (local) martingales when the price process has jumps. A
one-period example for
log-utility
[49, Example 5.1'] showed that the su-
permartingale solving the dual problem can indeed have nonvanishing drift.
In that example it is clear that this arises when the budget constraint becomes binding.
phenomenon.
For general models and
log-utility,
[25] comments on this
The calculations of this section yield an instructive local
picture also for power utility.
๐ฟ and the optimal strat(ห†
๐œ‹, ๐œ…
ห† ) solve the Bellman equation. Assume that C is convex and let
ห† = ๐‘‹(ห†
ห† ๐‘โˆ’1 , which was the solution to the dual
๐‘‹
๐œ‹, ๐œ…
ห† ). Consider ๐‘Œห† = ๐ฟ๐‘‹
ห† โ„ฐ(๐œ‹ โˆ™ ๐‘…) is a supermartingale
problem in Chapter II. We have shown that ๐‘Œ
Under Assumptions 3.1, the opportunity process
egy
III.6 Appendix A: Proof of Lemma 3.8
๐œ‹ โˆˆ ๐’œ, i.e., ๐‘Œห† is a supermartingale deator. Choosing ๐œ‹ = 0,
๐‘Œห† is itself a supermartingale, and by (5.3) its drift rate satises
for every
see that
57
we
ห†
โŠค
ห† ๐‘โˆ’1 ๐บ๐ฟ (0, ๐œ‹
ห† ๐‘โˆ’1 ๐œ‹
๐‘Ž๐‘Œ = ๐‘‹
ห† ) = โˆ’๐‘‹
๐œ‹ ).
โˆ’ ห† โˆ‡๐‘”(ห†
โˆ’
๐‘Œห† is a local martingale if and only if ๐œ‹
ห† โŠค โˆ‡๐‘”(ห†
๐œ‹ ) = 0. One can say
โŠค
that โˆ’ห†
๐œ‹ โˆ‡๐‘”(ห†
๐œ‹ ) < 0 means that the constraints are binding, whereas in an
ห† has nonvanishunconstrained case the gradient of ๐‘” would vanish; i.e., ๐‘Œ
ing drift rate at a given (๐œ”, ๐‘ก) whenever the constraints are binding. Even
๐‘‘
0 in the maximization of
if C = โ„ , we still have the budget constraint C
0
๐‘‘
๐‘” . If in addition ๐‘… is continuous, C = โ„ and we are truly in an unconห† is a local martingale; indeed, in the setting of
strained situation. Then ๐‘Œ
Hence
Corollary 3.12 we calculate
(
1
๐‘Œห† = ๐‘ฆ0 โ„ฐ โˆ’ ๐œ† โˆ™ ๐‘€ +
๐ฟโˆ’
Note how
๐‘ ๐ฟ,
the martingale part of
)
๐‘๐ฟ ,
โˆ™
๐ฟ
๐‘ฆ0 := ๐ฟ0 ๐‘ฅ๐‘โˆ’1
0 .
orthogonal to
๐‘…,
yields the solution
to the dual problem.
(ii) From the proof of Proposition 5.12 we have that the general formula
for the local martingale part of
ห†
ห† ๐‘โˆ’1
๐‘€๐‘Œ = ๐‘‹
โˆ’
๐‘Œห†
is
(
¯
๐‘€ ๐ฟ + ๐ฟโˆ’ (๐‘ โˆ’ 1)ห†
๐œ‹ โˆ™ ๐‘€ ๐‘… + (๐‘ โˆ’ 1)ห†
๐œ‹ โŠค ๐‘ฅโ€ฒ โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘…,๐ฟ โˆ’ ๐œˆ ๐‘…,๐ฟ )
)
{
}
โ€ฒ
โŠค ๐‘โˆ’1
โŠค
๐‘…,๐ฟ
๐‘…,๐ฟ
+ (๐ฟโˆ’ + ๐‘ฅ ) (1 + ๐œ‹
ห† ๐‘ฅ)
โˆ’ 1 โˆ’ (๐‘ โˆ’ 1)ห†
๐œ‹ โ„Ž(๐‘ฅ) โˆ— (๐œ‡
โˆ’๐œˆ ) .
โˆ™
This is relevant in the problem of
๐‘ž -optimal
equivalent martingale measures ;
cf. Goll and Rüschendorf [26] for a general perspective. Let
๐‘ข(๐‘ฅ0 ) < โˆž, ๐ท โ‰ก
1, ๐œ‡ = 0, C =
M of equivalent local martingale
measures for ๐‘† = โ„ฐ(๐‘…) is nonempty. Given ๐‘ž = ๐‘/(๐‘ โˆ’ 1) โˆˆ (โˆ’โˆž, 0) โˆช (0, 1)
โˆ’1 (๐‘‘๐‘„/๐‘‘๐‘ƒ )๐‘ž ] is nite and
conjugate to ๐‘, ๐‘„ โˆˆ M is called ๐‘ž -optimal if ๐ธ[โˆ’๐‘ž
minimal over M . If ๐‘ž < 0, i.e., ๐‘ โˆˆ (0, 1), then ๐‘ข(๐‘ฅ0 ) < โˆž is equivalent to
โˆ’1 (๐‘‘๐‘„/๐‘‘๐‘ƒ )๐‘ž ] < โˆž; moreover,
the existence of some ๐‘„ โˆˆ M such that ๐ธ[โˆ’๐‘ž
โ„๐‘‘ , and assume that the set
Assumptions 3.1 are satised (see Kramkov and Schachermayer [49, 50]).
Using [49, Theorem 2.2(iv)] we conclude that
(a) the
๐‘ž -optimal martingale measure exists if and only if ๐‘Ž๐‘Œ โ‰ก 0 and ๐‘€ ๐‘Œ
ห†
ห†
is a true martingale;
(b) in that case,
1 + ๐‘ฆ0โˆ’1 ๐‘€ ๐‘Œ
ห†
is its
๐‘ƒ -density
process.
This generalizes earlier results of [26] as well as of Grandits [28], Jeanblanc
et al. [35] and Choulli and Stricker [12].
III.6 Appendix A: Proof of Lemma 3.8
This main goal of this appendix is to construct a measurable maximizing
sequence for the random function
๐‘”
(cf. Lemma 3.8). The entire section is
III Bellman Equation
58
under Assumptions 3.1. Before beginning the proof, we discuss the properties
of
๐‘”;
recall that
๐‘”(๐‘ฆ) := ๐ฟโˆ’ ๐‘ฆ
โˆซ
+
โŠค
(
๐‘…
๐‘ +
๐‘๐‘…๐ฟ
๐ฟโˆ’
+
(๐‘โˆ’1) ๐‘…
2 ๐‘ ๐‘ฆ
)
โˆซ
+
๐‘ฅโ€ฒ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
โ„๐‘‘ ×โ„
{
}
(๐ฟโˆ’ + ๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )).
โ„๐‘‘ ×โ„
(6.1)
Lemma 6.1.
๐ฟโˆ’ + ๐‘ฅโ€ฒ
is strictly positive ๐น ๐ฟ (๐‘‘๐‘ฅโ€ฒ )-a.e.
๐ฟ
๐ฟ
โ€ฒ
๐ฟ
Proof.
[ โˆ‘ (๐‘ƒ โŠ— ๐œˆ ){๐ฟโˆ’ + ๐‘ฅ โ‰ค
] 0} = ๐ธ[1{๐ฟโˆ’ +๐‘ฅโ€ฒ โ‰ค0} โˆ— ๐œˆ๐‘‡ ] = ๐ธ[1{๐ฟโˆ’ +๐‘ฅโ€ฒ โ‰ค0} โˆ— ๐œ‡๐‘‡ ] =
๐ธ
๐‘ โ‰ค๐‘‡
1{๐ฟ๐‘  โ‰ค0} 1{ฮ”๐ฟ๐‘  โˆ•=0} = 0
as
๐ฟ>0
by Lemma 2.1.
(๐œ”, ๐‘ก) and let ๐‘™ := ๐ฟ๐‘กโˆ’ (๐œ”). Furthermore, let ๐น be any Lévy measure
๐‘…,๐ฟ
๐‘‘+1
on โ„
which is equivalent to ๐น๐‘ก
(๐œ”) and satises (2.5). Equivalence
0,โˆ—
0
implies that C๐‘ก (๐œ”), C๐‘ก (๐œ”), and N๐‘ก (๐œ”) are the same if dened with respect
๐‘…
to ๐น instead of ๐น . Given ๐œ€ > 0, let
Fix
๐ผ๐œ€๐น (๐‘ฆ)
โˆซ
{
}
(๐‘™ + ๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )),
:=
{โˆฃ๐‘ฅโˆฃ+โˆฃ๐‘ฅโ€ฒ โˆฃโ‰ค๐œ€}
๐น
๐ผ>๐œ€
(๐‘ฆ) :=
โˆซ
{
}
(๐‘™ + ๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )),
{โˆฃ๐‘ฅโˆฃ+โˆฃ๐‘ฅโ€ฒ โˆฃ>๐œ€}
so that
๐น
๐ผ ๐น (๐‘ฆ) := ๐ผ๐œ€๐น (๐‘ฆ) + ๐ผ>๐œ€
(๐‘ฆ)
is the last integral in (6.1) when
๐ผ
๐‘ > 0,
of Lemma 3.4 that
course, when
general
๐น , ๐ผ๐น
๐น ๐‘…,๐ฟ
(๐œ‹)
๐น = ๐น๐‘ก๐‘…,๐ฟ (๐œ”).
We know from the proof
is well dened and nite for any
๐œ‹ โˆˆ ๐’œ๐‘“ ๐ธ
this is essentially due to the assumption (2.2)).
(of
For
has the following properties.
Lemma 6.2. Consider a sequence ๐‘ฆ๐‘› โ†’ ๐‘ฆโˆž in C 0 .
(i) For any ๐‘ฆ โˆˆ C 0 , the integral ๐ผ ๐น (๐‘ฆ) is well dened in โ„ โˆช {sign(๐‘)โˆž}.
(ii) For ๐œ€ โ‰ค (2 sup๐‘› โˆฃ๐‘ฆ๐‘› โˆฃ)โˆ’1 we have ๐ผ๐œ€๐น (๐‘ฆ๐‘› ) โ†’ ๐ผ๐œ€๐น (๐‘ฆโˆž ).
(iii) If ๐‘ โˆˆ (0, 1), ๐ผ ๐น is l.s.c., that is, lim inf ๐‘› ๐ผ ๐น (๐‘ฆ๐‘› ) โ‰ฅ ๐ผ ๐น (๐‘ฆโˆž ).
(iv) If ๐‘ < 0, ๐ผ ๐น is u.s.c., that is, lim sup๐‘› ๐ผ ๐น (๐‘ฆ๐‘› ) โ‰ค ๐ผ ๐น (๐‘ฆโˆž ). Moreover,
๐‘ฆ โˆˆ C 0 โˆ– C 0,โˆ— implies ๐ผ ๐น (๐‘ฆ) = โˆ’โˆž.
Proof.
The rst item follows from the subsequent considerations.
(ii) We may assume that โ„Ž is the identity on {โˆฃ๐‘ฅโˆฃ โ‰ค ๐œ€}; then on this set
๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) =: ๐œ“(๐‘ง)โˆฃ๐‘ง=๐‘ฆโŠค ๐‘ฅ , where the function ๐œ“ is smooth
on {โˆฃ๐‘งโˆฃ โ‰ค 1/2} โŠ† โ„ satisfying
๐œ“(๐‘ง) = ๐‘โˆ’1 (1 + ๐‘ง)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ง =
๐‘โˆ’1 2
2 ๐‘ง
+ ๐‘œ(โˆฃ๐‘งโˆฃ3 )
III.6 Appendix A: Proof of Lemma 3.8
59
หœ
๐œ“(๐‘ง) = ๐‘ง 2 ๐œ“(๐‘ง)
with a function
๐œ“หœ that is continuous and in particular bounded on {โˆฃ๐‘งโˆฃ โ‰ค 1/2}.
โ€ฒ 2
2
As a Lévy measure, ๐น integrates (โˆฃ๐‘ฅ โˆฃ + โˆฃ๐‘ฅโˆฃ ) on compacts; in particular,
โ€ฒ
2
โ€ฒ
๐บ(๐‘‘(๐‘ฅ, ๐‘ฅ )) := โˆฃ๐‘ฅโˆฃ ๐น (๐‘‘(๐‘ฅ, ๐‘ฅ )) denes a nite measure on {โˆฃ๐‘ฅโˆฃ + โˆฃ๐‘ฅโ€ฒ โˆฃ โ‰ค ๐œ€}.
๐น
โˆ’1 , and dominated converHence ๐ผ๐œ€ (๐‘ฆ) is well dened and nite for โˆฃ๐‘ฆโˆฃ โ‰ค (2๐œ€)
โˆซ
๐น
โ€ฒ
โŠค
โ€ฒ
หœ
gence shows that ๐ผ๐œ€ (๐‘ฆ) =
{โˆฃ๐‘ฅโˆฃ+โˆฃ๐‘ฅโ€ฒ โˆฃโ‰ค๐œ€} (๐‘™+๐‘ฅ )๐œ“(๐‘ฆ ๐‘ฅ) ๐บ(๐‘‘(๐‘ฅ, ๐‘ฅ )) is continuous
โˆ’1 }.
in ๐‘ฆ on {โˆฃ๐‘ฆโˆฃ โ‰ค (2๐œ€)
๐น is bounded
(iii) For โˆฃ๐‘ฆโˆฃ bounded by a constant ๐ถ , the integrand in ๐ผ
โ€ฒ
โ€ฒ
โ€ฒ
from below by ๐ถ + โˆฃ๐‘ฅ โˆฃ for some constant ๐ถ depending on ๐‘ฆ only through
๐ถ . We choose ๐œ€ as before. As ๐ถ โ€ฒ + โˆฃ๐‘ฅโ€ฒ โˆฃ is ๐น -integrable on {โˆฃ๐‘ฅโˆฃ + โˆฃ๐‘ฅโ€ฒ โˆฃ > ๐œ€}
๐น
by (2.5), ๐ผ (๐‘ฆ) is well dened in โ„ โˆช {โˆž} and l.s.c. by Fatou's lemma.
because
1+๐‘ง
is bounded away from
0.
Thus
(iv) The rst part follows as in (iii), now the integrand is bounded from
๐ถ โ€ฒ + โˆฃ๐‘ฅโ€ฒ โˆฃ.
โˆ’โˆž on a set
above by
equals
๐‘ฆ โˆˆ C 0 โˆ– C 0,โˆ— , Lemma
of positive ๐น -measure.
If
6.1 shows that the integrand
Lemma 6.3. The function ๐‘” is concave. If C is convex, ๐‘” has at most one
maximum on C โˆฉ C 0 , modulo N .
Proof.
๐‘” need not
๐‘…๐‘ก = ๐‘ก(1, . . . , 1)โŠค was
We rst remark that the assertion is not trivial because
be strictly concave on
N
โŠฅ , for example, the process
not excluded.
๐‘”(๐‘ฆ) = ๐ป๐‘ฆ + ๐ฝ(๐‘ฆ), where ๐ป๐‘ฆ = ๐ฟโˆ’ ๐‘ฆ โŠค ๐‘๐‘… +
โŠค ๐‘…
๐น ๐‘…,๐ฟ (๐‘ฆ) is
๐‘ฆ โŠค ๐‘๐‘…๐ฟ + ๐‘ฅโ€ฒ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ is linear and ๐ฝ(๐‘ฆ) = (๐‘โˆ’1)
2 ๐ฟโˆ’ ๐‘ฆ ๐‘ ๐‘ฆ + ๐ผ
concave. We may assume that โ„Ž(๐‘ฅ) = ๐‘ฅ1{โˆฃ๐‘ฅโˆฃโ‰ค1} .
0 be such that ๐‘”(๐‘ฆ ) = ๐‘”(๐‘ฆ ) = sup ๐‘” =: ๐‘” โˆ— < โˆž,
Let ๐‘ฆ1 , ๐‘ฆ2 โˆˆ C โˆฉ C
1
2
โˆ—
our aim is to show ๐‘ฆ1 โˆ’ ๐‘ฆ2 โˆˆ N . By concavity, ๐‘” = ๐‘”((๐‘ฆ1 + ๐‘ฆ2 )/2)) =
[๐‘”(๐‘ฆ1 ) + ๐‘”(๐‘ฆ2 )]/2, which implies ๐ฝ((๐‘ฆ1 + ๐‘ฆ2 )/2)) = [๐ฝ(๐‘ฆ1 ) + ๐ฝ(๐‘ฆ2 )]/2 due to
the linearity of ๐ป . Using the denition of ๐ฝ , this shows that ๐ฝ is constant
on the line segment connecting ๐‘ฆ1 and ๐‘ฆ2 . A rst consequence is that ๐‘ฆ1 โˆ’ ๐‘ฆ2
}
โŠค ๐‘…
๐‘…
โŠค
lies in the set {๐‘ฆ : ๐‘ฆ ๐‘ = 0, ๐น {๐‘ฅ : ๐‘ฆ ๐‘ฅ โˆ•= 0} = 0 and a second is that
๐ป๐‘ฆ1 = ๐ป๐‘ฆ2 . It remains to show (๐‘ฆ1 โˆ’ ๐‘ฆ2 )โŠค ๐‘๐‘… = 0 to have ๐‘ฆ1 โˆ’ ๐‘ฆ2 โˆˆ N .
๐‘…
โŠค
๐‘…,๐ฟ {๐‘ฅ : ๐‘ฆ โŠค โ„Ž(๐‘ฅ) โˆ•= 0} = 0.
Note that ๐น {๐‘ฅ : ๐‘ฆ ๐‘ฅ โˆ•= 0} = 0 implies ๐น
โŠค
๐‘…
โŠค
๐‘…๐ฟ
Moreover, ๐‘ฆ ๐‘
= 0 implies ๐‘ฆ ๐‘
= 0 due to the absolute continuity
โŸจ๐‘…๐‘,๐‘– , ๐ฟ๐‘ โŸฉ โ‰ช โŸจ๐‘…๐‘,๐‘– โŸฉ which follows from the Kunita-Watanabe
inequality.
โˆซ โ€ฒ
โŠค
Therefore, the rst consequence above implies
๐‘ฅ (๐‘ฆ1 โˆ’ ๐‘ฆ2 ) โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ = 0
โŠค
๐‘…๐ฟ
and (๐‘ฆ1 โˆ’ ๐‘ฆ2 ) ๐‘
= 0, and now the second consequence and the denition
โŠค ๐‘…
โŠค ๐‘… = 0 as
of ๐ป yield 0 = ๐ป(๐‘ฆ1 โˆ’ ๐‘ฆ2 ) = ๐ฟโˆ’ (๐‘ฆ1 โˆ’ ๐‘ฆ2 ) ๐‘ . Thus (๐‘ฆ1 โˆ’ ๐‘ฆ2 ) ๐‘
๐ฟโˆ’ > 0 and this ends the proof.
Note that
๐‘”
is of the form
โˆซ
We can now move toward the main goal of this section. Clearly we need
some variant of the Measurable Maximum Theorem (see, e.g., [1, 18.19],
[41, Theorem 9.5], [64, 2K]). We state a version that is tailored to our needs
and has a simple proof; the technique is used also in Proposition 4.3.
60
III Bellman Equation
Lemma 6.4. Let๐‘‘D be a predictable set-valued process with nonempty compact values in 2โ„ . Let ๐‘“ (๐‘ฆ) = ๐‘“ (๐œ”, ๐‘ก, ๐‘ฆ) be a proper function on D with
values in โ„ โˆช {โˆ’โˆž} such that
(i) ๐‘“ (๐œ‘) is predictable whenever ๐œ‘ is a D -valued predictable process,
(ii) ๐‘ฆ 7โ†’ ๐‘“ (๐‘ฆ) is upper semicontinuous on D for xed (๐œ”, ๐‘ก).
Then there exists a D -valued predictable process ๐œ‹ such that ๐‘“ (๐œ‹) = maxD ๐‘“ .
Proof.
D : there exist
D -valued predictable processes (๐œ‘๐‘› )๐‘›โ‰ฅ1 such that {๐œ‘๐‘› : ๐‘› โ‰ฅ 1} = D for each
(๐œ”, ๐‘ก). By (i), ๐‘“ โˆ— := max๐‘› ๐‘“ (๐œ‘๐‘› ) is predictable, and ๐‘“ โˆ— = maxD ๐‘“ by (ii).
โˆ—
๐‘›
Fix ๐‘˜ โ‰ฅ 1 and let ฮ›๐‘› := {๐‘“ โˆ’ ๐‘“ (๐œ‘๐‘› ) โ‰ค 1/๐‘˜}, ฮ› := ฮ›๐‘› โˆ– (ฮ›1 โˆช โ‹… โ‹… โ‹… โˆช ฮ›๐‘›โˆ’1 ).
โˆ‘
๐‘˜
โˆ—
๐‘˜
๐‘˜
Dene ๐œ‹ :=
๐‘› ๐œ‘๐‘› 1ฮ›๐‘› , then ๐‘“ โˆ’ ๐‘“ (๐œ‹ ) โ‰ค 1/๐‘˜ and ๐œ‹ โˆˆ D .
๐‘˜
It remains to select a cluster point: By compactness, (๐œ‹ )๐‘˜โ‰ฅ1 is bounded
for each (๐œ”, ๐‘ก), so there is a convergent subsequence along random indices
๐œ๐‘˜ . More precisely, there exists a strictly increasing sequence of integerโˆ—
valued predictable processes ๐œ๐‘˜ = {๐œ๐‘˜ (๐œ”, ๐‘ก)} and a predictable process ๐œ‹
๐œ๐‘˜ (๐œ”,๐‘ก)
such that lim๐‘˜ ๐œ‹๐‘ก
(๐œ”) = ๐œ‹๐‘กโˆ— (๐œ”) for all (๐œ”, ๐‘ก). See, e.g., the proof of Föllmer
โˆ—
โˆ—
and Schied [22, Lemma 1.63]. We have ๐‘“ = ๐‘“ (๐œ‹ ) by (ii).
We start with the Castaing representation [64, 1B] of
Our random function
๐‘”
satises property (i) of Lemma 6.4 because the
characteristics are predictable (recall the denition [34, II.1.6]). We also note
that the intersection of closed predictable processes is predictable [64, 1M].
๐‘<0
and denote
๐‘” ; we start
๐ต๐‘Ÿ (โ„๐‘‘ ) = {๐‘ฅ โˆˆ โ„๐‘‘ : โˆฃ๐‘ฅโˆฃ โ‰ค ๐‘Ÿ}.
Proof of Lemma 3.8 for ๐‘ < 0.
In this case
๐‘” is u.s.c. on C โˆฉC 0 (Lemma 6.2).
The sign of
๐‘
is important as it switches the semicontinuity of
with the immediate case
C0
D(๐‘›) := C โˆฉ
โˆฉ ๐ต๐‘›
Lemma 6.4 yields a predictable process
โˆˆ arg maxD(๐‘›) ๐‘” for each ๐‘› โ‰ฅ 1, and clearly lim๐‘› ๐‘”(๐œ‹ ๐‘› ) = supC โˆฉC 0 ๐‘” . As
๐‘”(๐œ‹ ๐‘› ) โ‰ฅ ๐‘”(0) = 0, we have ๐œ‹ ๐‘› โˆˆ C 0,โˆ— by Lemma 6.2.
Let
(โ„๐‘‘ ).
๐œ‹๐‘›
III.6.1 Measurable Maximizing Sequence for ๐‘ โˆˆ (0, 1)
Fix
๐‘ โˆˆ (0, 1).
Since the continuity properties of
๐‘” are not clear, we will use an
approximating sequence of continuous functions. (See also Appendix III.7,
where an alternative approach is discussed and the continuity is claried
under an additional assumption on
C .)
We will approximate
๐‘”
using Lévy
measures with enhanced integrability, a method suggested by [41] in a similar
problem. This preserves monotonicity properties that will be useful to pass
to the limit.
๐‘… is locally bounded, or more generally if ๐น ๐‘…,๐ฟ
condition. We start with xed (๐œ”, ๐‘ก).
All this is not necessary if
satises the following
III.6 Appendix A: Proof of Lemma 3.8
Denition 6.5.
Let
๐น
โˆซ
(ii) The
โ„๐‘‘+1 which
๐‘-suitable if
be a Lévy measure on
๐น ๐‘…,๐ฟ and satises (2.5). (i) We say that
๐น
is
61
is equivalent to
(1 + โˆฃ๐‘ฅโ€ฒ โˆฃ)(1 + โˆฃ๐‘ฅโˆฃ)๐‘ 1{โˆฃ๐‘ฅโˆฃ>1} ๐น (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )) < โˆž.
๐‘-suitable
approximating sequence for ๐น
Lévy measures dened by
๐‘‘๐น๐‘› /๐‘‘๐น = ๐‘“๐‘› ,
is the sequence
(๐น๐‘› )๐‘›โ‰ฅ1
of
where
๐‘“๐‘› (๐‘ฅ) = 1{โˆฃ๐‘ฅโˆฃโ‰ค1} + ๐‘’โˆ’โˆฃ๐‘ฅโˆฃ/๐‘› 1{โˆฃ๐‘ฅโˆฃ>1} .
It is easy to see that each
addition being
quence
๐‘“๐‘›
๐‘-suitable
๐น๐‘›
in (ii) shares the properties of
because
(1 + โˆฃ๐‘ฅโˆฃ)๐‘ ๐‘’โˆ’โˆฃ๐‘ฅโˆฃ/๐‘›
is bounded.
is increasing, monotone convergence shows that
for any measurable function
๐‘‰ โ‰ฅ0
on
which is dened as in (6.1) but with
โ„๐‘‘+1 .
๐น ๐‘…,๐ฟ
โˆซ
We denote by
replaced by
๐น,
while in
As the se-
โˆซ
๐‘‰ ๐‘‘๐น๐‘› โ†‘ ๐‘‰ ๐‘‘๐น
๐‘” ๐น the function
๐น.
Lemma 6.6. If ๐น is ๐‘-suitable, ๐‘”๐น is real-valued and continuous on C 0 .
Proof.
C 0 . The only term in (6.1) for which continuity is not
๐น = ๐ผ ๐น +๐ผ ๐น , where we choose ๐œ€ as in Lemma 6.2. We
evident, is the integral ๐ผ
๐œ€
>๐œ€
๐น
๐น
have ๐ผ๐œ€ (๐‘ฆ๐‘› ) โ†’ ๐ผ๐œ€ (๐‘ฆ) by that lemma. When ๐น is ๐‘-suitable, the continuity
๐น
of ๐ผ>๐œ€ follows from the dominated convergence theorem.
Pick
๐‘ฆ๐‘› โ†’ ๐‘ฆ
Remark 6.7.
in
Dene the set
โˆช
(C โˆฉ C 0 )โ‹„ :=
๐œ‚(C โˆฉ C 0 ).
๐œ‚โˆˆ[0,1)
1 + ๐‘ฆ โŠค ๐‘ฅ is ๐น ๐‘… (๐‘‘๐‘ฅ)-essentially bounded
away from zero. Indeed, ๐‘ฆ = ๐œ‚๐‘ฆ0 with ๐œ‚ โˆˆ [0, 1) and ๐น ๐‘… {๐‘ฆ0โŠค ๐‘ฅ โ‰ฅ โˆ’1} = 0,
โŠค
๐‘…
0 โ‹„
0,โˆ— . If C is
hence 1 + ๐‘ฆ ๐‘ฅ โ‰ฅ 1 โˆ’ ๐œ‚ , ๐น -a.e. In particular, (C โˆฉ C ) โŠ† C
0 โ‹„
star-shaped with respect to the origin, we also have (C โˆฉ C ) โŠ† C .
Its elements
๐‘ฆ
have the property that
We introduce the compact-valued process
D(๐‘Ÿ) := C โˆฉ C 0 โˆฉ ๐ต๐‘Ÿ (โ„๐‘‘ ).
Lemma 6.8. Let ๐น be ๐‘-suitable. Under (C3), arg maxD(๐‘Ÿ) ๐‘”๐น โŠ† C 0,โˆ— .
More generally, this holds whenever ๐น is a Lévy measure equivalent to
๐น ๐‘…,๐ฟ satisfying (2.5) and ๐‘” ๐น is nite-valued.
Proof.
Assume that
as in the denition
derivative
๐ท๐‘ฆห‡,๐‘ฆ0 ๐‘”
๐‘ฆห‡ โˆˆ C 0 โˆ– C 0,โˆ— is a maximum of ๐‘” ๐น . Let ๐œ‚ โˆˆ (๐œ‚, 1) be
of (C3) and ๐‘ฆ0 := ๐œ‚ ๐‘ฆ
ห‡. By Lemma 5.14, the directional
can be calculated by dierentiating under the integral sign.
For the integrand of
๐ผ๐น
we have
{
}
{
}
๐ท๐‘ฆห‡,๐‘ฆ0 ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) = (1 โˆ’ ๐œ‚) (1 + ๐‘ฆห‡โŠค ๐‘ฅ)๐‘โˆ’1 ๐‘ฆห‡โŠค ๐‘ฅ โˆ’ ๐‘ฆห‡โŠค โ„Ž(๐‘ฅ) .
But this is innite on a set of positive measure as
๐น {ห‡
๐‘ฆโŠค๐‘ฅ
= โˆ’1} > 0,
๐‘ฆห‡ โˆˆ C 0 โˆ– C 0,โˆ—
means that
contradicting the last assertion of Lemma 5.14.
62
III Bellman Equation
Let
๐น
be a Lévy measure on
โ„๐‘‘+1
which is equivalent to
๐น ๐‘…,๐ฟ
and sat-
ises (2.5). The crucial step is
Lemma 6.9. Let (๐น๐‘› ) be the ๐‘-suitable approximating sequence for ๐น and
x ๐‘Ÿ > 0. For each ๐‘›, arg maxD(๐‘Ÿ) ๐‘”๐น๐‘› โˆ•= โˆ…, and for any ๐‘ฆ๐‘›โˆ— โˆˆ arg maxD(๐‘Ÿ) ๐‘”๐น๐‘›
it holds that lim sup๐‘› ๐‘”๐น (๐‘ฆ๐‘›โˆ— ) = supD(๐‘Ÿ) ๐‘”๐น .
Proof.
We rst show that
๐ผ ๐น๐‘› (๐‘ฆ) โ†’ ๐ผ ๐น (๐‘ฆ)
for any
๐‘ฆ โˆˆ C 0.
(6.2)
{
}
โˆซ
๐ผ ๐น๐‘› (๐‘ฆ) = (๐‘™+๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1+๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’๐‘โˆ’1 โˆ’๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐‘“๐‘› (๐‘ฅ) ๐น (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )),
where ๐‘“๐‘› is nonnegative and increasing in ๐‘›. As ๐‘“๐‘› = 1 in a neighbor๐น๐‘›
hood of the origin, we need to consider only ๐ผ>๐œ€ (for ๐œ€ = 1, say). Its
integrand is bounded below, simultaneously for all ๐‘›, by a negative conโ€ฒ
stant times (1 + โˆฃ๐‘ฅ โˆฃ), which is ๐น -integrable on the relevant domain. As
{
โ€ฒ
(๐‘“๐‘› ) is increasing, we can apply monotone
convergence on the set (๐‘ฅ, ๐‘ฅ ) :
}
โˆ’1
โŠค
๐‘
โˆ’1
โŠค
๐‘ (1 + ๐‘ฆ ๐‘ฅ) โˆ’ ๐‘ โˆ’ ๐‘ฆ โ„Ž(๐‘ฅ) โ‰ฅ 0 and dominated convergence on the
Recall that
complement to deduce (6.2).
๐‘ฆ๐‘›โˆ— โˆˆ arg maxD(๐‘Ÿ) ๐‘” ๐น๐‘› is clear by compactness of D(๐‘Ÿ) and
๐‘” ๐น๐‘› (Lemma 6.6). Let ๐‘ฆ โˆˆ D(๐‘Ÿ) be arbitrary. By denition of
Existence of
continuity of
๐‘ฆ๐‘›โˆ—
and (6.2),
lim sup ๐‘” ๐น๐‘› (๐‘ฆ๐‘›โˆ— ) โ‰ฅ lim sup ๐‘” ๐น๐‘› (๐‘ฆ) = ๐‘” ๐น (๐‘ฆ).
๐‘›
๐‘›
lim sup๐‘› ๐‘” ๐น (๐‘ฆ๐‘›โˆ— ) โ‰ฅ lim sup๐‘› ๐‘” ๐น๐‘› (๐‘ฆ๐‘›โˆ— ). We can split the integral
๐น
๐ผ ๐‘› (๐‘ฆ) into a sum of three terms: The integral over {โˆฃ๐‘ฅโˆฃ โ‰ค 1} is the same as
๐น
for ๐ผ , since ๐‘“๐‘› = 1 on this set. We can assume that the cut-o โ„Ž vanishes
We show
outside
{โˆฃ๐‘ฅโˆฃ โ‰ค 1}.
The second term is then
โˆซ
(๐‘™ + ๐‘ฅโ€ฒ )๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ ๐‘“๐‘› ๐‘‘๐น,
{โˆฃ๐‘ฅโˆฃ>1}
here the integrand is nonnegative and hence increasing in
the third term is
โˆซ
๐‘›,
for all
๐‘ฆ;
and
(๐‘™ + ๐‘ฅโ€ฒ )(โˆ’๐‘โˆ’1 )๐‘“๐‘› ๐‘‘๐น,
{โˆฃ๐‘ฅโˆฃ>1}
which is decreasing in
๐‘›
but converges to
โˆซ
{โˆฃ๐‘ฅโˆฃ>1} (๐‘™
+ ๐‘ฅโ€ฒ )(โˆ’๐‘โˆ’1 ) ๐‘‘๐น .
Thus
๐‘” ๐น (๐‘ฆ๐‘›โˆ— ) โ‰ฅ ๐‘” ๐น๐‘› (๐‘ฆ๐‘›โˆ— ) โˆ’ ๐œ€๐‘›
โˆซ
๐œ€๐‘› := {โˆฃ๐‘ฅโˆฃ>1} (๐‘™ + ๐‘ฅโ€ฒ )(โˆ’๐‘โˆ’1 )(๐‘“๐‘› โˆ’ 1) ๐‘‘๐น โ†“ 0. Together,
supD(๐‘Ÿ) ๐‘” ๐น โ‰ฅ lim sup๐‘› ๐‘” ๐น (๐‘ฆ๐‘›โˆ— ) โ‰ฅ lim sup๐‘› ๐‘” ๐น๐‘› (๐‘ฆ๐‘›โˆ— ) โ‰ฅ supD(๐‘Ÿ) ๐‘” ๐น .
with the sequence
conclude
we
III.7 Appendix B: Parametrization by Representative Portfolios
63
Proof of Lemma 3.8 for ๐‘ โˆˆ (0, 1).
Fix ๐‘Ÿ > 0. By Lemma 6.4 we can nd
arg maxD(๐‘Ÿ) ๐‘” ๐น๐‘› , i.e., ๐œ‹๐‘ก๐‘›,๐‘Ÿ (๐œ”) plays the role
โˆ—
๐‘› := ๐œ‹ ๐‘›,๐‘› and noting D(๐‘›) โ†‘ C โˆฉ C 0 ,
of ๐‘ฆ๐‘› in Lemma 6.9.
Taking ๐œ‹
๐‘› are C โˆฉ C 0 -valued predictable processes such
Lemma 6.9 shows that ๐œ‹
๐‘›
๐‘›
that lim sup๐‘› ๐‘”(๐œ‹ ) = supC โˆฉC 0 ๐‘” ๐‘ƒ โŠ— ๐ด-a.e. Lemma 6.8 shows that ๐œ‹ takes
0,โˆ—
values in C
.
measurable selectors
๐œ‹ ๐‘›,๐‘Ÿ
for
III.7 Appendix B: Parametrization by Representative Portfolios
This appendix introduces an equivalent transformation of the model
(๐‘…, C )
with specic properties (Theorem 7.3); the main idea is to substitute the
given assets by wealth processes that represent the investment opportunities
of the model. While the result is of independent interest, the main conclusion
in our context is that the approximation technique from Appendix III.6.1 for
๐‘ โˆˆ (0, 1) can be avoided, at least under slightly stronger assumptions on C : If the utility maximization problem is nite, the corresponding
Lévy measure in the transformed model is ๐‘-suitable (cf. Denition 6.5) and
hence the corresponding function ๐‘” is continuous. This is not only an althe case
ternative argument to prove Lemma 3.8. In applications, continuity can be
useful to construct a maximizer for
if one does not know
a priori
๐‘”
(rather than a maximizing sequence)
that there exists an optimal strategy. A static
version of our construction is carried out in Chapter IV.
In this appendix we use the following assumptions on the set-valued
process
(C1)
(C2)
(C4)
C
of constraints:
C
is predictable.
C
is closed.
C
is star-shaped with respect to the origin:
๐œ‚C โŠ† C
for all
๐œ‚ โˆˆ [0, 1].
Since we already obtained a proof of Lemma 3.8, we do not strive for
minimal conditions here.
Clearly (C4) implies condition (C3) from Sec-
tion III.2.4, but its main implication is that we can select a bounded (hence
๐‘…-integrable)
process in the subsequent lemma. The following result is the
๐‘— th representative portfolio, a portfolio with the property
the ๐‘— th asset whenever this is feasible.
construction of the
that it invests in
Lemma 7.1. Fix 1 โ‰ค ๐‘— โ‰ค ๐‘‘ and let ๐ป ๐‘— = {๐‘ฅ โˆˆ โ„๐‘‘ : ๐‘ฅ๐‘— โˆ•= 0}. There exists
a bounded predictable C โˆฉ C 0,โˆ— -valued process ๐œ™ satisfying
{
}
{๐œ™๐‘— = 0} = C โˆฉ C 0,โˆ— โˆฉ ๐ป ๐‘— = โˆ… .
Proof.
๐‘—
๐ต{1 = ๐ต1 (โ„๐‘‘ ) be the} closed
{ unit ball0,โˆ—and ๐ป :=}๐ป . Condition
0,โˆ—
(C4) implies C โˆฉ C
โˆฉ ๐ป = โˆ… = C โˆฉ ๐ต1 โˆฉ C โˆฉ ๐ป = โˆ… , hence we may
๐‘‘
๐‘—
โˆ’1 }
substitute C by C โˆฉ ๐ต1 . Dene the closed sets ๐ป๐‘˜ = {๐‘ฅ โˆˆ โ„ : โˆฃ๐‘ฅ โˆฃ โ‰ฅ ๐‘˜
Let
III Bellman Equation
64
for
๐‘˜ โ‰ฅ 1,
then
โˆช
๐‘˜
๐ป๐‘˜ = ๐ป .
Moreover, let
D๐‘˜ = C โˆฉ C 0 โˆฉ ๐ป๐‘˜ .
This is a
compact-valued predictable process, so there exists a predictable process
๐‘—
such that ๐œ™๐‘˜ โˆˆ D๐‘˜ (hence ๐œ™๐‘˜ โˆ•=
๐‘˜
the complement. Dene ฮ› :=
๐œ™๐‘˜
0) on the set ฮ›๐‘˜ := {D๐‘˜ โˆ•= โˆ…} and โˆ‘
๐œ™๐‘˜ = 0 on
โ€ฒ :=
ฮ›
โˆ–
(ฮ›
โˆช
โ‹…
โ‹…
โ‹…
โˆช
ฮ›
)
and
๐œ™
1
๐‘˜ ๐œ™}๐‘˜ 1ฮ›๐‘˜ .
{ ๐‘˜
} ๐‘˜โˆ’1
{
โ€ฒ
โ€ฒ๐‘—
0
0,โˆ—
Then โˆฃ๐œ™ โˆฃ โ‰ค 1 and {๐œ™ = 0} = C โˆฉ C โˆฉ ๐ป = โˆ… = C โˆฉ C
โˆฉ ๐ป = โˆ… ; the
second equality uses (C4) and Remark 6.7. These two facts also show that
๐œ™ := 12 ๐œ™โ€ฒ
has the same property while in addition being
Remark 7.2.
diameter of
C
C โˆฉ C 0,โˆ— -valued.
The previous proof also applies if instead of (C4), e.g., the
is uniformly bounded and
C 0 = C 0,โˆ— .
ฮฆ is a ๐‘‘×๐‘‘-matrix with columns ๐œ™1 , . . . , ๐œ™๐‘‘ โˆˆ ๐ฟ(๐‘…), the matrix stochasหœ = ฮฆ โˆ™ ๐‘… is the โ„๐‘‘ -valued process given by ๐‘…
หœ๐‘— = ๐œ™๐‘— โˆ™ ๐‘…. If
๐‘…
๐œ“ โˆˆ ๐ฟ(ฮฆ โˆ™ ๐‘…) is โ„๐‘‘ -valued, then ฮฆ๐œ“ โˆˆ ๐ฟ(๐‘…) and
If
tic integral
๐œ“ โˆ™ (ฮฆ โˆ™ ๐‘…) = (ฮฆ๐œ“) โˆ™ ๐‘….
If
D
(7.1)
is a set-valued process which is predictable, closed and contains the ori-
gin, then the preimage
ฮฆโˆ’1 D
shares these properties (cf. [64, 1Q]). Convexity
and star-shape are also preserved.
We obtain the following model if we sequentially replace the given assets
by representative portfolios; here
1โ‰ค๐‘—โ‰ค๐‘‘
๐‘–
(i.e., ๐‘’๐‘—
๐‘’๐‘—
denotes the
๐‘— th
unit vector in
โ„๐‘‘
for
= ๐›ฟ๐‘–๐‘— ).
Theorem 7.3. There exists a predictable โ„๐‘‘×๐‘‘ -valued uniformly bounded
process ฮฆ such that the nancial market model with returns
หœ := ฮฆ โˆ™ ๐‘…
๐‘…
and constraints Cหœ := ฮฆโˆ’1 C has the following properties: for all 1 โ‰ค ๐‘— โ‰ค ๐‘‘,
(i) ฮ”๐‘…หœ๐‘— > โˆ’1 (positive prices),
(ii) ๐‘’๐‘— โˆˆ Cหœ โˆฉ Cหœ0,โˆ— , where Cหœ0,โˆ— = ฮฆโˆ’1 C 0,โˆ— (entire wealth can be invested in
each asset),
หœ Cหœ) admits the same wealth processes as (๐‘…, C ).
(iii) the model (๐‘…,
Proof.
We treat the components one by one. Let
๐‘— =1
and let
1
be as in Lemma 7.1. We replace the rst asset ๐‘… by the process
equivalently, we replace
๐‘…
ฮฆ โˆ™ ๐‘…, where ฮฆ = ฮฆ(1)
โŽ› 1
โŽž
๐œ™
โŽœ ๐œ™2 1
โŽŸ
โŽœ
โŽŸ
ฮฆ=โŽœ .
โŽŸ.
..
โŽ ..
โŽ 
.
๐‘‘
๐œ™
1
by
๐‘‘ × ๐‘‘-matrix
ฮฆโˆ’1 C 0 and we replace C by ฮฆโˆ’1 C . Note
๐‘’1 โˆˆ ฮฆโˆ’1 (C โˆฉ C 0,โˆ— ) because ฮฆ๐‘’1 = ๐œ™ โˆˆ C โˆฉ C 0,โˆ— by construction.
The new natural constraints are
that
is the
๐œ™ = ๐œ™(1)
๐œ™ โˆ™ ๐‘…, or
III.7 Appendix B: Parametrization by Representative Portfolios
65
C โˆฉ C 0,โˆ— -valued process ๐œ‹ โˆˆ ๐ฟ(๐‘…) there exists
ฮฆ๐œ“ = ๐œ‹ . In view of (7.1), this will imply that the
We show that for every
๐œ“
predictable such that
new model admits the same wealth processes as the old one.
On the set
{๐œ™1 โˆ•= 0} = {ฮฆ is invertible} we take ๐œ“ = ฮฆโˆ’1 ๐œ‹ and on the complement we
1
๐‘—
๐‘—
choose ๐œ“ โ‰ก 0 and ๐œ“ = ๐œ‹ for ๐‘— โ‰ฅ 2; this is the same as inverting ฮฆ on its
1
1
image. Note that {๐œ™ = 0} โŠ† {๐œ‹ = 0} by the choice of ๐œ™.
We proceed with the second component of the new model in the same
We obtain matrices ฮฆ(๐‘—) for
ฮฆฬ‚ = ฮฆ(1) โ‹… โ‹… โ‹… ฮฆ(๐‘‘). Then ฮฆฬ‚ has the required properties.
โˆ’1 (C โˆฉC 0,โˆ— ).
Indeed, the construction and ฮฆ(๐‘–)๐‘’๐‘— = ๐‘’๐‘— for ๐‘– โˆ•= ๐‘— imply ๐‘’๐‘— โˆˆ ฮฆฬ‚
way, and then continue until the last one.
1โ‰ค๐‘— โ‰ค๐‘‘
and set
This is (ii), and (i) is a consequence of (ii).
Coming back to the utility maximization problem, note that property
(iii) implies that the value functions and the opportunity processes for the
models
(๐‘…, C )
and
หœ Cหœ)
(๐‘…,
the sequel. Furthermore, if
coincide up to evanescence; we identify them in
๐‘”หœ denotes
the analogue of
๐‘”
in the model
หœ Cหœ),
(๐‘…,
cf. (6.1), we have the relation
๐‘”หœ(๐‘ฆ) = ๐‘”(ฮฆ๐‘ฆ),
๐‘ฆ โˆˆ Cหœ0 .
๐‘”หœ is equivalent to nding one for ๐‘” and if (หœ
๐œ‹ , ๐œ…) is
หœ Cหœ) then (ฮฆหœ
(๐‘…,
๐œ‹ , ๐œ…) is optimal for (๐‘…, C ). In fact,
หœ Cหœ), in particular
interest carry over from (๐‘…, C ) to (๐‘…,
Finding a maximizer for
an optimal strategy for
most properties of
any no-arbitrage property that is dened via the set of admissible (positive)
wealth processes.
Remark 7.4. A classical no-arbitrage condition dened in a slightly dierent
๐‘„ โ‰ˆ ๐‘ƒ under which โ„ฐ(๐‘…) is a
หœ is even
๐œŽ -martingale; cf. Delbaen and Schachermayer [17]. In this case, โ„ฐ(๐‘…)
a local martingale under ๐‘„, as it is a ๐œŽ -martingale with positive components.
way is that there exist a probability measure
Property (ii) from Theorem 7.3 is useful to apply the following result.
Lemma 7.5. Let ๐‘ โˆˆ (0, 1) and assume ๐‘’๐‘— โˆˆ C โˆฉ C 0,โˆ— for 1 โ‰ค ๐‘— โ‰ค ๐‘‘. Then
๐‘ข(๐‘ฅ0 ) < โˆž implies that ๐น ๐‘…,๐ฟ is ๐‘-suitable. If, in
there exists a
โˆซ addition,
๐‘
constant ๐‘˜1 such that ๐ท โ‰ฅ ๐‘˜1 > 0, it follows that {โˆฃ๐‘ฅโˆฃ>1} โˆฃ๐‘ฅโˆฃ ๐น ๐‘… (๐‘‘๐‘ฅ) < โˆž.
Proof.
As
๐‘ > 0
Section III.2.2.
and
๐‘ข(๐‘ฅ0 ) < โˆž, ๐ฟ
is well dened and
๐ฟ, ๐ฟโˆ’ > 0
by
No further properties were used to establish Lemma 3.4,
๐‘ƒ โŠ— ๐ด-a.e.
๐’œ = ๐’œ๐‘“ ๐ธ . In
โˆซ for allโ€ฒ ๐œ‹{โˆˆโˆ’1
particular, from the denition of ๐‘” , it follows that (๐ฟโˆ’ +๐‘ฅ ) ๐‘
(1+๐œ‹ โŠค ๐‘ฅ)๐‘ โˆ’
}
๐‘โˆ’1 โˆ’ ๐œ‹ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )) is nite. If ๐ท โ‰ฅ ๐‘˜1โˆซ, {
Lemma II.3.5 shows that
๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’
๐ฟ โ‰ฅ ๐‘˜1}, hence ๐ฟโˆ’ + ๐‘ฅโ€ฒ โ‰ฅ ๐‘˜1 ๐น ๐ฟ (๐‘‘๐‘ฅโ€ฒ )-a.e. and
โŠค
๐‘…
๐œ‹ โ„Ž(๐‘ฅ) ๐น (๐‘‘๐‘ฅ) < โˆž. We choose ๐œ‹ = ๐‘’๐‘— (and ๐œ… arbitrary) for 1 โ‰ค ๐‘— โ‰ค ๐‘‘ to
whose formula shows that
deduce the result.
๐‘”(๐œ‹)
is nite
66
III Bellman Equation
๐‘ข(๐‘ฅ0 ) < โˆž does not imply any properties of ๐‘…;
C = {0} or C 0,โˆ— = {0}. The transformation
C and C 0,โˆ— such that Theorem 7.3(ii) holds, and
In general, the condition
for instance, in the trivial cases
changes the geometry of
then the situation is dierent.
Corollary 7.6. Let ๐‘ โˆˆ (0, 1) and ๐‘ข(๐‘ฅ0 ) < โˆž. In the model
หœ
Theorem 7.3, ๐น ๐‘…,๐ฟ
is ๐‘-suitable and hence ๐‘”หœ is continuous.
หœ Cหœ)
(๐‘…,
Therefore, to prove Lemma 3.8 under (C4), we may substitute
หœ Cหœ)
(๐‘…,
of
(๐‘…, C )
๐‘-suitable approximating sequences. In some
(๐‘…, C ). In particular, if the asset prices
๐‘—
are strictly positive (ฮ”๐‘… > โˆ’1 for 1 โ‰ค ๐‘— โ‰ค ๐‘‘), then the positive orthant of
โ„๐‘‘ is contained in C 0,โˆ— and the condition of Lemma 7.5 is satised as soon
as ๐‘’๐‘— โˆˆ C for 1 โ‰ค ๐‘— โ‰ค ๐‘‘.
by
and avoid the use of
cases, Lemma 7.5 applies directly in
III.8 Appendix C: Omitted Calculation
This appendix contains a calculation which was omitted in the proof of
Proposition 5.12.
Lemma 8.1. Let (โ„“, ๐œ‹ห‡ , ๐œ…ห‡) be a solution of the Bellman equation, (๐œ‹, ๐œ…) โˆˆ ๐’œ,
ห‡ := ๐‘‹(ห‡
¯ = ๐‘… โˆ’ (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘… as well as
๐‘‹ := ๐‘‹(๐œ‹, ๐œ…) and ๐‘‹
๐œ‹, ๐œ…
ห‡ ). Dene ๐‘…
ห‡ ๐‘โˆ’1 ๐‘‹ satises
๐œ‹
¯ := (๐‘ โˆ’ 1)ห‡
๐œ‹ + ๐œ‹ and ๐œ…
¯ := (๐‘ โˆ’ 1)ห‡
๐œ… + ๐œ…. Then ๐œ‰ := โ„“๐‘‹
(
ห‡ ๐‘โˆ’1 ๐‘‹โˆ’
๐‘‹
โˆ’
)โˆ’1
โˆ™
๐œ‰=
(
)โŠค
¯ โˆ’ โ„“โˆ’ ๐œ…
¯ โˆ™ ๐œ‡ + โ„“โˆ’ (๐‘ โˆ’ 1) ๐‘โˆ’2
โ„“ โˆ’ โ„“0 + โ„“โˆ’ ๐œ‹
¯โˆ™๐‘…
ห‡ + ๐œ‹ ๐‘๐‘… ๐œ‹
ห‡ โˆ™๐ด+๐œ‹
¯ โŠค ๐‘๐‘…โ„“ โˆ™ ๐ด
2 ๐œ‹
{
}
+๐œ‹
¯ โŠค๐‘ฅโ€ฒ โ„Ž(๐‘ฅ) โˆ— ๐œ‡๐‘…,โ„“ + (โ„“โˆ’ + ๐‘ฅโ€ฒ ) (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ) โˆ’ 1 โˆ’ ๐œ‹
¯ โŠค โ„Ž(๐‘ฅ) โˆ— ๐œ‡๐‘…,โ„“ .
Proof.
We may assume
๐‘ฅ0 = 1 .
This calculation is similar to the one in the
proof of Lemma 3.4 and therefore we shall be brief. By Itô's formula we have
ห‡ ๐‘โˆ’1 = โ„ฐ(๐œ)
๐‘‹
Thus
for
๐œ = (๐‘ โˆ’ 1)(ห‡
๐œ‹ โˆ™๐‘…โˆ’๐œ…
ห‡ โˆ™ ๐œ‡) + (๐‘โˆ’1)(๐‘โˆ’2)
๐œ‹
ห‡ โŠค ๐‘๐‘… ๐œ‹
ห‡โˆ™๐ด
2
{
}
+ (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 โˆ’ 1 โˆ’ (๐‘ โˆ’ 1)ห‡
๐œ‹ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘… .
(
)
ห‡ ๐‘โˆ’1 ๐‘‹ = โ„ฐ ๐œ + ๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡ + [๐œ, ๐œ‹ โˆ™ ๐‘…] =: โ„ฐ(ฮจ) with
๐‘‹
โˆ‘
[๐‘…, ๐œ] = [๐‘…๐‘ , ๐œ ๐‘ ] +
ฮ”๐‘…ฮ”๐œ
= (๐‘ โˆ’ 1)๐‘๐‘… ๐œ‹
ห‡ โˆ™ ๐ด + (๐‘ โˆ’ 1)ห‡
๐œ‹ โŠค ๐‘ฅ๐‘ฅ โˆ— ๐œ‡๐‘…
{
}
+ ๐‘ฅ (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 โˆ’ 1 โˆ’ ๐œ‹
ห‡ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘…
and recombining the terms yields
)โŠค
(
ห‡ + ๐œ‹ ๐‘๐‘… ๐œ‹
ห‡โˆ™๐ด
ฮจ=๐œ‹
¯ โˆ™๐‘…โˆ’๐œ…
¯ โˆ™ ๐œ‡ + (๐‘ โˆ’ 1) ๐‘โˆ’2
2 ๐œ‹
{
}
+ (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ) โˆ’ 1 โˆ’ ๐œ‹
¯ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘… .
III.8 Appendix C: Omitted Calculation
Then
(
ห‡ ๐‘โˆ’1 ๐‘‹โˆ’
๐‘‹
โˆ’
)โˆ’1
โˆ™
๐œ‰ = โ„“ โˆ’ โ„“0 + โ„“โˆ’ โˆ™ ฮจ + [โ„“, ฮจ],
[โ„“, ฮจ] = [โ„“๐‘ , ฮจ๐‘ ] +
โˆ‘
67
where
ฮ”โ„“ฮ”ฮจ
=๐œ‹
¯ โŠค ๐‘๐‘…โ„“ โˆ™ ๐ด + ๐œ‹
¯ โŠค ๐‘ฅโ€ฒ ๐‘ฅ โˆ— ๐œ‡๐‘…,โ„“
{
}
+ ๐‘ฅโ€ฒ (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ) โˆ’ 1 โˆ’ ๐œ‹
¯ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘…,โ„“ .
We arrive at
(
ห‡ ๐‘โˆ’1 ๐‘‹โˆ’
๐‘‹
โˆ’
)โˆ’1
โˆ™
๐œ‰=
(
)โŠค
โ„“ โˆ’ โ„“0 + โ„“โˆ’ ๐œ‹
¯ โˆ™ ๐‘… โˆ’ โ„“โˆ’ ๐œ…
¯ โˆ™ ๐œ‡ + โ„“โˆ’ (๐‘ โˆ’ 1) ๐‘โˆ’2
ห‡ + ๐œ‹ ๐‘๐‘… ๐œ‹
ห‡ โˆ™๐ด+๐œ‹
¯ โŠค ๐‘๐‘…โ„“ โˆ™ ๐ด
2 ๐œ‹
{
}
+๐œ‹
¯ โŠค ๐‘ฅโ€ฒ ๐‘ฅ โˆ— ๐œ‡๐‘…,โ„“ + (โ„“โˆ’ + ๐‘ฅโ€ฒ ) (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘โˆ’1 (1 + ๐œ‹ โŠค ๐‘ฅ) โˆ’ 1 โˆ’ ๐œ‹
¯ โŠค ๐‘ฅ โˆ— ๐œ‡๐‘…,โ„“ .
The result follows by writing
๐‘ฅ = โ„Ž(๐‘ฅ) + ๐‘ฅ โˆ’ โ„Ž(๐‘ฅ).
68
III Bellman Equation
Chapter IV
Lévy Models
In this chapter, which corresponds to the article [60], we study power utility
maximization for exponential Lévy models with portfolio constraints and
construct an explicit solution in terms of the Lévy triplet.
IV.1 Introduction
Lévy process and the investor's preferences are given by a power utility function. This
We consider the case when the asset prices follow an exponential
problem was rst studied by Merton [55] for drifted geometric Brownian motion and by Mossin [57] and Samuelson [65] for the discrete-time analogues.
A consistent observation was that when the asset returns are i.i.d., the optimal portfolio and consumption are given by a constant and a deterministic
function, respectively.
This result was subsequently extended to various
classes of Lévy models and its general validity was readily conjecturedwe
note that the existence of an optimal strategy is known also for much more
general models (see Karatzas and ยšitkovi¢ [43]), but
is
some
a priori
that strategy
stochastic process without a constructive description.
We prove this conjecture for general Lévy models under
minimal assump-
tions; in addition, we consider the case where the choice of the portfolio is
investment portfolio
constrained to a convex set. The optimal
ized as the maximizer of a deterministic concave function
of the Lévy triplet; and the maximum of
๐”ค
๐”ค
is character-
dened in terms
yields the optimal
consumption.
Moreover, the Lévy triplet characterizes the niteness of the value function,
i.e., the maximal expected utility.
๐‘ž -optimal
We also draw the conclusions for the
equivalent martingale measures
mization by convex duality (๐‘ž
that are linked to utility maxi-
โˆˆ (โˆ’โˆž, 1) โˆ– {0});
this results in an explicit
existence characterization and a formula for the density process.
Finally,
some generalizations to non-convex constraints are studied.
Our method consists in solving the
Bellman equation,
which was intro-
duced for general semimartingale models in Chapter III. In the Lévy setting,
70
IV Lévy Models
this equation reduces to a Bernoulli ordinary dierential equation. There are
two main mathematical diculties. The rst one is to construct the maximizer for
๐”ค, i.e., the optimal portfolio.
The necessary compactness is obtained
from a minimal no-free-lunch condition (no unbounded increasing prot)
via scaling arguments which were developed by Kardaras [44] for
log-utility.
In our setting these arguments require certain integrability properties of the
asset returns. Without compromising the generality, integrability is achieved
by a linear transformation of the model which replaces the given assets by
certain portfolios.
We construct the maximizer for
๐”ค
in the transformed
model and then revert to the original one.
verify
the optimality of the constructed con-
sumption and investment portfolio.
Here we use the general verication
The second diculty is to
theory of Chapter III and exploit a well-known property of Lévy processes,
namely that any Lévy local martingale is a true martingale.
This chapter is organized as follows. The next section species the optimization problem and the notation related to the Lévy triplet.
recall the no-free-lunch condition
NUIPC
We also
and the opportunity process. Sec-
tion IV.3 states the main result for utility maximization under convex constraints and relates the triplet to the niteness of the value function. The
transformation of the model is described in Section IV.4 and the main theorem is proved in Section IV.5. Section IV.6 gives the application to
๐‘ž -optimal
measures while non-convex constraints are studied in Section IV.7. Related
literature is discussed in the concluding Section IV.8 as this necessitates
technical terminology introduced in the body of the chapter.
IV.2 Preliminaries
๐‘ฅ, ๐‘ฆ โˆˆ โ„ are reals, ๐‘ฅ+ = max{๐‘ฅ, 0} and
๐‘ฅ โˆง ๐‘ฆ = min{๐‘ฅ, ๐‘ฆ}. We set 1/0 := โˆž where necessary. If ๐‘ง โˆˆ โ„๐‘‘ is a
๐‘‘-dimensional vector, ๐‘ง ๐‘– is its ๐‘–th coordinate and ๐‘ง โŠค its transpose. Given
๐ด โŠ† โ„๐‘‘ , ๐ดโŠฅ denotes the Euclidean orthogonal complement and ๐ด is said
to be star-shaped (with respect to the origin) if ๐œ†๐ด โŠ† ๐ด for all ๐œ† โˆˆ [0, 1].
๐‘‘
๐‘‘
If ๐‘‹ is an โ„ -valued semimartingale and ๐œ‹ โˆˆ ๐ฟ(๐‘‹) is an โ„ -valued preThe following notation is used. If
dictable integrand, the vector stochastic integral is a scalar semimartingale
with initial value zero and denoted by
โˆซ
๐œ‹ ๐‘‘๐‘‹
or by
๐œ‹ โˆ™ ๐‘‹.
Relations between
measurable functions hold almost everywhere unless otherwise stated. Our
reference for any unexplained notion or notation from stochastic calculus is
Jacod and Shiryaev [34].
IV.2.1 The Optimization Problem
We x the time horizon
ltration
(โ„ฑ๐‘ก )๐‘กโˆˆ[0,๐‘‡ ]
๐‘‡ โˆˆ (0, โˆž)
and a probability space
(ฮฉ, โ„ฑ, ๐‘ƒ )
with a
satisfying the usual assumptions of right-continuity and
completeness, as well as
โ„ฑ0 = {โˆ…, ฮฉ} ๐‘ƒ -a.s.
We consider an
โ„๐‘‘ -valued
Lévy
IV.2 Preliminaries
process ๐‘… = (๐‘…1 , . . . , ๐‘…๐‘‘ ) with ๐‘…0 = 0.
That is,
71
๐‘… is a càdlàg semimartingale
with stationary independent increments as dened in [34, II.4.1(b)]. It is not
๐‘… generates the ltration. The stochastic exponential
๐‘† = โ„ฐ(๐‘…) = (โ„ฐ(๐‘…1 ), . . . , โ„ฐ(๐‘…๐‘‘ )) represents the discounted price processes of
๐‘‘ risky assets, while ๐‘… stands for their returns. If one wants to model only
relevant for us whether
positive prices, one can equivalently use the ordinary exponential (see, e.g.,
Kallsen [38, Lemma 4.2]). Our agent also has a bank account paying zero
interest at his disposal.
๐‘ฅ0
๐‘‘
is a predictable ๐‘…-integrable โ„ -valued process
The agent is endowed with a deterministic initial capital
trading strategy
the
๐‘–th
> 0. A
๐œ‹ , where
component is interpreted as the fraction of wealth (or the portfolio
proportion) invested in the
๐‘–th
risky asset.
We want to consider two cases. Either consumption occurs only at the
terminal time
๐‘‡
(utility from terminal wealth only); or there is intermediate
consumption plus a bulk consumption at the time horizon.
1
notation, we dene
{
1
๐›ฟ :=
0
To unify the
in the case with intermediate consumption,
otherwise.
It will be convenient to parametrize the consumption strategies as a fraction
of the current wealth.
process
๐œ…
sponding to a pair
Let
C โŠ† โ„๐‘‘
A
propensity to consume
is a nonnegative optional
0 ๐œ…๐‘  ๐‘‘๐‘  < โˆž ๐‘ƒ -a.s. The wealth process ๐‘‹(๐œ‹, ๐œ…)
(๐œ‹, ๐œ…) is dened by the stochastic exponential
(
)
โˆซ
๐‘‹(๐œ‹, ๐œ…) = ๐‘ฅ0 โ„ฐ ๐œ‹ โˆ™ ๐‘… โˆ’ ๐›ฟ ๐œ…๐‘  ๐‘‘๐‘  .
satisfying
โˆซ๐‘‡
be a (constant) set containing the origin; we refer to
constraints. The set of (constrained)
{
๐’œ(๐‘ฅ0 ) := (๐œ‹, ๐œ…) : ๐‘‹(๐œ‹, ๐œ…) > 0
and
admissible
๐œ‹๐‘ก (๐œ”) โˆˆ C
C
corre-
as the
strategies is
for all
}
(๐œ”, ๐‘ก) โˆˆ ฮฉ × [0, ๐‘‡ ] .
(๐œ‹, ๐œ…) โˆˆ ๐’œ,
๐‘ := ๐œ…๐‘‹(๐œ‹, ๐œ…) is the corresponding consumption
โˆซ ๐‘ก rate and ๐‘‹ =
โˆซ ๐‘ก ๐‘‹(๐œ‹, ๐œ…)
satises the self-nancing condition ๐‘‹๐‘ก = ๐‘ฅ0 +
๐‘‹
๐œ‹
๐‘‘๐‘…
โˆ’
๐›ฟ
๐‘ โˆ’ ๐‘ 
๐‘ 
0
0 ๐‘๐‘  ๐‘‘๐‘  as
well as ๐‘‹โˆ’ > 0.
Let ๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1). We use the power utility function
We x the initial capital
๐‘ฅ0
and usually write
๐‘ˆ (๐‘ฅ) := ๐‘1 ๐‘ฅ๐‘ ,
๐’œ for ๐’œ(๐‘ฅ0 ).
Given
๐‘ฅ โˆˆ (0, โˆž)
to model the preferences of the agent. Note that we exclude the well-studied
๐‘ = 0.
๐‘ˆ.
The constant
๐›ฟ ๐‘‘๐‘ก
๐œ‡(๐‘‘๐‘ก)
logarithmic utility (see [44]) which corresponds to
๐›ฝ := (1 โˆ’ ๐‘)โˆ’1 > 0
1
is the relative risk tolerance of
In this chapter, it is more convenient to use the notation
the other chapters.
instead of
as in
IV Lévy Models
72
(๐œ‹, ๐œ…) โˆˆ ๐’œ and ๐‘‹ = ๐‘‹(๐œ‹, ๐œ…), ๐‘ = ๐œ…๐‘‹ . The corresponding expected
[ โˆซ๐‘‡
]
is ๐ธ ๐›ฟ
0 ๐‘ˆ (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ (๐‘‹๐‘‡ ) . The value function is given by
Let
utility
[ โˆซ
๐‘ข(๐‘ฅ0 ) := sup ๐ธ ๐›ฟ
๐’œ(๐‘ฅ0 )
๐‘‡
]
๐‘ˆ (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ (๐‘‹๐‘‡ ) ,
0
(๐‘, ๐‘‹) which correspond to some
(๐œ‹, ๐œ…) โˆˆ ๐’œ(๐‘ฅ0 ). We say that the utility maximization problem is nite if
๐‘ข(๐‘ฅ0 ) < โˆž. This always holds if ๐‘ < 0 as then ๐‘ˆ < 0. If ๐‘ข(๐‘ฅ0 ) < โˆž, (๐œ‹, ๐œ…) is
[ โˆซ
]
optimal if the corresponding (๐‘, ๐‘‹) satisfy ๐ธ ๐›ฟ 0๐‘‡ ๐‘ˆ (๐‘๐‘ก ) ๐‘‘๐‘ก+๐‘ˆ (๐‘‹๐‘‡ ) = ๐‘ข(๐‘ฅ0 ).
where the supremum is taken over all
IV.2.2 Lévy Triplet, Constraints, No-Free-Lunch Condition
(๐‘๐‘… , ๐‘๐‘… , ๐น ๐‘… ) be the Lévy triplet of ๐‘… with respect to some xed cut-o
๐‘‘
๐‘‘
function โ„Ž : โ„ โ†’ โ„ (i.e., โ„Ž is bounded and โ„Ž(๐‘ฅ) = ๐‘ฅ in a neighborhood
๐‘… โˆˆ โ„๐‘‘ , ๐‘๐‘… โˆˆ โ„๐‘‘×๐‘‘ is a nonnegative denite
of ๐‘ฅ = 0). This means that ๐‘
๐‘…
๐‘‘
๐‘…
matrix, and ๐น
is a Lévy measure on โ„ , i.e., ๐น {0} = 0 and
โˆซ
1 โˆง โˆฃ๐‘ฅโˆฃ2 ๐น ๐‘… (๐‘‘๐‘ฅ) < โˆž.
(2.1)
Let
โ„๐‘‘
The process
๐‘…
can be represented as
๐‘…
๐‘…
๐‘…๐‘ก = ๐‘๐‘… ๐‘ก + ๐‘…๐‘ก๐‘ + โ„Ž(๐‘ฅ) โˆ— (๐œ‡๐‘…
๐‘ก โˆ’ ๐œˆ๐‘ก ) + (๐‘ฅ โˆ’ โ„Ž(๐‘ฅ)) โˆ— ๐œ‡๐‘ก .
Here
of
๐‘…
๐œ‡๐‘…
is the integer-valued random measure associated with the jumps
and
๐œˆ๐‘ก๐‘… = ๐‘ก๐น ๐‘…
is its compensator.
๐‘
martingale part, in fact, ๐‘…๐‘ก
๐‘Š
is a
๐‘‘-dimensional
= ๐œŽ๐‘Š๐‘ก ,
where
Moreover,
๐œŽโˆˆ
๐‘…๐‘
is the continuous
โ„๐‘‘×๐‘‘ satises
๐œŽ๐œŽ โŠค = ๐‘๐‘…
and
standard Brownian motion. We refer to [34, II.4] or
Sato [68] for background material concerning Lévy processes.
โ„๐‘‘ to be used in the sequel; the terminology
The rst two are related to the budget constraint ๐‘‹(๐œ‹, ๐œ…) > 0.
We introduce some subsets of
follows [44].
The
natural constraints
are given by
{
}
[
]
C 0 := ๐‘ฆ โˆˆ โ„๐‘‘ : ๐น ๐‘… ๐‘ฅ โˆˆ โ„๐‘‘ : ๐‘ฆ โŠค ๐‘ฅ < โˆ’1 = 0 ;
clearly
C0
is closed, convex, and contains the origin. We also consider the
slightly smaller set
C
0,โˆ—
{
}
[
]
๐‘‘
๐‘…
๐‘‘
โŠค
:= ๐‘ฆ โˆˆ โ„ : ๐น ๐‘ฅ โˆˆ โ„ : ๐‘ฆ ๐‘ฅ โ‰ค โˆ’1 = 0 .
It is convex, contains the origin, and its closure equals
C 0,
but it is a proper
subset in general. The meaning of these sets is explained by
Lemma 2.1. A process ๐œ‹ โˆˆ ๐ฟ(๐‘…) satises โ„ฐ(๐œ‹ โˆ™ ๐‘…) โ‰ฅ 0 (> 0) if and only if
๐œ‹ takes values in C 0 (C 0,โˆ— ) ๐‘ƒ โŠ— ๐‘‘๐‘ก-a.e.
IV.2 Preliminaries
See, e.g., Lemma III.2.5 for the proof. The linear space of
73
null-investments
is dened by
{
}
N := ๐‘ฆ โˆˆ โ„๐‘‘ : ๐‘ฆ โŠค ๐‘๐‘… = 0, ๐‘ฆ โŠค ๐‘๐‘… = 0, ๐น ๐‘… [๐‘ฅ : ๐‘ฆ โŠค ๐‘ฅ โˆ•= 0] = 0 .
๐ป โˆˆ ๐ฟ(๐‘…) satises ๐ป โˆ™ ๐‘… โ‰ก 0 if and only if ๐ป takes values in N
๐‘ƒ โŠ— ๐‘‘๐‘ก-a.e. In particular, two portfolios ๐œ‹ and ๐œ‹ โ€ฒ generate the same wealth
โ€ฒ
process (for given ๐œ…) if and only if ๐œ‹ โˆ’ ๐œ‹ is N -valued.
๐‘‘
๐‘‘
We recall the set 0 โˆˆ C โŠ† โ„ of portfolio constraints. The set J โŠ† โ„
Then
of
immediate arbitrage opportunities
is dened by
โˆซ
{
}
โŠค ๐‘…
๐‘… โŠค
โŠค ๐‘…
J = ๐‘ฆ : ๐‘ฆ ๐‘ = 0, ๐น [๐‘ฆ ๐‘ฅ < 0] = 0, ๐‘ฆ ๐‘ โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ) โ‰ฅ 0 โˆ–N .
๐‘ฆ โˆˆ J , the process ๐‘ฆ โŠค ๐‘… is increasing and
nonconstant.
๐‘‘
ห‡ := โˆฉ
a subset ๐บ of โ„ , its recession cone is given by ๐บ
๐‘Žโ‰ฅ0 ๐‘Ž๐บ. Now
condition NUIPC (no unbounded increasing prot) can be dened by
Note that for
NUIPC
(cf. [44, Theorem 4.5]).
J โŠ†
For
the
J โˆฉ Cห‡ = โˆ…
โ‡โ‡’
This is equivalent to
J โˆฉ (C โˆฉ C 0 )ห‡ = โˆ…
because
Cห‡0 , and it means that if a strategy leads to an increasing nonconstant
wealth process, then that strategy cannot be scaled arbitrarily. This is a very
weak no-free-lunch condition; we refer to [44] for more information about free
lunches in exponential Lévy models. We give a simple example to illustrate
the objects.
Example 2.2.
Assume there is only one asset (๐‘‘
= 1),
that its price is
strictly positive, and that it can jump arbitrarily close to zero and arbitrarily
๐น ๐‘… (โˆ’โˆž, โˆ’1] = 0 and for all ๐œ€ > 0, ๐น ๐‘… (โˆ’1, โˆ’1 + ๐œ€] > 0
> 0.
C 0,โˆ— = [0, 1] and N = {0}. In this situation NUIPC is
satised for any set C , both because J = โˆ… and because Cห‡0 = {0}. If the
๐‘…
0,โˆ— = [0, 1)
price process is merely nonnegative and ๐น {โˆ’1} > 0, then C
high. In formulas,
๐‘… โˆ’1 , โˆž)
and ๐น [๐œ€
0 =
Then C
while the rest stays the same.
In fact, most of the scalar models presented in Schoutens [70] correspond
to the rst part of Example 2.2 for nondegenerate choices of the parameters.
IV.2.3 Opportunity Process
Assume
๐‘ข(๐‘ฅ0 ) < โˆž
(๐œ‹, ๐œ…) โˆˆ{ ๐’œ. For xed ๐‘ก โˆˆ [0, ๐‘‡ ], the set }
of
๐’œ(๐œ‹, ๐œ…, ๐‘ก) := (หœ
๐œ‹, ๐œ…
หœ ) โˆˆ ๐’œ : (หœ
๐œ‹, ๐œ…
หœ ) = (๐œ‹, ๐œ…) on [0, ๐‘ก] .
and let
compatible controls is
By Proposition II.3.1 and Remark II.3.7 there exists a unique càdlàg semimartingale
๐ฟ,
๐ฟ๐‘ก ๐‘1
called
(
opportunity process, such that
[
)๐‘
๐‘‹๐‘ก (๐œ‹, ๐œ…) = ess sup ๐ธ ๐›ฟ
๐’œ(๐œ‹,๐œ…,๐‘ก)
โˆซ
๐‘ก
๐‘‡
]
หœ ๐‘‡ )โ„ฑ๐‘ก ,
๐‘ˆ (หœ
๐‘๐‘  ) ๐‘‘๐‘  + ๐‘ˆ (๐‘‹
IV Lévy Models
74
where the supremum is taken over all consumption and wealth pairs
corresponding to some
(หœ
๐œ‹, ๐œ…
หœ ) โˆˆ ๐’œ(๐œ‹, ๐œ…, ๐‘ก).
หœ
(หœ
๐‘, ๐‘‹)
We shall see that in the present
Lévy setting, the opportunity process is simply a deterministic function. The
right hand side above is known as the
value process
in particular the dynamic value function at time
of our control problem;
๐‘ก
is
๐‘ข๐‘ก (๐‘ฅ) = ๐ฟ๐‘ก ๐‘1 ๐‘ฅ๐‘ .
IV.3 Main Result
We can now formulate the main theorem for the convex-constrained case; the
proofs are given in the two subsequent sections. We consider the following
conditions.
Assumptions 3.1.
(i)
C
is convex.
(ii) The orthogonal projection of
(iii)
NUIPC
(iv)
๐‘ข(๐‘ฅ0 ) < โˆž,
C โˆฉ C0
onto
N
โŠฅ is closed.
holds.
i.e., the utility maximization problem is nite.
To state the result, we dene for
๐”ค(๐‘ฆ) := ๐‘ฆ โŠค ๐‘๐‘… +
(๐‘โˆ’1) โŠค ๐‘…
2 ๐‘ฆ ๐‘ ๐‘ฆ
โˆซ
+
๐‘ฆ โˆˆ C0
the deterministic function
{ โˆ’1
}
๐‘ (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ).
โ„๐‘‘
(3.1)
As we will see later, this concave function is well dened with values in
โ„ โˆช {sign(๐‘)โˆž}.
Theorem 3.2. Under Assumptions 3.1, there exists an optimal strategy
(ห†
๐œ‹, ๐œ…
ห† ) such that ๐œ‹
ห† is a constant vector and ๐œ…
ห† is deterministic. Here ๐œ‹
ห† is
characterized by
๐œ‹
ห† โˆˆ arg maxC โˆฉC 0 ๐”ค
and, in the case with intermediate consumption,
(
)โˆ’1
๐œ…
ห† ๐‘ก = ๐‘Ž (1 + ๐‘Ž)๐‘’๐‘Ž(๐‘‡ โˆ’๐‘ก) โˆ’ 1 ,
where ๐‘Ž :=
๐‘
1โˆ’๐‘
maxC โˆฉC 0 ๐”ค.
The opportunity process is given by
{
(
)
exp ๐‘Ž(1 โˆ’ ๐‘)(๐‘‡ โˆ’ ๐‘ก)
[
]1โˆ’๐‘
๐ฟ๐‘ก =
๐‘Ž๐‘โˆ’1 (1 + ๐‘Ž)๐‘’๐‘Ž(๐‘‡ โˆ’๐‘ก) โˆ’ 1
without intermediate consumption,
with intermediate consumption.
Concerning the question of uniqueness, we recall the following from Remark III.3.3.
Remark 3.3.
and
๐œ…
ห† is unique. The optimal portfolio
N ; i.e., if ๐œ‹ โˆ— is another optimal
๐œ‹
ห† โˆ’ ๐œ‹ โˆ— takes values in N . Equivalently, the
The propensity to consume
arg maxC โˆฉC 0 ๐”ค
are unique modulo
portfolio (or maximizer), then
wealth processes coincide.
IV.3 Main Result
75
We comment on Assumptions 3.1.
Remark 3.4.
(a)
Convexity of
C
is of course not necessary to have a
solution. We give some generalizations in Section IV.7.
(b) Without the closedness in (ii), there are examples with non-existence
of an optimal strategy even for drifted Brownian motion and closed convex
cone constraints; see Example 3.5 below. One can note that closedness of
C
C = C + N , see
C to N โŠฅ
0 = C 0 + N yields ฮ (C โˆฉ C 0 ) =
is closed: if ฮ  denotes the projector, C
0
0
(ฮ C ) โˆฉ C and C is closed. This includes the cases where C is closed and
implies (ii) if
N โŠ†C
and
C
is convex (as this implies
[44, Remark 2.4]). Similarly, (ii) holds whenever the projection of
polyhedral, or compact.
(c)
NUIPC does not hold. If ๐‘ โˆˆ (0, 1), it is obvious that
๐‘ < 0, there exists no optimal strategy, essentially because
Suppose that
๐‘ข(๐‘ฅ0 ) = โˆž.
If
adding a suitable arbitrage strategy would always yield a higher expected
utility. See Karatzas and Kardaras [41, Proposition 4.19] for a proof.
(d)
If
๐‘ข(๐‘ฅ0 ) = โˆž,
either there is no optimal strategy, or there are in-
nitely many strategies yielding innite expected utility. It would be inconvenient to call the latter optimal. Indeed, using that
๐‘ข(๐‘ฅ0 /2) = โˆž,
one can
typically construct such strategies which also exhibit intuitively suboptimal
behavior (such as throwing away money by a suicide strategy; see Harrison
and Pliska [32, ยŸ6.1]). Hence we require (iv) to have a meaningful solution
to our problemthe relevant question is how to characterize this condition
in terms of the model.
The following example is based on Czichowsky et al. [15, ยŸ2.2] and illustrates how non-existence of an optimal portfolio may occur when Assumption 3.1(ii) is violated. We denote by
๐‘–
i.e., ๐‘’๐‘—
๐‘’๐‘— , 1 โ‰ค ๐‘— โ‰ค ๐‘‘
the unit vectors in
โ„๐‘‘ ,
= ๐›ฟ๐‘–๐‘— .
Example 3.5 (๐›ฟ = 0).
Let
โŽ›
โŽž
1 0
0
๐œŽ = โŽ0 1 โˆ’1โŽ  ;
0 โˆ’1 1
๐‘Š
be a standard Brownian motion in
โ„3
and
{
}
2 2 2
C = ๐‘ฆ โˆˆ โ„3 : ๐‘ฆ 1 + ๐‘ฆ 2 โ‰ค ๐‘ฆ 3 , ๐‘ฆ 3 โ‰ฅ 0 .
๐‘…๐‘ก = ๐‘๐‘ก + ๐œŽ๐‘Š๐‘ก , where ๐‘ := ๐‘’1 is orthogonal to ker ๐œŽ โŠค = โ„(0, 1, 1)โŠค .
โŠค
โŠฅ is spanned by ๐‘’ and ๐‘’ โˆ’ ๐‘’ . The closed convex
Thus N = ker ๐œŽ and N
1
2
3
โŠฅ and the orthogonal projection of
cone C is leaning against the plane N
C onto N โŠฅ is an open half-plane plus the origin. The vectors ๐›ผ๐‘’1 with
๐›ผ โˆˆ โ„ โˆ– {0} are not contained in this half-plane but in its closure.
The optimal portfolio ๐œ‹
ห† for the unconstrained problem lies on this boundary. Indeed, NUIPโ„3 holds and Theorem 3.2 yields ๐œ‹
ห† = ๐›ฝ(๐œŽ๐œŽ โŠค )โˆ’1 ๐‘’1 = ๐›ฝ๐‘’1 ,
โˆ’1
where ๐›ฝ = (1 โˆ’ ๐‘)
. This is simply Merton's optimal portfolio in the
Let
market consisting only of the rst asset.
By construction we nd vectors
76
IV Lévy Models
๐œ‹ ๐‘› โˆˆ C whose projections to N โŠฅ
๐ธ[๐‘ˆ (๐‘‹๐‘‡ (๐œ‹ ๐‘› ))] โ†’ ๐ธ[๐‘ˆ (๐‘‹๐‘‡ (ห†
๐œ‹ ))].
converge to
๐œ‹
ห†
and it is easy to see that
Hence the value functions for the con-
strained and the unconstrained problem are identical. Since the solution
of the unconstrained problem is unique modulo
N
constrained problem has a solution, it has to agree with
({ห†
๐œ‹ } + N ) โˆฉ C = โˆ…,
๐œ‹
ห†
, this implies that if the
๐œ‹
ห†,
modulo
N
. But
so there is no solution.
The rest of the section is devoted to the characterization of Assumption 3.1(iv) by the jump characteristic
๐น๐‘…
and the set
C;
this is intimately
related to the moment condition
โˆซ
โˆฃ๐‘ฅโˆฃ๐‘ ๐น ๐‘… (๐‘‘๐‘ฅ) < โˆž.
(3.2)
{โˆฃ๐‘ฅโˆฃ>1}
We start with a partial result; again
๐‘’๐‘— , 1 โ‰ค ๐‘— โ‰ค ๐‘‘ denote the unit vectors.
Proposition 3.6. Let ๐‘ โˆˆ (0, 1).
(i) Under Assumptions 3.1(i)-(iii), (3.2) implies ๐‘ข(๐‘ฅ0 ) < โˆž.
(ii) If ๐‘’๐‘— โˆˆ C โˆฉ C 0,โˆ— for all 1 โ‰ค ๐‘— โ‰ค ๐‘‘, then ๐‘ข(๐‘ฅ0 ) < โˆž implies
By Lemma 2.1 the
๐‘— th
asset has a positive price if and only if
.
(3.2)
๐‘’๐‘— โˆˆ C 0,โˆ— .
Hence we have the following consequence of Proposition 3.6.
Corollary 3.7. In an unconstrained exponential Lévy model with positive
asset prices satisfying NUIPโ„๐‘‘ , ๐‘ข(๐‘ฅ0 ) < โˆž is equivalent to (3.2).
The implication
๐‘ข(๐‘ฅ0 ) < โˆž โ‡’ (3.2) is essentially true also in the general
case; more precisely, it holds in an equivalent model. As a motivation, consider the case where either
C = {0}
or
C 0 = {0}.
The latter occurs, e.g., if
๐‘‘ = 1 and the asset has jumps which are unbounded in both directions. Then
the statement ๐‘ข(๐‘ฅ0 ) < โˆž carries no information about ๐‘… because ๐œ‹ โ‰ก 0 is
the only admissible portfolio. On the other hand, we are not interested in
assets that cannot be traded, and may as well remove them from the model.
This is part of the following result.
หœ Cหœ) of the model
Proposition 3.8. There exists a linear transformation (๐‘…,
(๐‘…, C ), which is equivalent in that it admits the same wealth processes, and
has the following properties:
(i) the prices are strictly positive,
(ii) the wealth can be invested in each asset (i.e., ๐œ‹ โ‰ก ๐‘’๐‘— is admissible),
หœ Cหœ) and
(iii) if ๐‘ข(๐‘ฅ0 ) < โˆž holds for (๐‘…, C ), it holds also in the model (๐‘…,
โˆซ
หœ
๐‘
๐‘…
{โˆฃ๐‘ฅโˆฃ>1} โˆฃ๐‘ฅโˆฃ ๐น (๐‘‘๐‘ฅ) < โˆž.
The details of the construction are given in the next section, where we
also show that Assumptions 3.1 carry over to
หœ Cหœ).
(๐‘…,
IV.4 Transformation of the Model
77
IV.4 Transformation of the Model
This section contains the announced linear transformation of the market
model.
Assumptions 3.1 are not used.
We rst describe how any linear
transformation aects our objects.
Lemma 4.1. Let ฮ› be a ๐‘‘×๐‘‘-matrix and dene ๐‘…หœ := ฮ›๐‘…. Then ๐‘…หœ is a Lévy
process with triplet ๐‘๐‘…หœ = ฮ›๐‘๐‘… , ๐‘๐‘…หœ = ฮ›๐‘๐‘… ฮ›โŠค and ๐น ๐‘…หœ (โ‹…) = ๐น ๐‘… (ฮ›โˆ’1 โ‹…). Moreover, the corresponding natural constraints and null-investments are given
by Cหœ0 := (ฮ›๐‘‡ )โˆ’1 C 0 and Nหœ := (ฮ›๐‘‡ )โˆ’1 N and the corresponding function ๐”คหœ
satises ๐”คหœ(๐‘ง) = ๐”ค(ฮ›โŠค ๐‘ง).
The proof is straightforward and omitted.
the preimage if
ฮ›
ฮ›โˆ’1
refers to
ฮ›, we keep the notation from
๐‘‡
โˆ’1
หœ
C := (ฮ› ) C as well as Cหœ0,โˆ— := (ฮ›๐‘‡ )โˆ’1 C 0,โˆ—
is not invertible.
Lemma 4.1 and introduce also
Of course,
Given
(which is consistent with Section IV.2.2).
Theorem 4.2. There exists a matrix ฮ› โˆˆ โ„๐‘‘×๐‘‘ such that for 1 โ‰ค ๐‘— โ‰ค ๐‘‘,
(i) ฮ”๐‘…หœ๐‘— > โˆ’1 up to evanescence,
(ii) ๐‘’๐‘— โˆˆ Cหœ โˆฉ Cหœ0,โˆ— ,
หœ Cหœ) admits the same wealth processes as (๐‘…, C ).
(iii) the model (๐‘…,
Proof.
๐‘ฆ1 โˆˆ C โˆฉ C 0,โˆ—
1
such that ๐‘ฆ1 โˆ•= 0, if there is no such vector, set ๐‘ฆ1 = 0. We replace the rst
1
โŠค
asset ๐‘… by the process ๐‘ฆ1 ๐‘…. In other words, we replace ๐‘… by ฮ›1 ๐‘…, where
ฮ›1
We treat the components one by one. Pick any vector
is the matrix
โŽ› 1 2
๐‘ฆ1 ๐‘ฆ1 โ‹… โ‹… โ‹…
โŽœ
1
โŽœ
ฮ›1 = โŽœ
..
โŽ
.
๐‘ฆ1๐‘‘
โŽž
โŽŸ
โŽŸ
โŽŸ.
โŽ 
1
โˆ’1 0
โŠค โˆ’1
(ฮ›โŠค
1 ) C and we replace C by (ฮ›1 ) C .
โˆ’1
0,โˆ—
โŠค
0,โˆ—
โŠค
Note that ๐‘’1 โˆˆ (ฮ›1 )
(C โˆฉ C ) because ฮ›1 ๐‘’1 = ๐‘ฆ1 โˆˆ C โˆฉ C by construcโŠค
โŠค
tion. Similarly, (ฮ›1 ๐œ“) โˆ™ ๐‘… = ๐œ“ โˆ™ (ฮ›1 ๐‘…) whenever ฮ›1 ๐œ“ โˆˆ ๐ฟ(๐‘…). Therefore,
The new natural constraints are
to show that the new model admits the same wealth processes as the old
C โˆฉ C 0,โˆ— -valued process ๐œ‹ โˆˆ ๐ฟ(๐‘…) there
โŠค
exists a predictable ๐œ“ such that ฮ›1 ๐œ“ = ๐œ‹ ; note that this already implies
โˆ’1
0,โˆ— ). If ฮ›โŠค is in๐œ“ โˆˆ ๐ฟ(ฮ›1 ๐‘…) and that ๐œ“ takes values in (ฮ›โŠค
1 ) (C โˆฉ C
1
โŠค
โˆ’1
1
vertible, we take ๐œ“ := (ฮ›1 )
๐œ‹ . Otherwise ๐œ‹ โ‰ก 0 by construction and we
1
๐‘—
๐‘—
โŠค
choose ๐œ“ โ‰ก 0 and ๐œ“ = ๐œ‹ for ๐‘— โ‰ฅ 2; this is the same as inverting ฮ›1 on its
one, we have to show that for every
image.
We proceed with the second component of the new model in the same
way, and then continue until the last one. We obtain matrices
ฮ› = ฮ›๐‘‘ โ‹… โ‹… โ‹… ฮ›1 .
โŠค
๐‘’๐‘— โˆˆ (ฮ› )โˆ’1 (C โˆฉ C 0,โˆ— ),
and set
The construction and
ฮ›โŠค
๐‘– ๐‘’๐‘— = ๐‘’๐‘—
ฮ›๐‘— , 1 โ‰ค ๐‘— โ‰ค ๐‘‘
๐‘– โˆ•= ๐‘— imply
for
which is (ii), and (i) is a consequence of (ii).
78
IV Lévy Models
From now on let
ฮ›
and
หœ
๐‘…
be as in Theorem 4.2.
หœ Cหœ) coincide.
Corollary 4.3. (i) The value functions for (๐‘…, C ) and (๐‘…,
หœ Cหœ) coincide.
(ii) The opportunity processes for (๐‘…, C ) and (๐‘…,
(iii) supCหœโˆฉCหœ0,โˆ— ๐”คหœ = supC โˆฉC 0,โˆ— ๐”ค.
(iv) ๐‘ง โˆˆ arg maxCหœโˆฉCหœ0,โˆ— ๐”คหœ if and only if ฮ›โŠค ๐‘ง โˆˆ arg maxC โˆฉC 0,โˆ— ๐”ค.
หœ Cหœ) if and only if (ฮ›โŠค ๐œ‹, ๐œ…) is optimal
(v) (๐œ‹, ๐œ…) is an optimal strategy for (๐‘…,
for (๐‘…, C ).
(vi) NUIPCหœ holds for ๐‘…หœ if and only if NUIPC holds for ๐‘….
Proof.
This follows from Theorem 4.2(iii) and Lemma 4.1.
The transformation also preserves certain properties of the constraints.
Remark 4.4.
(a) If
C
is closed (star-shaped, convex), then
Cหœ is also closed
(star-shaped, convex).
Cหœ is compact only if ฮ› is invertible. However,
โŠค
โŠค ๐‘‘
the relevant properties for Theorem 4.2 are that ฮ› Cหœ = C โˆฉ (ฮ› โ„ ) and
that ๐‘’๐‘— โˆˆ Cหœ for 1 โ‰ค ๐‘— โ‰ค ๐‘‘; we can equivalently substitute Cหœ by a compact set
โŠค is considered as a mapping โ„๐‘‘ โ†’ ฮ›โŠค โ„๐‘‘ , it
having these properties. If ฮ›
โŠค โˆ’1 C = ๐‘“ (C โˆฉฮ›โŠค โ„๐‘‘ )+ker(ฮ›โŠค ).
admits a continuous right-inverse ๐‘“ , and (ฮ› )
โŠค ๐‘‘
๐‘‘
Here ๐‘“ (C โˆฉ ฮ› โ„ ) is compact and contained in ๐ต๐‘Ÿ = {๐‘ฅ โˆˆ โ„ : โˆฃ๐‘ฅโˆฃ โ‰ค ๐‘Ÿ}
[
]
โŠค
๐‘‘
โŠค
ห†
for some ๐‘Ÿ โ‰ฅ 1. The set C := ๐‘“ (C โˆฉ ฮ› โ„ ) + ker(ฮ› ) โˆฉ ๐ต๐‘Ÿ has the two
(b) Let
C
be compact, then
desired properties.
Next, we deal with the projection of
Cหœ โˆฉ Cหœ0
onto
NหœโŠฅ .
We begin with
a coordinate-free description for its closedness; it can be seen as a simple
static version of Czichowsky and Schweizer [14].
Lemma 4.5. Let D โŠ† โ„๐‘‘ be a nonempty set and let D โ€ฒ be its orthogonal
projection onto N โŠฅ . Then D โ€ฒ is closed in โ„๐‘‘ if and only if {๐‘ฆโŠค ๐‘…๐‘‡ : ๐‘ฆ โˆˆ D}
is closed for convergence in probability.
Proof.
N , we may assume that D = D โ€ฒ . If (๐‘ฆ๐‘› )
โŠค
โŠค
limit ๐‘ฆโˆ— , clearly ๐‘ฆ๐‘› ๐‘…๐‘‡ โ†’ ๐‘ฆโˆ— ๐‘…๐‘‡ in probability.
it follows that ๐‘ฆโˆ— โˆˆ D because D โˆฉ N = {0};
Recalling the denition of
D with some
{๐‘ฆ โŠค ๐‘…๐‘‡ : ๐‘ฆ โˆˆ D} is closed,
hence D is closed.
โŠค
Conversely, let ๐‘ฆ๐‘› โˆˆ D and assume ๐‘ฆ๐‘› ๐‘…๐‘‡ โ†’ ๐‘Œ in probability for some
0
๐‘Œ โˆˆ ๐ฟ (โ„ฑ). With D โˆฉ N = {0} it follows that (๐‘ฆ๐‘› ) is bounded, therefore, it
has a subsequence which converges to some ๐‘ฆโˆ— . If D is closed, then ๐‘ฆโˆ— โˆˆ D
โŠค
and we conclude that ๐‘Œ = ๐‘ฆโˆ— ๐‘…๐‘‡ , showing closedness in probability.
is a sequence in
If
Lemma 4.6. Assume that C โˆฉ C 0,โˆ— is dense in C โˆฉ C 0 . Then the orthogonal
projection of C โˆฉ C 0 onto N โŠฅ is closed if and only if this holds for Cหœ โˆฉ Cหœ0
and NหœโŠฅ .
IV.4 Transformation of the Model
Proof.
(i)
ฮ› = ฮ›๐‘‘ โ‹… โ‹… โ‹… ฮ›1 from the proof of The1 โ‰ค ๐‘– โ‰ค ๐‘‘. In a rst step, we
Recall the construction of
orem 4.2. Assume rst that
ฮ› = ฮ›๐‘–
79
for some
show
ฮ›โŠค (Cหœ โˆฉ Cหœ0 ) = C โˆฉ C 0 .
(4.1)
ฮ› is invertible, in which case the claim is clear,
ฮ›โŠค is the orthogonal projection of โ„๐‘‘ onto the hyperplane
๐ป๐‘– = {๐‘ฆ โˆˆ โ„๐‘‘ : ๐‘ฆ ๐‘– = 0} and C โˆฉ C 0,โˆ— โŠ† ๐ป๐‘– . By the density assumption it
0
โŠค โˆ’1 (C โˆฉ C 0 ) = Cหœ โˆฉ Cหœ0 + ๐ป โŠฅ , the sum
follows that C โˆฉ C โŠ† ๐ป๐‘– . Thus (ฮ› )
๐‘–
โŠค
โŠค
โˆ’1
being orthogonal, and ฮ› [(ฮ› )
(C โˆฉ C 0 )] = C โˆฉ C 0 as claimed. We also
By construction, either
or otherwise
note that
(ฮ›โŠค )โˆ’1 (C โˆฉ C 0,โˆ— ) โŠ† (ฮ›โŠค )โˆ’1 (C โˆฉ C 0 )
is dense.
(4.2)
หœ๐‘‡ : ๐‘ฆหœ โˆˆ Cหœโˆฉ Cหœ0 } and now
{๐‘ฆ โŠค ๐‘…๐‘‡ : ๐‘ฆ โˆˆ C โˆฉ C 0 } = {หœ
๐‘ฆโŠค๐‘…
the result follows from Lemma 4.5, for the special case ฮ› = ฮ›๐‘– .
0
โŠค โˆ’1 โˆ˜ โ‹… โ‹… โ‹… โˆ˜ (ฮ›โŠค )โˆ’1 (C โˆฉ C 0 ).
(ii) In the general case, we have Cหœโˆฉ Cหœ = (ฮ›๐‘‘ )
1
We apply part (i) successively to ฮ›1 , . . . , ฮ›๐‘‘ to obtain the result, here (4.2)
Using (4.1) we have
ensures that the density assumption is satised in each step.
Lemmaโˆซ 4.7. Let ๐‘ โˆˆ (0, 1) and assume
implies {โˆฃ๐‘ฅโˆฃ>1} โˆฃ๐‘ฅ๐‘— โˆฃ๐‘ ๐น ๐‘… (๐‘‘๐‘ฅ) < โˆž.
Proof.
๐‘’๐‘— โˆˆ C โˆฉ C 0,โˆ— .
Then ๐‘ข(๐‘ฅ0 ) < โˆž
โˆซ
๐‘’๐‘— โˆˆ C 0,โˆ— implies ฮ”๐‘…๐‘— > โˆ’1, i.e., {โˆฃ๐‘ฅโˆฃ>1} โˆฃ๐‘ฅ๐‘— โˆฃ๐‘ ๐น ๐‘… (๐‘‘๐‘ฅ) =
]
[ ๐‘
โˆซ
๐‘— ๐‘
๐‘— + ๐‘ ๐‘…
{โˆฃ๐‘ฅโˆฃ>1} ((๐‘ฅ ) ) ๐น (๐‘‘๐‘ฅ). Moreover, we have ๐ธ ๐‘ฅ0 โ„ฐ(๐‘… )๐‘‡[ โ‰ค ๐‘ข(๐‘ฅ]0 ). There
๐‘— ๐‘
๐‘
๐‘— ๐‘
exists a Lévy process ๐‘ such that โ„ฐ(๐‘… ) = ๐‘’ , hence ๐ธ โ„ฐ(๐‘… )๐‘‡ < โˆž im๐‘—
๐‘
๐‘— ๐‘
plies that โ„ฐ(๐‘… ) is of class (D) (cf. [38, Lemma 4.4]). In particular, โ„ฐ(๐‘… )
Note that
has a Doob-Meyer decomposition with a well dened drift (predictable nite variation part).
๐‘Œ =
โ„ฐ(๐‘…๐‘— )โˆ’๐‘
โˆ’
the integrand
The stochastic logarithm
๐‘Œ
of
โ„ฐ(๐‘…๐‘— )๐‘
is given by
โ„ฐ(๐‘…๐‘— )๐‘ and the drift of ๐‘Œ is again well dened because
โ„ฐ(๐‘…๐‘— )โˆ’๐‘
โˆ’ is locally bounded. Itô's formula shows that ๐‘Œ is a
โˆ™
Lévy process with drift
๐ด๐‘Œ๐‘ก
(
= ๐‘ก ๐‘(๐‘๐‘… )๐‘— +
In particular,
โˆซ
๐‘(๐‘โˆ’1) ๐‘… ๐‘—๐‘—
(๐‘ )
2
{โˆฃ๐‘ฅโˆฃ>1} ((๐‘ฅ
โˆซ
{
+
(1 + ๐‘ฅ ) โˆ’ 1 โˆ’
โ„๐‘‘
๐‘— )+ )๐‘ ๐น ๐‘… (๐‘‘๐‘ฅ)
๐‘— ๐‘
๐‘๐‘’โŠค
๐‘— โ„Ž(๐‘ฅ)
}
)
๐น (๐‘‘๐‘ฅ) .
๐‘…
< โˆž.
Note that the previous lemma shows Proposition 3.6(ii); moreover, in the
general case, we obtain the desired integrability in the transformed model.
Corollary 4.8.
Proof.
๐‘ข(๐‘ฅ0 ) < โˆž
implies
โˆซ
หœ
๐‘ ๐‘…
{โˆฃ๐‘ฅโˆฃ>1} โˆฃ๐‘ฅโˆฃ ๐น (๐‘‘๐‘ฅ)
< โˆž.
By Theorem 4.2(ii) and Corollary 4.3(i) we can apply Lemma 4.7 in
the model
หœ Cหœ).
(๐‘…,
80
IV Lévy Models
Remark 4.9.
The transformation in Theorem 4.2 preserves the Lévy struc-
ture. Theorem 4.2 and Lemma 4.7 were generalized to semimartingale models in Appendix B of Chapter III. There, additional assumptions are required
for measurable selections and particular structures of the model are not preserved in general.
IV.5 Proof of Theorem 3.2
Our aim is to prove Theorem 3.2 and Proposition 3.6(i). We shall see that
we may replace Assumption 3.1(iv) by the integrability condition (3.2). Under (3.2), we will obtain the fact that
๐‘ข(๐‘ฅ0 ) < โˆž as we construct the optimal
strategies, and that will yield the proof for both results.
IV.5.1 Solution of the Bellman Equation
We start with informal considerations to construct a candidate solution,
which we then verify. If there is an optimal strategy, Theorem III.3.2 states
that the drift rate
๐‘Ž๐ฟ
of the opportunity process
๐ฟ (a special semimartingale
in general) satises the Bellman equation
๐‘/(๐‘โˆ’1)
๐‘Ž๐ฟ ๐‘‘๐‘ก = ๐›ฟ(๐‘ โˆ’ 1)๐ฟโˆ’
where
๐‘”
๐‘‘๐‘ก โˆ’ ๐‘ max ๐‘”(๐‘ฆ) ๐‘‘๐‘ก;
๐‘ฆโˆˆC โˆฉC 0
๐ฟ๐‘‡ = 1,
(5.1)
is the following function, stated in terms of the joint dierential
semimartingale characteristics of
๐‘”(๐‘ฆ) = ๐ฟโˆ’ ๐‘ฆ
โˆซ
+
โŠค
(
๐‘…
๐‘ +
๐‘๐‘…๐ฟ
๐ฟโˆ’
+
(๐‘…, ๐ฟ):
(๐‘โˆ’1) ๐‘…
2 ๐‘ ๐‘ฆ
)
โˆซ
๐‘ฅโ€ฒ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ ))
+
โ„๐‘‘ ×โ„
{
}
(๐ฟโˆ’ + ๐‘ฅโ€ฒ ) ๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘…,๐ฟ (๐‘‘(๐‘ฅ, ๐‘ฅโ€ฒ )).
โ„๐‘‘ ×โ„
In this equation one should see the characteristics of
and
๐ฟ
tics of
๐‘…
as the driving terms
as the solution. In the present Lévy case, the dierential characteris-
๐‘…
are given by the Lévy triplet, in particular, they are deterministic.
To wit, there is
no exogenous stochasticity
in (5.1). Therefore we can expect
that the opportunity process is deterministic.
We make a smooth
Ansatz
โ„“ for ๐ฟ. As โ„“ has no jumps and vanishing martingale part, ๐‘” reduces to
๐‘”(๐‘ฆ) = โ„“๐”ค(๐‘ฆ), where ๐”ค is as (3.1). We show later that this deterministic function is well dened.
For the maximization in (5.1), we have the following
result.
Lemma 5.1. Suppose that Assumptions 3.1(i)-(iv) hold, or alternatively
that Assumptions 3.1(i)-(iii) and (3.2) hold. Then ๐”คโˆ— := supC โˆฉC 0 ๐”ค < โˆž
and there exists a nonrandom vector ๐œ‹ห‡ โˆˆ C โˆฉ C 0,โˆ— such that ๐”ค(ห‡๐œ‹) = ๐”คโˆ— .
IV.5 Proof of Theorem 3.2
The proof is given in the subsequent section. As
โ„“
81
is a smooth function,
its drift rate is simply the time derivative, hence we deduce from (5.1)
๐‘/(๐‘โˆ’1)
๐‘‘โ„“๐‘ก = ๐›ฟ(๐‘ โˆ’ 1)โ„“๐‘ก
๐‘‘๐‘ก โˆ’ ๐‘๐”คโˆ— โ„“๐‘ก ๐‘‘๐‘ก;
This is a Bernoulli ODE. If we denote
๐‘“ (๐‘ก) :=
โ„“๐‘‡ = 1.
๐›ฝ := 1/(1 โˆ’ ๐‘),
the transformation
โ„“๐›ฝ๐‘‡ โˆ’๐‘ก produces the forward linear equation
๐‘‘
๐‘‘๐‘ก ๐‘“ (๐‘ก)
which has, with
๐‘Ž=
=๐›ฟ+
(
)
๐‘
โˆ—
1โˆ’๐‘ ๐”ค ๐‘“ (๐‘ก);
๐‘“ (0) = 1,
๐‘
โˆ—
1โˆ’๐‘ ๐”ค , the unique solution
๐‘“ (๐‘ก) = โˆ’๐›ฟ/๐‘Ž + (1 + ๐›ฟ/๐‘Ž)๐‘’๐‘Ž๐‘ก .
Therefore,
{
โˆ—
๐‘’(๐‘Ž/๐›ฝ)(๐‘‡ โˆ’๐‘ก) = ๐‘’๐‘๐”ค (๐‘‡ โˆ’๐‘ก)
โ„“๐‘ก =
(
)1/๐›ฝ
๐‘Žโˆ’1/๐›ฝ (1 + ๐‘Ž)๐‘’๐‘Ž(๐‘‡ โˆ’๐‘ก) โˆ’ 1
If we dene
(
)โˆ’1
๐œ…
ห‡ ๐‘ก := โ„“โˆ’๐›ฝ
= ๐‘Ž (1 + ๐‘Ž)๐‘’๐‘Ž(๐‘‡ โˆ’๐‘ก) โˆ’ 1 ,
๐‘ก
if
๐›ฟ = 0,
if
๐›ฟ = 1.
then
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
is a solution
of the Bellman equation in the sense of Denition III.4.1; note that the martingale part vanishes.
process. We want to
Let also
verify
nity process and that
ห‡ = ๐‘‹(ห‡
๐‘‹
๐œ‹, ๐œ…
ห‡)
be the corresponding wealth
this solution, i.e., to show that
(ห‡
๐œ‹, ๐œ…
ห‡)
โ„“
is the opportu-
is optimal. We shall use the following result; it
is a special case of Proposition III.4.7 and Theorem III.5.15.
Lemma 5.2. The process
ห‡๐‘ + ๐›ฟ
ฮ“ := โ„“๐‘‹
โˆซ
ห‡ ๐‘ ๐‘ ๐‘‘๐‘ 
๐œ…
ห‡ ๐‘  โ„“๐‘  ๐‘‹
(5.2)
is a local martingale. If C is convex, then ฮ“ is a martingale if and only if
๐‘ข(๐‘ฅ0 ) < โˆž and (ห‡
๐œ‹, ๐œ…
ห‡ ) is optimal and โ„“ is the opportunity process.
To check that
ฮ“
is a martingale, it is convenient to consider the closely
related process
{
}
ฮจ := ๐‘ห‡
๐œ‹ โŠค ๐‘…๐‘ + (1 + ๐œ‹
ห‡ โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ).
Remark III.5.9 shows that
๐œ‹
ห‡
a positive local martingale and that
ฮ“
Now ฮจ has an advantageous structure:
ฮจ is a semimartingale with constant characteristics. In
โ„ฐ(ฮจ) is an exponential Lévy local martingale, therefore auto-
is a martingale as soon as
as
โ„ฐ(ฮจ) is
โ„ฐ(ฮจ) is.
is constant,
other words,
matically a true martingale (e.g., [38, Lemmata 4.2, 4.4]).
Hence we can
apply Lemma 5.2 to nish the proofs of Theorem 3.2 and Proposition 3.6(i),
modulo Lemma 5.1.
IV Lévy Models
82
IV.5.2 Proof of Lemma 5.1: Construction of the Maximizer
Our goal is to show Lemma 5.1. In this section we will use that
C
is star-
shaped, but not its convexity. For convenience, we state again the denition
๐”ค(๐‘ฆ) = ๐‘ฆ โŠค ๐‘๐‘… +
(๐‘โˆ’1) โŠค ๐‘…
2 ๐‘ฆ ๐‘ ๐‘ฆ
โˆซ
{
+
}
๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ).
โ„๐‘‘
(5.3)
The following lemma is a direct consequence of this formula and does not
depend on Assumptions 3.1; it is a simplied version of Lemma III.6.2.
Lemma 5.3. (i) If ๐‘ โˆˆ (0, 1), ๐”ค is well dened with values in (โˆ’โˆž, โˆž]
and lower semicontinuous on C 0 . If (3.2) holds, ๐”ค is nite and continuous on C 0 .
(ii) If ๐‘ < 0, ๐”ค is well dened with values in [โˆ’โˆž, โˆž) and upper semicontinuous on C 0 . Moreover, ๐”ค is nite on Cห‡ and ๐”ค(๐‘ฆ) = โˆ’โˆž for
๐‘ฆ โˆˆ C 0 โˆ– C 0,โˆ— .
Proof.โˆซ Fix {a sequence ๐‘ฆ๐‘› โ†’ ๐‘ฆ in C 0 . A Taylor
expansion and (2.1) show
}
๐‘โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ) is nite and continuous
โˆ’1 .
along (๐‘ฆ๐‘› ) for ๐œ€ small enough, e.g., ๐œ€ = (2 sup๐‘› โˆฃ๐‘ฆ๐‘› โˆฃ)
โˆ’1 (1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ ๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) โ‰ค โˆ’๐‘โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) and
If ๐‘ < 0, we have ๐‘
{ โˆ’1
}
โˆซ
โŠค ๐‘
โˆ’1 โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ)
Fatou's lemma shows that
โˆฃ๐‘ฅโˆฃ>๐œ€ ๐‘ (1 + ๐‘ฆ ๐‘ฅ) โˆ’ ๐‘
is upper semicontinuous of with respect to ๐‘ฆ . For ๐‘ > 0 we have the converse inequality and the same argument yields lower semicontinuity. If ๐‘ > 0
that
โˆฃ๐‘ฅโˆฃโ‰ค๐œ€
and (3.2) holds, the integral is nite and dominated convergence yields continuity.
๐‘ < 0.
๐œ† โˆˆ [0, 1).
Let
For niteness on
Cห‡ we
note that
๐”ค
is even nite on
๐œ†C 0
for
0
โŠค
๐‘…
any
Indeed, ๐‘ฆ โˆˆ ๐œ†C implies ๐‘ฆ ๐‘ฅ โ‰ฅ โˆ’๐œ† > โˆ’1 ๐น (๐‘‘๐‘ฅ)-a.e., hence
๐‘…
the integrand in (5.3) is bounded ๐น -a.e. and we conclude by (2.1). The last
0
0,โˆ— as well as (5.3).
claim is immediate from the denitions of C and C
Assume the alternative version of Lemma 5.1 under Assumptions 3.1(i)(iii) and (3.2) has already been proved; we argue that the complete claim
of that lemma then follows.
Indeed, suppose that Assumptions 3.1(i)-(iv)
C โˆฉ C 0,โˆ— is dense in C โˆฉ C 0 . To see this, note
โˆ–
๐‘› โˆˆ โ„• we have ๐‘ฆ๐‘› := (1 โˆ’ ๐‘›โˆ’1 )๐‘ฆ โ†’ ๐‘ฆ and
denition) and also in C , due to the star-shape. Using
hold. We rst observe that
that for
๐‘ฆ๐‘›
is in
C0
๐‘ฆโˆˆC โˆฉ
C 0,โˆ— (by the
C 0,โˆ— and
Section IV.4 and its notation, Assumptions 3.1(i)-(iv) now imply that the
transformed model
หœ Cหœ)
(๐‘…,
satises Assumptions 3.1(i)-(iii) and (3.2).
We
apply the above alternative version of Lemma 5.1 in that model to obtain
หœโˆ— := supCหœโˆฉCหœ0 ๐”ค
หœ < โˆž and a vector ๐œ‹
หœ(หœ
หœโˆ— . The
๐”ค
หœ โˆˆ Cหœ โˆฉ Cหœ0,โˆ— such that ๐”ค
๐œ‹) = ๐”ค
0,โˆ— observed above and the semicontinuity from Lemma 5.3(i)
density of C โˆฉC
โˆ—
imply that ๐”ค := supC โˆฉC 0 ๐”ค = supC โˆฉC 0,โˆ— ๐”ค. Using this argument also for
หœ Cหœ), Corollary 4.3(iii) yields ๐”คโˆ— = ๐”ค
หœโˆ— < โˆž. Moreover, Corollary 4.3(iv)
(๐‘…,
states that ๐œ‹
ห‡ := ฮ›โŠค ๐œ‹
หœ โˆˆ C โˆฉ C 0,โˆ— is a maximizer for ๐”ค.
IV.5 Proof of Theorem 3.2
83
Summarizing this discussion, it suces to prove Lemma 5.1 under Assumptions 3.1(i)-(iii) and (3.2); hence these will be our assumptions for the
rest of the section.
Formally, by dierentiation under the integral, the directional derivatives
of
๐”ค
(หœ
๐‘ฆ โˆ’ ๐‘ฆ)โŠค โˆ‡๐”ค(๐‘ฆ) = ๐”Š(หœ
๐‘ฆ , ๐‘ฆ), with
(
)
๐”Š(หœ
๐‘ฆ , ๐‘ฆ) := (หœ
๐‘ฆ โˆ’ ๐‘ฆ)โŠค ๐‘๐‘… + (๐‘ โˆ’ 1)๐‘๐‘… ๐‘ฆ
โˆซ {
}
(หœ
๐‘ฆ โˆ’ ๐‘ฆ)โŠค ๐‘ฅ
โŠค
+
โˆ’
(หœ
๐‘ฆ
โˆ’
๐‘ฆ)
โ„Ž(๐‘ฅ)
๐น ๐‘… (๐‘‘๐‘ฅ).
โŠค ๐‘ฅ)1โˆ’๐‘
(1
+
๐‘ฆ
๐‘‘
โ„
are given by
We take this as the denition of
๐”Š(หœ
๐‘ฆ , ๐‘ฆ)
(5.4)
whenever the integral makes sense.
Remark 5.4. Formally setting ๐‘ = 0, we see that ๐”Š corresponds to the relative rate of return of two portfolios in the theory of log-utility [41, Eq. (3.2)].
Lemma 5.5. Let ๐‘ฆหœ โˆˆ C 0 . On the set C 0 โˆฉ {๐”ค > โˆ’โˆž}, ๐”Š(หœ๐‘ฆ, โ‹…) is well
dened with values in (โˆ’โˆž, โˆž]. Moreover, ๐”Š(0, โ‹…) is lower semicontinuous
on C 0 โˆฉ {๐”ค > โˆ’โˆž}.
Proof.
The rst part follows by rewriting
๐”Š(หœ
๐‘ฆ , ๐‘ฆ)
as
โˆซ
(
)
{
}
(หœ
๐‘ฆ โˆ’ ๐‘ฆ)โŠค ๐‘๐‘… + (๐‘ โˆ’ 1)๐‘๐‘… ๐‘ฆ โˆ’
(1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘ โˆ’ 1 โˆ’ ๐‘๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ)
โˆซ {
}
1 + ๐‘ฆหœโŠค ๐‘ฅ
โŠค
+
โˆ’
1
โˆ’
(หœ
๐‘ฆ
+
(๐‘
โˆ’
1)๐‘ฆ)
โ„Ž(๐‘ฅ)
๐น ๐‘… (๐‘‘๐‘ฅ)
(1 + ๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
1+ ๐‘ฆหœโŠค ๐‘ฅ โ‰ฅ 0 ๐น ๐‘… -a.e. by denition
denition of ๐”Š. Using
because the rst integral occurs in (5.3) and
0
of C . Let
๐‘ โˆˆ (0, 1)
and
๐‘ฆหœ = 0
in the
โˆ’๐‘ฆ โŠค ๐‘ฅ
1 + ๐‘ฆโŠค๐‘ฅ
โ‰ฅ
โˆ’
= โˆ’(1 + ๐‘ฆ โŠค ๐‘ฅ)๐‘
(1 + ๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
(1 + ๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
and (3.2), Fatou's lemma yields that
๐‘ง/(1 +
๐‘ง)1โˆ’๐‘
โ‰ค1
for
๐‘งโ‰ฅ0
implies
๐”Š(0, โ‹…)
is l.s.c. on
โˆ’๐‘ฆ โŠค ๐‘ฅ
(1+๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
โ‰ฅ โˆ’1.
C 0.
If
๐‘ < 0,
then
Again, Fatou's lemma
yields the claim.
As our goal is to prove Lemma 5.1, we may assume in the following that
C โˆฉ C0 โŠ† N
โŠฅ
.
๐”ค(๐‘ฆ) = ๐”ค(๐‘ฆ + ๐‘ฆ โ€ฒ ) for ๐‘ฆ โ€ฒ โˆˆ N , we may substitute C โˆฉ C 0
โŠฅ . The remainder of the section parallels the case
to N
Indeed, noting that
by its projection
log-utility as treated in [44, Lemmata 5.2,5.1]. By Lemmata 5.3 and 5.5,
๐”Š(0, ๐‘ฆ) is well dened with values in (โˆ’โˆž, โˆž] for ๐‘ฆ โˆˆ Cห‡, so the following
of
statement makes sense.
Lemma 5.6. Let ๐‘ฆ โˆˆ (C โˆฉ C 0 )ห‡, then ๐‘ฆ โˆˆ J if and only if ๐”Š(0, ๐‘Ž๐‘ฆ) โ‰ค 0 for
all ๐‘Ž โ‰ฅ 0.
84
IV Lévy Models
Proof.
๐‘ฆ โˆˆ J , then ๐”Š(0, ๐‘Ž๐‘ฆ) โ‰ค 0 by the denitions of J and ๐”Š; we prove
0ห‡
๐‘…
โŠค
the converse. As ๐‘ฆ โˆˆ (C โˆฉ C ) implies ๐น [๐‘Ž๐‘ฆ ๐‘ฅ < โˆ’1] = 0 for all ๐‘Ž, we
๐‘ฆโŠค ๐‘ฅ
๐‘… โŠค
โŠค
have ๐น [๐‘ฆ ๐‘ฅ < 0] = 0. Since ๐‘ฆ ๐‘ฅ โ‰ฅ 0 entails โˆฃ
โˆฃ โ‰ค 1, dominated
(1+๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
If
convergence yields
lim
๐‘Žโ†’โˆž
โˆซ {
}
๐‘ฆโŠค๐‘ฅ
โŠค
โˆ’
๐‘ฆ
โ„Ž(๐‘ฅ)
๐น ๐‘… (๐‘‘๐‘ฅ) = โˆ’
(1 + ๐‘Ž๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
โˆซ
๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น ๐‘… (๐‘‘๐‘ฅ).
โˆ’๐‘Žโˆ’1 ๐”Š(0, ๐‘Ž๐‘ฆ) โ‰ฅ 0, i.e,
โˆซ {
}
๐‘ฆโŠค๐‘ฅ
โŠค
๐‘ฆ โŠค ๐‘๐‘… + ๐‘Ž(๐‘ โˆ’ 1)๐‘ฆ โŠค ๐‘๐‘… ๐‘ฆ +
โˆ’
๐‘ฆ
โ„Ž(๐‘ฅ)
๐น ๐‘… (๐‘‘๐‘ฅ) โ‰ฅ 0.
(1 + ๐‘Ž๐‘ฆ โŠค ๐‘ฅ)1โˆ’๐‘
By assumption,
(๐‘ โˆ’ โˆซ1)๐‘ฆ โŠค ๐‘๐‘… ๐‘ฆ โ‰ค 0, taking ๐‘Ž โ†’ โˆž
๐‘ฆ โŠค ๐‘๐‘… โˆ’ ๐‘ฆ โŠค โ„Ž(๐‘ฅ) ๐น (๐‘‘๐‘ฅ) โ‰ฅ 0.
As
Proof of Lemma 5.1.
shows
๐‘ฆ โŠค ๐‘๐‘… = 0
and then we also see
(๐‘ฆ๐‘› ) โŠ‚ C โˆฉ C 0 be such that ๐”ค(๐‘ฆ๐‘› ) โ†’ ๐”คโˆ— . We may
0,โˆ— by Lemma 5.3, since C 0,โˆ— โŠ† C 0 is dense.
assume ๐”ค(๐‘ฆ๐‘› ) > โˆ’โˆž and ๐‘ฆ๐‘› โˆˆ C
We claim that (๐‘ฆ๐‘› ) has a bounded subsequence. By way of contradiction,
suppose that (๐‘ฆ๐‘› ) is unbounded. Without loss of generality, ๐œ‰๐‘› := ๐‘ฆ๐‘› /โˆฃ๐‘ฆ๐‘› โˆฃ
converges to some ๐œ‰ .
Moreover, we may assume by redening ๐‘ฆ๐‘› that
๐”ค(๐‘ฆ๐‘› ) = max๐œ†โˆˆ[0,1] ๐”ค(๐œ†๐‘ฆ๐‘› ), because ๐”ค is continuous on each of the compact
sets ๐ถ๐‘› = {๐œ†๐‘ฆ๐‘› : ๐œ† โˆˆ [0, 1]}. Indeed, if ๐‘ < 0, continuity follows by domiโŠค
โŠค
โŠค
nated convergence using 1 + ๐œ†๐‘ฆ ๐‘ฅ โ‰ฅ 1 + ๐‘ฆ ๐‘ฅ on {๐‘ฅ : ๐‘ฆ ๐‘ฅ โ‰ค 0}; while for
๐‘ โˆˆ (0, 1), ๐”ค is continuous by Lemma 5.3.
Using concavity one can check that ๐”Š(0, ๐‘Ž๐œ‰๐‘› ) is indeed the directional
derivative of the function ๐”ค at ๐‘Ž๐œ‰๐‘› (cf. Lemma III.5.14). In particular,
๐”ค(๐‘ฆ๐‘› ) = max๐œ†โˆˆ[0,1] ๐”ค(๐œ†๐‘ฆ๐‘› ) implies that ๐”Š(0, ๐‘Ž๐œ‰๐‘› ) โ‰ค 0 for ๐‘Ž > 0 and all ๐‘›
such that โˆฃ๐‘ฆ๐‘› โˆฃ โ‰ฅ ๐‘Ž (and hence ๐‘Ž๐œ‰๐‘› โˆˆ ๐ถ๐‘› ). By the star-shape and closedness
0
0ห‡
of C โˆฉ C we have that ๐œ‰ โˆˆ (C โˆฉ C ) . Lemmata 5.3 and 5.5 yield the semicontinuity to pass from ๐”Š(0, ๐‘Ž๐œ‰๐‘› ) โ‰ค 0 to ๐”Š(0, ๐‘Ž๐œ‰) โ‰ค 0 and now Lemma 5.6
0ห‡
shows ๐œ‰ โˆˆ J , contradicting the NUIPC condition that J โˆฉ (C โˆฉ C ) = โˆ….
Let
We have shown that after passing to a subsequence, there exists a limit
= lim๐‘› ๐‘ฆ๐‘› . Lemma 5.3 shows ๐”คโˆ— = lim๐‘› ๐”ค(๐‘ฆ๐‘› ) = ๐”ค(๐‘ฆ โˆ— ) < โˆž; and ๐‘ฆ โˆ— โˆˆ C 0,โˆ—
โˆ—
0,โˆ— follows as in Lemma III.6.8.
for ๐‘ < 0. For ๐‘ โˆˆ (0, 1), ๐‘ฆ โˆˆ C
๐‘ฆโˆ—
IV.6 ๐‘ž-Optimal Martingale Measures
In this section we consider
๐›ฟ = 0
(no consumption) and
C = โ„๐‘‘ .
Then
Assumptions 3.1 are equivalent to
NUIPโ„d
holds
and
๐‘ข(๐‘ฅ0 ) < โˆž
M be the
๐‘† = โ„ฐ(๐‘…). Then NUIPโ„d
and these conditions are in force for the following discussion. Let
set of all equivalent local martingale measures for
(6.1)
IV.6 ๐‘ž -Optimal Martingale Measures
is equivalent to
๐‘…
M โˆ•= โˆ…,
more precisely, there exists
๐‘„ โˆˆ M
85
under which
is a Lévy martingale (see [44, Remark 3.8]). In particular, we are in the
setting of Kramkov and Schachermayer [49].
๐‘ž = ๐‘/(๐‘ โˆ’ 1) โˆˆ (โˆ’โˆž, 0) โˆช (0, 1) be the exponent conjugate to ๐‘,
โˆ’1 (๐‘‘๐‘„/๐‘‘๐‘ƒ )๐‘ž ] is nite and minimal
then ๐‘„ โˆˆ M is called ๐‘ž -optimal if ๐ธ[โˆ’๐‘ž
over M . If ๐‘ž < 0, i.e., ๐‘ โˆˆ (0, 1), then ๐‘ข(๐‘ฅ0 ) < โˆž is equivalent to the
โˆ’1 (๐‘‘๐‘„/๐‘‘๐‘ƒ )๐‘ž ] < โˆž (cf. Kramkov
existence of some ๐‘„ โˆˆ M such that ๐ธ[โˆ’๐‘ž
Let
and Schachermayer [50]).
This minimization problem over
M
is linked to power utility maximiza-
tion by convex duality in the sense of [49]. More precisely, that article considers a dual problem over an enlarged domain of certain supermartingales.
We recall from Proposition II.4.2 that the solution to that dual problem is
ห† ๐‘โˆ’1 ,
๐‘Œห† = ๐ฟ๐‘‹
where ๐ฟ is the opportuห† = ๐‘ฅ0 โ„ฐ(ห†
๐‘‹
๐œ‹ โˆ™ ๐‘…) is the optimal wealth process corresponding
to ๐œ‹
ห† as in Theorem 3.2. It follows from [49, Theorem 2.2(iv)] that the ๐‘ž ห† exists if and only if ๐‘Œห† is a martingale, and in
optimal martingale measure ๐‘„
ห† /๐‘Œห†0 is the ๐‘ƒ -density process of ๐‘„
ห† . Recall the functions ๐”ค and ๐”Š
that case ๐‘Œ
given by the positive supermartingale
nity process and
from (3.1) and (5.4). A direct calculation (or Remark III.5.18) shows
(
)
{
๐‘Œห† /๐‘Œห†0 = โ„ฐ ๐”Š(0, ๐œ‹
ห† )๐‘ก + (๐‘ โˆ’ 1)ห†
๐œ‹ โŠค ๐‘…๐‘ + (1 + ๐œ‹
ห† โŠค ๐‘ฅ)๐‘โˆ’1 โˆ’ 1} โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) .
Here absence of drift is equivalent to
โŠค ๐‘…
โŠค ๐‘…
โˆซ
๐œ‹
ห† ๐‘ + (๐‘ โˆ’ 1)ห†
๐œ‹ ๐‘ ๐œ‹
ห†+
โ„๐‘‘
{
๐”Š(0, ๐œ‹
ห† ) = 0,
or more explicitly,
}
๐œ‹
ห†โŠค๐‘ฅ
โŠค
โˆ’
๐œ‹
ห†
โ„Ž(๐‘ฅ)
๐น ๐‘… (๐‘‘๐‘ฅ) = 0,
(1 + ๐œ‹
ห† โŠค ๐‘ฅ)1โˆ’๐‘
(6.2)
and in that case
(
)
{
}
๐‘Œห† /๐‘Œห†0 = โ„ฐ (๐‘ โˆ’ 1)ห†
๐œ‹ โŠค ๐‘…๐‘ + (1 + ๐œ‹
ห† โŠค ๐‘ฅ)๐‘โˆ’1 โˆ’ 1 โˆ— (๐œ‡๐‘… โˆ’ ๐œˆ ๐‘… ) .
This is an exponential Lévy martingale because
๐œ‹
ห†
(6.3)
is a constant vector; in
particular, it is indeed a true martingale.
Theorem 6.1. Let ๐‘ž โˆˆ (โˆ’โˆž, 0) โˆช (0, 1). The following are equivalent:
(i) The ๐‘ž-optimal martingale measure ๐‘„ห† exists,
(ii) (6.1) and (6.2) hold,
(iii) there exists ๐œ‹ห† โˆˆ C 0 such that ๐”ค(ห†๐œ‹) = maxC 0 ๐”ค < โˆž and (6.2) holds.
Under these equivalent conditions, (6.3) is the ๐‘ƒ -density process of ๐‘„ห†.
Proof.
exists
We have just argued the equivalence of (i) and (ii). Under (6.1), there
๐œ‹
ห†
satisfying (iii) by Theorem 3.2. Conversely, given (iii) we construct
the solution to the utility maximization problem as before and (6.1) follows;
recall Remark 3.4(c).
IV Lévy Models
86
Remark 6.2.
๐‘ƒ
(i)
If
ห†
๐‘„
exists, (6.3) shows that the change of measure from
ห†
๐‘„
and time-independent) Girsanov param( has constant (deterministic
)
โŠค
๐‘โˆ’1
eters (๐‘ โˆ’ 1)ห†
๐œ‹ , (1 + ๐œ‹
ห† ๐‘ฅ)
; compare [34, III.3.24] or Jeanblanc et al. [35,
ห† . This result was
ยŸA.1, ยŸA.2]. Therefore, ๐‘… is again a Lévy process under ๐‘„
to
previously obtained in [35] by an abstract argument (cf. Section IV.8 below).
ห† is a fairly delicate question compared to the existence
๐‘„
ห† . Recalling the denition of ๐”Š, (6.2) essentially exof the supermartingale ๐‘Œ
0
presses that the budget constraint C in the maximization of ๐”ค is not bind(ii) Existence of
ing. Theorem 6.1 gives an explicit and sharp description for the existence
of
ห†;
๐‘„
this appears to be missing in the previous literature.
IV.7 Extensions to Non-Convex Constraints
In this section we consider the utility maximization problem for some cases
where the constraints
0 โˆˆ C โŠ† โ„๐‘‘
are not convex. Let us rst recapitulate
where the convexity assumption was used above. The proof of Lemma 5.1
used the star-shape of
the shape of
C
C,
but not convexity. In the rest of Section IV.5.1,
was irrelevant except in Lemma 5.2.
We denote by
co (C )
the closed convex hull of
C.
Corollary 7.1. Let ๐‘ < 0 and suppose that either (i) or (ii) below hold:
(i) (a) C is star-shaped,
(b) the orthogonal projection of co (C ) โˆฉ C 0 onto N โŠฅ is closed,
(c) NUIPco (C ) holds.
(ii) C โˆฉ C 0 is compact.
Then the assertion of Theorem 3.2 remains valid.
Proof.
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡ ) is as above; we have to substitute the
Lemma 5.2. The model (๐‘…, co (C )) satises the
(i) The construction of
verication step which used
assumptions of Theorem 3.2. Hence the corresponding opportunity process
๐ฟco (C )
is deterministic and bounded away from zero. The denition of the
opportunity process and the inclusion
process
๐ฟ = ๐ฟC
for
(๐‘…, C )
(โ„“, ๐œ‹
ห‡, ๐œ…
ห‡)
shape of C .
bounded and we can verify
assumptions about the
C โŠ† co (C ) imply that the opportunity
โ„“/๐ฟ is
is also bounded away from zero. Hence
by Corollary III.5.4, which makes no
C = C โˆฉ C 0 . In (i),
0
the star-shape was used only to construct a maximizer for ๐”ค. When C โˆฉC is
(ii) We may assume without loss of generality that
compact, its existence is clear by the upper semicontinuity from Lemma 5.3,
which also shows that any maximizer is necessarily in
C 0,โˆ— .
To proceed as
co (C ) โˆฉ C 0
ห‡
(co (C )) = {0} since
in (i), it remains to note that the projection of the compact set
N
co (C ) is
onto
โŠฅ is compact, and
bounded.
NUIPco (C )
holds because
IV.7 Extensions to Non-Convex Constraints
When the constraints are not star-shaped and
87
๐‘ > 0, an additional con๐”ค is not attained on
dition is necessary to ensure that the maximum of
C 0 โˆ– C 0,โˆ— ,
or equivalently, to obtain a positive optimal wealth process. In
Section III.2.4 we introduced the following condition:
(C3)
๐œ‚ โˆˆ (0, 1)
๐œ‚ โˆˆ (๐œ‚, 1).
There exists
for all
This is clearly satised if
C
such that
๐‘ฆ โˆˆ (C โˆฉ C 0 ) โˆ– C 0,โˆ—
is star-shaped or if
implies
๐œ‚๐‘ฆ โˆˆ C
C 0,โˆ— = C 0 .
Corollary 7.2. Let ๐‘ โˆˆ (0, 1) and suppose that either (i) or (ii) below hold:
(i) Assumptions 3.1 hold except that C is star-shaped instead of being convex.
(ii) C โˆฉ C 0 is compact and satises (C3) and ๐‘ข(๐‘ฅ0 ) < โˆž.
Then the assertion of Theorem 3.2 remains valid.
Proof. (i) The assumptions carry over to the transformed model as before,
hence again we only need to substitute the verication argument. In view
of
๐‘ โˆˆ (0, 1),
we can use Theorem III.5.2, which makes no assumptions
about the shape of
C.
Note that we have already checked its condition (cf.
Remark III.5.16).
(ii) We may again assume
C = C โˆฉC 0 and Remark 4.4 shows that we can
Cหœโˆฉ Cหœ0 to be compact in the transformed model satisfying (3.2). That
is, we can again assume (3.2) without loss of generality. Then ๐”ค is continuous
0 is clear. Under (C3), any
and hence existence of a maximizer on C โˆฉ C
0,โˆ—
maximizer is in C
by the same argument as in the proof of Lemma 5.1.
choose
The following result covers
all
closed constraints and applies to most of
the standard models (cf. Example 2.2).
Corollary 7.3. Let C be closed and assume that C 0 is compact and that
๐‘ข(๐‘ฅ0 ) < โˆž. Then the assertion of Theorem 3.2 remains valid.
Proof. Note that (C3) holds for all sets C when C 0,โˆ— is closed (and hence
equal to
C 0 ).
It remains to apply part (ii) of the two previous corollaries.
Remark 7.4.
(i)
For
๐‘ โˆˆ (0, 1)
we also have the analogue of Proposi-
tion 3.6(i): under the assumptions of Corollary 7.2 excluding
(3.2) implies
(ii)
๐‘ข(๐‘ฅ0 ) < โˆž,
๐‘ข(๐‘ฅ0 ) < โˆž.
The optimal propensity to consume
๐œ…
ห†
remains unique even when
the constraints are not convex (cf. Theorem III.3.2). However, there is no
uniqueness for the optimal portfolio.
corollaries, any constant vector
(by the same proofs); and when
๐œ‹
need not be in
portfolios.
N
In fact, in the setting of the above
๐œ‹ โˆˆ arg maxC โˆฉC 0 ๐”ค is an optimal
C is not convex, the dierence of
portfolio
two such
. See also Remark III.3.3 for statements about dynamic
Conversely, by Theorem III.3.2 any optimal portfolio, possibly
dynamic, takes values in
arg maxC โˆฉC 0 ๐”ค.
IV Lévy Models
88
IV.8 Related Literature
We discuss some related literature in a highly selective manner; an exhaustive
overview is beyond our scope. For the unconstrained utility maximization
problem with general Lévy processes, Kallsen [38] gave a result of verication type: If there exists a vector
๐œ‹
satisfying a certain equation,
๐œ‹
is the
optimal portfolio. This equation is essentially our (6.2) and therefore holds
only if the corresponding dual element
sure.
๐‘Œห†
is the density process of a mea-
Muhle-Karbe [58, Example 4.24] showed that this condition fails in
a particular model, in other words, the supermartingale
๐‘Œห†
is not a martin-
gale in that example. In the one-dimensional case, he introduced a weaker
inequality condition [58, Corollary 4.21], but again existence of
๐œ‹ was not dis-
cussed. (In fact, our proofs show the necessity of that inequality condition;
cf. Remark III.5.16.)
Numerous variants of our utility maximization problem were also studied
along more traditional lines of dynamic programming. E.g., Benth et al. [6]
solve a similar problem with innite time horizon when the Lévy process
satises additional integrability properties and the portfolios are chosen in
[0, 1].
This part of the literature generally requires technical conditions,
which we sought to avoid.
Jeanblanc et al. [35] study the
when
๐‘ž < 0
or
๐‘ž > 1
๐‘ž -optimal
measure
ห†
๐‘„
for Lévy processes
(note that the considered parameter range does not
coincide with ours). They show that the Lévy structure is preserved under
ห†,
๐‘„
if the latter exists; a result we recovered in Remark 6.2 above for our values
of
๐‘ž.
In [35] this is established by showing that starting from any equivalent
change of measure, a suitable choice of constant Girsanov parameters reduces
the
๐‘ž -divergence
of the density.
This argument does not seem to extend
to our general dual problem which involves supermartingales rather than
measures; in particular, it cannot be used to show that the optimal portfolio
is a constant vector.
ห†
๐‘„
A deterministic, but not explicit characterization of
is given in [35, Theorem 2.7]. The authors also provide a more explicit
candidate for the
๐‘ž -optimal
measure [35, Theorem 2.9], but the condition of
that theorem fails in general (see Bender and Niethammer [5]).
In the Lévy setting the
๐‘ž -optimal measures (๐‘ž โˆˆ โ„) coincide with the min-
imal Hellinger measures and hence the pertinent results apply. See Choulli
and Stricker [12] and in particular their general sucient condition [12, Theorem 2.3].
We refer to [35, p. 1623] for a discussion.
that both the existence of
ห†
๐‘„
in terms of the Lévy triplet.
Our result diers in
and its density process are described explicitly
Chapter V
Risk Aversion Asymptotics
In this chapter, which corresponds to the article [62], we use the tools from
Chapters II and III to study the optimal strategy as the relative risk aversion
tends to innity or to one.
V.1 Introduction
We study preferences given by power utility random elds for an agent who
can invest in a nancial market which is modeled by a general semimartingale.
We defer the precise formulation to the next section to allow for a
brief presentation of the contents and focus on the power utility function
๐‘ˆ (๐‘) (๐‘ฅ) =
1 ๐‘
๐‘ ๐‘ฅ , where
there exists for each
๐‘
๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1).
Under standard assumptions,
an optimal trading and consumption strategy that
๐‘ˆ (๐‘) . Our main interest
concerns the behavior of these strategies in the limits ๐‘ โ†’ โˆ’โˆž and ๐‘ โ†’ 0.
(๐‘) tends to innity for ๐‘ โ†’ โˆ’โˆž. Hence
The relative risk aversion of ๐‘ˆ
maximizes the expected utility corresponding to
economic intuition suggests that the agent should become reluctant to take
risks and, in the limit, not invest in the risky assets. Our rst main result
conrms this intuition. More precisely, we prove in a general semimartingale
model that the optimal consumption, expressed as a proportion of current
wealth, converges pointwise to a deterministic function. This function corresponds to the consumption which would be optimal in the case where
trading is not allowed. In the continuous semimartingale case, we show that
the optimal trading strategy tends to zero in a local
๐ฟ2 -sense
and that the
corresponding wealth process converges in the semimartingale topology.
Our second result pertains to the same limit
๐‘ โ†’ โˆ’โˆž
but concerns the
problem without intermediate consumption. In the continuous case, we show
that the optimal trading strategy scaled by
1โˆ’๐‘ converges to a strategy which
is optimal for exponential utility. We provide economic intuition for this fact
via a sequence of auxiliary power utility functions with shifted domains.
The limit
๐‘โ†’0
is related to the logarithmic utility function. Our third
V Risk Aversion Asymptotics
90
main result is the convergence of the corresponding optimal consumption for
the general semimartingale case, and the convergence of the trading strategy
and the wealth process in the continuous case.
All these results are readily observed for special models where the optimal
strategies can be calculated explicitly.
While the corresponding economic
a priori unclear how to go about
Indeed, the problem is to get our hands on the optimal
intuition extends to general models, it is
proving the results.
controls, which is a notorious question in stochastic optimal control.
Our main tool is the opportunity process. We prove its convergence using control-theoretic arguments and convex analysis. On the one hand, this
yields the convergence of the value function. On the other hand, we deduce
the convergence of the optimal consumption, which is directly related to
the opportunity process. The optimal trading strategy is also linked to this
process, by the Bellman equation. We study the asymptotics of this backward stochastic dierential equation (BSDE) to obtain the convergence of
the strategy. This involves nonstandard arguments to deal with nonuniform
quadratic growth in the driver and solutions that are not locally bounded.
To derive the results in the stated generality, it is important to
combine
ideas from optimal control, convex analysis and BSDE theory rather than to
rely on only one of these ingredients; and one may see the problem at hand
as a
model problem of control
in a semimartingale setting.
The chapter is organized as follows. In the next section, we specify the
optimization problem in detail. Section V.3 summarizes the main results on
the risk aversion asymptotics of the optimal strategies and indicates connections to the literature. Section V.4 introduces the main tools, the opportunity process and the Bellman equation, and explains the general approach
for the proofs. In Section V.5 we study the dependence of the opportunity
๐‘ and establish some related estimates. Sections V.6 deals with
๐‘ โ†’ โˆ’โˆž; we prove the main results stated in Section V.3 and, in
process on
the limit
addition, the convergence of the opportunity process and the solution to the
dual problem (in the sense of convex duality). Similarly, Section V.7 contains
the proof of the main theorem for
๐‘โ†’0
and additional renements.
V.2 Preliminaries
If ๐‘ฅ, ๐‘ฆ โˆˆ โ„ are reals, ๐‘ฅ โˆง ๐‘ฆ = min{๐‘ฅ, ๐‘ฆ}
๐‘ฅ โˆจ ๐‘ฆ = max{๐‘ฅ, ๐‘ฆ}. We use 1/0 := โˆž where necessary. If ๐‘ง โˆˆ โ„๐‘‘
๐‘–
โŠค its transpose, and
is a ๐‘‘-dimensional vector, ๐‘ง is its ๐‘–th coordinate, ๐‘ง
โŠค
1/2
๐‘‘
โˆฃ๐‘งโˆฃ = (๐‘ง ๐‘ง)
the Euclidean norm. If ๐‘‹ is an โ„ -valued semimartingale
๐‘‘
and ๐œ‹ is an โ„ -valued predictable integrand, the vector stochastic integral,
โˆซ
denoted by
๐œ‹ ๐‘‘๐‘‹ or ๐œ‹ โˆ™ ๐‘‹ , is a scalar semimartingale with initial value
The following notation is used.
and
zero. Relations between measurable functions hold almost everywhere unless
otherwise mentioned. Dellacherie and Meyer [18] and Jacod and Shiryaev [34]
V.2 Preliminaries
91
are references for unexplained notions from stochastic calculus.
V.2.1 The Optimization Problem
๐‘‡ โˆˆ (0, โˆž) and a ltered probability space
(ฮฉ, โ„ฑ, ๐”ฝ = (โ„ฑ๐‘ก )๐‘กโˆˆ[0,๐‘‡ ] , ๐‘ƒ ) satisfying the usual assumptions of right-continuity
๐‘‘
and completeness, as well as โ„ฑ0 = {โˆ…, ฮฉ} ๐‘ƒ -a.s. Let ๐‘… be an โ„ -valued càdlàg
semimartingale with ๐‘…0 = 0. Its components are interpreted as the returns
1
๐‘‘
of ๐‘‘ risky assets and the stochastic exponential ๐‘† = (โ„ฐ(๐‘… ), . . . , โ„ฐ(๐‘… ))
represents their prices. Let M be the set of equivalent ๐œŽ -martingale measures
for ๐‘† . We assume
M โˆ•= โˆ…,
(2.1)
We consider a xed time horizon
so that arbitrage is excluded in the sense of the NFLVR condition (see Delbaen and Schachermayer [17]).
Our agent also has a bank account at his
disposal. As usual in mathematical nance, the interest rate is assumed to
be zero.
The agent is endowed with a deterministic initial capital
ing strategy
is a predictable
๐‘…-integrable โ„๐‘‘ -valued
๐‘ฅ0 > 0. A trad๐œ‹ , where ๐œ‹ ๐‘– is
process
interpreted as the fraction of the current wealth (or the portfolio proportion)
invested in the
๐‘โ‰ฅ0
such that
๐‘–โˆซth
๐‘‡
0
risky asset. A
consumption rate
๐‘๐‘ก ๐‘‘๐‘ก < โˆž ๐‘ƒ -a.s.
is an optional process
We want to consider two cases simulta-
neously: Either consumption occurs only at the terminal time
๐‘‡
(utility from
terminal wealth only); or there is intermediate and a bulk consumption at
the time horizon. To unify the notation, we dene the measure
{
0
๐œ‡(๐‘‘๐‘ก) :=
๐‘‘๐‘ก
๐œ‡
on
[0, ๐‘‡ ],
in the case without intermediate consumption,
in the case with intermediate consumption.
๐œ‡โˆ˜ := ๐œ‡ + ๐›ฟ{๐‘‡ } , where ๐›ฟ{๐‘‡ } is the unit Dirac measure at ๐‘‡ .
process ๐‘‹(๐œ‹, ๐‘) of a pair (๐œ‹, ๐‘) is dened by the linear equation
Moreover, let
The
wealth
โˆซ
๐‘ก
โˆซ
๐‘‹๐‘ โˆ’ (๐œ‹, ๐‘)๐œ‹๐‘  ๐‘‘๐‘…๐‘  โˆ’
๐‘‹๐‘ก (๐œ‹, ๐‘) = ๐‘ฅ0 +
0
The set of
admissible
๐‘๐‘  ๐œ‡(๐‘‘๐‘ ),
0 โ‰ค ๐‘ก โ‰ค ๐‘‡.
0
trading and consumption pairs is
{
๐’œ(๐‘ฅ0 ) = (๐œ‹, ๐‘) : ๐‘‹(๐œ‹, ๐‘) > 0
The convention
๐‘ก
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘)
and
}
๐‘๐‘‡ = ๐‘‹๐‘‡ (๐œ‹, ๐‘) .
is merely for notational convenience and
๐‘‡ . We x the
๐’œ for ๐’œ(๐‘ฅ0 ). Moreover, ๐‘ โˆˆ ๐’œ indicates
(๐œ‹, ๐‘) โˆˆ ๐’œ; an analogous convention is used for
means that all the remaining wealth is consumed at time
initial capital
๐‘ฅ0
that there exists
and usually write
๐œ‹
such that
similar expressions.
92
V Risk Aversion Asymptotics
It will be convenient to parametrize the consumption strategies as frac-
tions of the wealth. Let
(๐œ‹, ๐‘) โˆˆ ๐’œ and let ๐‘‹ = ๐‘‹(๐œ‹, ๐‘) be the corresponding
wealth process. Then
๐‘
๐‘‹
๐œ… :=
propensity to consume
is called the
corresponding to
propensity to consume is an optional process
๐‘ƒ -a.s.
and
๐œ…๐‘‡ = 1.
The parametrizations by
(๐œ‹, ๐‘).
๐œ… โ‰ฅ 0 such
๐‘ and by ๐œ…
Remark II.2.1) and we abuse the notation by identifying
that
In general, a
โˆซ๐‘‡
๐œ…๐‘  ๐‘‘๐‘  < โˆž
0
are equivalent (see
๐‘
and
๐œ…
when
๐œ‹
is
given. Note that the wealth process can be expressed as
(
)
๐‘‹(๐œ‹, ๐œ…) = ๐‘ฅ0 โ„ฐ ๐œ‹ โˆ™ ๐‘… โˆ’ ๐œ… โˆ™ ๐œ‡ .
(2.2)
The preferences of the agent are modeled by a random utility function with constant relative risk aversion. More precisely, let
adapted positive process and x
๐‘ โˆˆ (โˆ’โˆž, 0) โˆช (0, 1).
๐ท
be a càdlàg
We dene the utility
random eld
(๐‘)
๐‘ˆ๐‘ก (๐‘ฅ) := ๐‘ˆ๐‘ก (๐‘ฅ) := ๐ท๐‘ก ๐‘1 ๐‘ฅ๐‘ ,
๐‘ฅ โˆˆ (0, โˆž), ๐‘ก โˆˆ [0, ๐‘‡ ],
where we assume that there are constants
๐‘˜1 โ‰ค ๐ท๐‘ก โ‰ค ๐‘˜2 ,
The process
๐ท
0 < ๐‘˜1 โ‰ค ๐‘˜2 < โˆž
๐‘
in
such that
0 โ‰ค ๐‘ก โ‰ค ๐‘‡.
is taken to be independent of
in Remark II.2.2. The parameter
(2.3)
๐‘ˆ (๐‘)
๐‘;
(2.4)
interpretations are discussed
will sometimes be suppressed in
the notation and made explicit when we want to recall the dependence. The
same applies to other quantities in this paper.
The constant
expected
utility
[โˆซ
๐ธ
๐‘‡
0
1โˆ’๐‘ > 0
is called the
relative risk aversion
of
๐‘ˆ.
The
corresponding to a consumption rate ๐‘ โˆˆ ๐’œ is given by
]
โˆซ๐‘‡
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) , which is either ๐ธ[๐‘ˆ๐‘‡ (๐‘๐‘‡ )] or ๐ธ[ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐‘‘๐‘ก + ๐‘ˆ๐‘‡ (๐‘๐‘‡ )].
We will always assume that the optimization problem is nondegenerate, i.e.,
๐‘ข๐‘ (๐‘ฅ0 ) := sup ๐ธ
๐‘โˆˆ๐’œ(๐‘ฅ0 )
[โˆซ
0
๐‘‡
]
(๐‘)
๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) < โˆž.
(2.5)
๐‘, but not on ๐‘ฅ0 . Note that
๐‘ข๐‘ (๐‘ฅ0 ) < โˆž for any ๐‘ < ๐‘0 ; and for ๐‘ < 0 the condi(๐‘) < 0. A strategy (๐œ‹, ๐‘) โˆˆ ๐’œ(๐‘ฅ ) is optimal if
tion (2.5) is void since then ๐‘ˆ
0
[โˆซ๐‘‡
]
๐ธ 0 ๐‘ˆ๐‘ก (๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) = ๐‘ข(๐‘ฅ0 ). Note that ๐‘ˆ๐‘ก is irrelevant for ๐‘ก < ๐‘‡ when there
This condition depends on the choice of
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž
implies
is no intermediate consumption. We recall the following existence result.
Proposition 2.1 (Karatzas and ยšitkovi¢ [43]). For each ๐‘, if ๐‘ข๐‘ (๐‘ฅ0 ) < โˆž,
there exists an optimal strategy (ห†๐œ‹, ๐‘ห†) โˆˆ ๐’œ. The corresponding wealth process
ห† = ๐‘‹(ห†
๐‘‹
๐œ‹ , ๐‘ห†) is unique. The consumption rate ๐‘ห† can be chosen to be càdlàg
and is unique ๐‘ƒ โŠ— ๐œ‡โˆ˜ -a.e.
ห† = ๐‘‹(ห†
๐‘ห† denotes this càdlàg version, ๐‘‹
๐œ‹ , ๐‘ห†) is the
ห†
and ๐œ…
ห† = ๐‘ห†/๐‘‹ is the optimal propensity to consume.
In the sequel,
wealth process
optimal
V.2 Preliminaries
93
V.2.2 Decompositions and Spaces of Processes
In some of the statements, we will assume that the price process
alently
that
๐‘…
๐‘…) is continuous.
satises the
๐‘†
(or equiv-
In this case, it follows from (2.1) and Schweizer [71]
structure condition, i.e.,
โˆซ
๐‘‘โŸจ๐‘€ โŸฉ๐œ†,
๐‘…=๐‘€+
๐‘€
Let ๐œ‰
where
is a continuous local martingale with
(2.6)
๐‘€0 = 0
and
๐œ† โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ).
be a scalar special semimartingale, i.e., there exists a (unique)
๐œ‰ = ๐œ‰0 + ๐‘€ ๐œ‰ + ๐ด๐œ‰ , where ๐œ‰0 โˆˆ โ„, ๐‘€ ๐œ‰ is a local
๐œ‰
๐œ‰
๐œ‰
martingale, ๐ด is predictable of nite variation, and ๐‘€0 = ๐ด0 = 0. As ๐‘€ is
๐œ‰
continuous, ๐‘€ has a Kunita-Watanabe (KW) decomposition with respect
canonical decomposition
to
๐‘€,
๐œ‰ = ๐œ‰ 0 + ๐‘ ๐œ‰ โˆ™ ๐‘€ + ๐‘ ๐œ‰ + ๐ด๐œ‰ ,
where
[๐‘€ ๐‘– , ๐‘ ๐œ‰ ] = 0
Stricker [2,
cas
for
1 โ‰ค ๐‘– โ‰ค ๐‘‘
and
(2.7)
๐‘ ๐œ‰ โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ );
see Ansel and
3]. Analogous notation will be used for other special semi-
martingales and, with a slight abuse of terminology, we will refer to (2.7) as
the KW decomposition of
๐‘ƒ -semimartingales and ๐‘Ÿ โˆˆ [1, โˆž). If
๐‘‹ = ๐‘‹0 + ๐‘€ ๐‘‹ + ๐ด๐‘‹ , we dene
โˆซ ๐‘‡
1/2 โˆฅ๐‘‹โˆฅโ„‹๐‘Ÿ := โˆฃ๐‘‹0 โˆฃ + 0 โˆฃ๐‘‘๐ด๐‘‹ โˆฃ๐ฟ๐‘Ÿ + [๐‘€ ๐‘‹ ]๐‘‡ ๐ฟ๐‘Ÿ .
[
]
2
In particular, we will often use that โˆฅ๐‘ โˆฅ 2 = ๐ธ [๐‘ ]๐‘‡ for a local martingale
โ„‹
๐‘ with ๐‘0 = 0. If ๐‘‹ is a non-special semimartingale, โˆฅ๐‘‹โˆฅโ„‹๐‘Ÿ := โˆž. We
๐‘Ÿ
can now dene โ„‹ := {๐‘‹ โˆˆ ๐’ฎ : โˆฅ๐‘‹โˆฅโ„‹๐‘Ÿ < โˆž}. The same space is some๐‘Ÿ
times denoted by ๐’ฎ in the literature; moreover, there are many equivalent
๐‘Ÿ
๐‘Ÿ
denitions for โ„‹ (see [18, VII.98]). The localized spaces โ„‹๐‘™๐‘œ๐‘ are dened
๐‘›
๐‘›
๐‘Ÿ
in the usual way. In particular, if ๐‘‹, ๐‘‹ โˆˆ ๐’ฎ we say that ๐‘‹ โ†’ ๐‘‹ in โ„‹๐‘™๐‘œ๐‘
if there exists a localizing sequence of stopping times (๐œ๐‘š )๐‘šโ‰ฅ1 such that
lim๐‘› โˆฅ(๐‘‹ ๐‘› โˆ’ ๐‘‹)๐œ๐‘š โˆฅโ„‹๐‘Ÿ = 0 for all ๐‘š. The localizing sequence may depend
๐‘›
on the sequence (๐‘‹ ), causing this convergence to be non-metrizable. On
๐’ฎ , the Émery distance is dened by
[
]
โˆ™
๐‘‘(๐‘‹, ๐‘Œ ) := โˆฃ๐‘‹0 โˆ’ ๐‘Œ0 โˆฃ + sup ๐ธ sup 1 โˆง โˆฃ๐ป (๐‘‹ โˆ’ ๐‘Œ )๐‘ก โˆฃ ,
Let
๐‘‹โˆˆ๐’ฎ
๐’ฎ
๐œ‰.
be the space of all càdlàg
has the canonical decomposition
โˆฃ๐ปโˆฃโ‰ค1
๐‘กโˆˆ[0,๐‘‡ ]
where the supremum is taken over all predictable processes bounded by one
in absolute value.
topology
This complete metric induces on
(cf. Émery [20]).
An optional process
๐‘‹
satises a certain property
satises this property for
local simply means local.
the
semimartingale
prelocally
if there ex-
๐œ๐‘š such that ๐‘‹ ๐œ๐‘š โˆ’ := ๐‘‹1[0,๐œ๐‘š ) +
each ๐‘š. When ๐‘‹ is continuous, pre-
ists a localizing sequence of stopping times
๐‘‹๐œ๐‘š โˆ’ 1[๐œ๐‘š ,๐‘‡ ]
๐’ฎ
94
V Risk Aversion Asymptotics
Proposition 2.2 ([20]). Let ๐‘‹, ๐‘‹ ๐‘› โˆˆ ๐’ฎ and ๐‘Ÿ โˆˆ [1, โˆž). Then ๐‘‹ ๐‘› โ†’ ๐‘‹ in
the semimartingale topology if and only if every subsequence of (๐‘‹ ๐‘› ) has a
subsequence which converges to ๐‘‹ prelocally in โ„‹๐‘Ÿ .
We denote by
๐ต๐‘€ ๐‘‚
โˆฅ๐‘ โˆฅ2๐ต๐‘€ ๐‘‚
where
๐œ
norm).
๐‘ with ๐‘0 = 0
]
[
:= sup ๐ธ [๐‘ ]๐‘‡ โˆ’ [๐‘ ]๐œ โˆ’ โ„ฑ๐œ โˆž < โˆž,
the space of martingales
satisfying
๐ฟ
๐œ
๐ต๐‘€ ๐‘‚2 โ„‹๐œ” be the
๐‘‹0 = 0 and
ranges over all stopping times (more precisely, this is the
There exists a similar notion for semimartingales:
let
โ„‹1 consisting of all special semimartingales ๐‘‹ with
]
[(
)1/2 โˆซ ๐‘‡
:= sup ๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐œ โˆ’
+ ๐œ โˆ’ โˆฃ๐‘‘๐ด๐‘‹ โˆฃ โ„ฑ๐œ subspace of
โˆฅ๐‘‹โˆฅ2โ„‹๐œ”
Finally, let
๐ฟโˆž
๐œ
โ„›๐‘Ÿ
< โˆž.
be the space of scalar adapted processes which are right-
continuous and such that
โˆฅ๐‘‹โˆฅโ„›๐‘Ÿ := sup โˆฃ๐‘‹๐‘ก โˆฃ
0โ‰ค๐‘กโ‰ค๐‘‡
๐ฟ๐‘Ÿ
< โˆž.
With a mild abuse of notation, we will use the same norm also for leftcontinuous processes.
V.3 Main Results
In this section we present the main results about the limits of the optimal
strategies. To state an assumption in the results, we rst recall the
nity process ๐ฟ(๐‘).
opportu-
๐‘ข๐‘ (๐‘ฅ0 ) < โˆž. Then by Proposition
there exists a unique càdlàg semimartingale ๐ฟ(๐‘) such that
]
[โˆซ ๐‘‡
(
)๐‘
1
๐ฟ๐‘ก (๐‘) ๐‘ ๐‘‹๐‘ก (๐œ‹, ๐‘) = ess sup ๐ธ
๐‘ˆ๐‘  (หœ
๐‘๐‘  ) ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก , 0 โ‰ค ๐‘ก โ‰ค ๐‘‡
Fix
๐‘
such that
๐‘หœโˆˆ๐’œ(๐œ‹,๐‘,๐‘ก)
for all
(๐œ‹, ๐‘) โˆˆ ๐’œ,
where
II.3.1
(3.1)
๐‘ก
{
๐’œ(๐œ‹, ๐‘, ๐‘ก) := (หœ
๐œ‹ , ๐‘หœ) โˆˆ ๐’œ : (หœ
๐œ‹ , ๐‘หœ) = (๐œ‹, ๐‘)
on
}
[0, ๐‘ก] .
We can now proceed to state the main results. The proofs are postponed
to Sections V.6 and V.7. Those sections also contain statements about the
convergence of the opportunity processes and the solutions to the dual problems, as well as some renements of the results below.
V.3.1 The Limit ๐‘ โ†’ โˆ’โˆž
The relative risk aversion
1โˆ’๐‘
of
๐‘ˆ (๐‘)
increases to innity as
๐‘ โ†’ โˆ’โˆž.
Therefore we expect that in the limit, the agent does not invest at all. In
that situation the optimal propensity to consume is
๐œ…๐‘ก = (1 + ๐‘‡ โˆ’ ๐‘ก)โˆ’1
since
this corresponds to a constant consumption rate. Our rst result shows that
this coincides with the limit of the
๐‘ˆ (๐‘) -optimal
propensities to consume.
V.3 Main Results
95
Theorem 3.1. The following convergences hold as ๐‘ โ†’ โˆ’โˆž.
(i) Let ๐‘ก โˆˆ [0, ๐‘‡ ]. In the case with intermediate consumption,
1
1+๐‘‡ โˆ’๐‘ก
๐œ…
ห† ๐‘ก (๐‘) โ†’
๐‘ƒ -a.s.
If ๐”ฝ is continuous, the convergence is uniform in ๐‘ก, ๐‘ƒ -a.s.; and holds
also in โ„›๐‘Ÿ๐‘™๐‘œ๐‘ for all ๐‘Ÿ โˆˆ [1, โˆž).
(ii) If ๐‘† is continuous and ๐ฟ(๐‘) is continuous for all ๐‘ < 0, then
ห†
and ๐‘‹(๐‘)
โ†’ ๐‘ฅ0 exp โˆ’
(
๐œ‹
ห† (๐‘) โ†’ 0
โˆซ โ‹… ๐œ‡(๐‘‘๐‘ ) )
0 1+๐‘‡ โˆ’๐‘ 
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ )
in the semimartingale topology.
The continuity assumptions in (ii) are always satised if the ltration
๐”ฝ
is generated by a Brownian motion; see also Remark 4.2.
Literature.
We are not aware of a similar result in the continuous-time litera-
ture, with the exception that when the strategies can be calculated explicitly,
the convergences mentioned in this section are often straightforward to obtain. E.g., Grasselli [31] carries out such a construction in a complete market
model. There are also related systematic results. Carassus and Rásonyi [10]
and Grandits and Summer [30] study convergence to the superreplication
problem for increasing (absolute) risk aversion of general utility functions
in discrete models. Note that superreplicating the contingent claim
๐ต โ‰ก0
corresponds to not trading at all. For the maximization of exponential utility
โˆ’ exp(โˆ’๐›ผ๐‘ฅ)
without claim, the optimal strategy is proportional to the
inverse of the absolute risk aversion
in the limit
๐›ผ โ†’ โˆž.
๐›ผ
and hence trivially converges to zero
The case with claim has also been studied. See, e.g.,
Mania and Schweizer [53] for a continuous model, and Becherer [4] for a
related result.
The references given here and later in this section do not
consider intermediate consumption.
We continue with our second main result, which concerns only the case
without intermediate consumption.
We rst introduce in detail the expo-
nential hedging problem already mentioned above. Let
contingent claim.
utility (here with
๐ต โˆˆ ๐ฟโˆž (โ„ฑ๐‘‡ )
Then the aim is to maximize the expected exponential
๐›ผ = 1)
of the terminal wealth including the claim,
)]
[
(
max ๐ธ โˆ’ exp ๐ต โˆ’ ๐‘ฅ0 โˆ’ (๐œ— โˆ™ ๐‘…)๐‘‡ ,
(3.2)
๐œ—โˆˆฮ˜
where
๐œ—
be a
is the trading strategy parametrized by the
vested in the assets (setting
๐‘–
๐œ— :=
corresponds to the more customary
To describe the set
ฮ˜,
monetary amounts
๐‘– yields
1{๐‘† ๐‘– โˆ•=0} ๐œ—๐‘– /๐‘†โˆ’
โˆ’
๐œ—
โˆ™
๐‘† =๐œ—
[ ๐‘‘๐‘„
๐‘‘๐‘ƒ
log
๐‘…
and
number of shares of the assets).
entropy of ๐‘„ โˆˆ M relative to ๐‘ƒ
we dene the
๐ป(๐‘„โˆฃ๐‘ƒ ) := ๐ธ
โˆ™
( ๐‘‘๐‘„ )]
๐‘‘๐‘ƒ
[
( ๐‘‘๐‘„ )]
= ๐ธ ๐‘„ log
๐‘‘๐‘ƒ
in-
by
V Risk Aversion Asymptotics
96
and let
{
}
M ๐‘’๐‘›๐‘ก = ๐‘„ โˆˆ M : ๐ป(๐‘„โˆฃ๐‘ƒ ) < โˆž .
We assume in the following that
M ๐‘’๐‘›๐‘ก โˆ•= โˆ….
(3.3)
}
๐‘„-supermartingale for all ๐‘„ โˆˆ M ๐‘’๐‘›๐‘ก is
of admissible strategies for (3.2). If ๐‘† is locally bounded, there
ห† โˆˆ ฮ˜ for (3.2) by Kabanov and Stricker [37,
optimal strategy ๐œ—
{
ฮ˜ := ๐œ— โˆˆ ๐ฟ(๐‘…) : ๐œ—
Now
the class
exists an
โˆ™
๐‘…
is a
Theorem 2.1]. (See Biagini and Frittelli [7, 8] for the unbounded case.)
As there is no intermediate consumption, the process
to a random variable
๐ท๐‘‡ โˆˆ ๐ฟโˆž (โ„ฑ๐‘‡ ).
๐ท
in (2.3) reduces
If we choose
๐ต := log(๐ท๐‘‡ ),
(3.4)
we have the following result.
Theorem 3.2. Let ๐‘† be continuous and assume that ๐ฟ(๐‘) is continuous for
all ๐‘ < 0. Under (3.3) and (3.4),
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ).
(1 โˆ’ ๐‘) ๐œ‹
ห† (๐‘) โ†’ ๐œ—ห†
Here ๐œ‹ห† (๐‘) is in the fractions of wealth parametrization, while ๐œ—ห† denotes the
monetary amounts invested for the exponential utility.
As this convergence may seem surprising at rst glance, we give the
following heuristics.
Remark 3.3.
๐ต = log(๐ท๐‘‡ ) = 0 for simplicity. The preferences
=
โ„+ are not directly comparable to the ones
exponential utility, which are dened on โ„. We consider the
Assume
(๐‘) (๐‘ฅ)
induced by ๐‘ˆ
given by the
1 ๐‘
๐‘ ๐‘ฅ on
shifted power utility functions
(
)
หœ (๐‘) (๐‘ฅ) := ๐‘ˆ (๐‘) ๐‘ฅ + 1 โˆ’ ๐‘ ,
๐‘ˆ
Then
หœ (๐‘)
๐‘ˆ
again has relative risk aversion
denition increases to
โ„
as
หœ (๐‘) (๐‘ฅ) =
(1 โˆ’ ๐‘)1โˆ’๐‘ ๐‘ˆ
๐‘ โ†’ โˆ’โˆž.
( ๐‘ฅ
1โˆ’๐‘
๐‘
1โˆ’๐‘
๐‘ฅ โˆˆ (๐‘ โˆ’ 1, โˆž).
1โˆ’๐‘ > 0
and its domain of
Moreover,
+1
)๐‘
โ†’ โˆ’๐‘’โˆ’๐‘ฅ ,
๐‘ โ†’ โˆ’โˆž,
(3.5)
and the multiplicative constant does not aect the preferences.
Let the agent with utility function
โˆ—
capital ๐‘ฅ0
หœ (๐‘) be
๐‘ˆ
๐‘ฅโˆ—0 < 0,
endowed with some initial
โˆˆ โ„ independent of ๐‘. (If
we consider only values of
๐‘ such that ๐‘ โˆ’ 1 < ๐‘ฅโˆ—0 .) The change of variables ๐‘ฅ = ๐‘ฅ
หœ + 1 โˆ’ ๐‘ yields
(๐‘)
(๐‘)
หœ
ห†
๐‘ˆ (๐‘ฅ) = ๐‘ˆ (หœ
๐‘ฅ). Hence the corresponding optimal wealth processes ๐‘‹(๐‘)
หœ
หœ
ห†
and ๐‘‹(๐‘) are related by ๐‘‹(๐‘) = ๐‘‹(๐‘) โˆ’ 1 + ๐‘ if we choose the initial capital
๐‘ฅ0 := ๐‘ฅโˆ—0 + 1 โˆ’ ๐‘ > 0 for the agent with ๐‘ˆ (๐‘) . We conclude
(
)
หœ
ห†
ห† ๐œ‹ (๐‘) ๐‘‘๐‘… = ๐‘‹(๐‘)
หœ
๐‘‘๐‘‹(๐‘)
= ๐‘‘๐‘‹(๐‘)
= ๐‘‹(๐‘)ห†
+1โˆ’๐‘ ๐œ‹
ห† (๐‘) ๐‘‘๐‘…,
V.3 Main Results
i.e., the optimal monetary investment
หœ
๐œ—(๐‘)
for
หœ (๐‘)
๐‘ˆ
97
is given by
)
(
หœ
หœ
+1โˆ’๐‘ ๐œ‹
ห† (๐‘).
๐œ—(๐‘)
= ๐‘‹(๐‘)
In view of (3.5), it is reasonable that
หœ
๐œ—(๐‘)
monetary investment for the exponential utility.
fractions of wealth) does not depend on
conditions of Theorem 3.1.
๐œ—ห†,
should converge to
๐‘ฅ0
the optimal
We recall that
๐œ‹
ห† (๐‘)
(in
and converges to zero under the
Thus, loosely speaking,
หœ ๐œ‹ (๐‘) โ‰ˆ 0
๐‘‹(๐‘)ห†
for
โˆ’๐‘
large, and hence
หœ
๐œ—(๐‘)
โ‰ˆ (1 โˆ’ ๐‘)ห†
๐œ‹ (๐‘).
)
(
หœ ๐œ‹ (๐‘)
show that lim๐‘โ†’โˆ’โˆž ๐‘‹(๐‘)ห†
More precisely, one can
โˆ™
๐‘… = 0
in the
semimartingale topology, using arguments as in Appendix V.8.
Literature.
To the best of our knowledge, the statement of Theorem 3.2
is new in the systematic literature.
the dual side for the case
๐ต = 0.
However, there are known results on
The problem dual to exponential utility
maximization is the minimization of
๐‘„๐ธ
โˆˆ
๐ป(๐‘„โˆฃ๐‘ƒ )
over
M ๐‘’๐‘›๐‘ก
and the optimal
M ๐‘’๐‘›๐‘ก is called minimal entropy martingale measure.
tional assumptions on the model, the solution
๐‘Œห† (๐‘)
Under addi-
of the dual problem for
๐‘„๐‘ž
ห†
ห†
= ๐‘Œ๐‘‡ (๐‘)/๐‘Œ0 (๐‘) is called ๐‘ž -optimal martingale measure,
where ๐‘ž < 1 is conjugate to ๐‘. This measure can be dened also for ๐‘ž > 1,
๐‘ž
in which case it is not connected to power utility. The convergence of ๐‘„ to
๐ธ
๐‘„ for ๐‘ž โ†’ 1+ was proved by Grandits and Rheinländer [29] for continuous
power utility (4.3) introduced below is a martingale and then the measure
๐‘ž
dened by ๐‘‘๐‘„ /๐‘‘๐‘ƒ
semimartingale models satisfying a reverse Hölder inequality. Under the additional assumption that
๐‘žโ†’1
๐”ฝ
is continuous, the convergence of
and more generally the continuity of
๐‘„๐‘ž for
๐‘ž 7โ†’
๐‘žโ‰ฅ0
๐‘„๐‘ž
to
๐‘„๐ธ
for
were obtained
by Mania and Tevzadze [54] (see also Santacroce [67]) using BSDE convergence together with
๐ต๐‘€ ๐‘‚
arguments.
The latter are possible due to the
reverse Hölder inequality; an assumption which is not present in our results.
V.3.2 The Limit ๐‘ โ†’ 0
As
๐‘
tends to zero, the relative risk aversion of the power utility tends to
log(๐‘ฅ).
which corresponds to the utility function
๐‘ขlog (๐‘ฅ0 ) := sup ๐ธ
โˆ’โˆž
]
log(๐‘๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) ;
0
๐‘โˆˆ๐’œ(๐‘ฅ0 )
here integrals are set to
๐‘‡
[โˆซ
1,
Hence we consider
if they are not well dened in
โ„.
A
log-utility
agent exhibits a very special (myopic) behavior, which allows for an explicit
solution of the utility maximization problem (cf. Goll and Kallsen [24, 25]).
If in particular
๐‘†
is continuous, the
๐œ‹๐‘ก = ๐œ†๐‘ก ,
log-optimal
๐œ…๐‘ก =
strategy is
1
1+๐‘‡ โˆ’๐‘ก
V Risk Aversion Asymptotics
98
by [24, Theorem 3.1], where
๐œ†
is dened by (2.6). Our result below shows
๐ท โ‰ก 1 converges to the logoptimal one as ๐‘ โ†’ 0. In general, the randomness of ๐ท is an additional source
that the optimal strategy for power utility with
of risk and will cause an excess hedging demand.
Consider the bounded
semimartingale
๐œ‚๐‘ก := ๐ธ
๐‘‡
[โˆซ
๐‘ก
]
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก .
๐œ‚ = ๐œ‚0 + ๐‘ ๐œ‚ โˆ™ ๐‘€ + ๐‘ ๐œ‚ + ๐ด๐œ‚ denotes the Kunita-Watanabe
decomposition of ๐œ‚ with respect to ๐‘€ and the standard case ๐ท โ‰ก 1 correโˆ˜
๐œ‚
sponds to ๐œ‚๐‘ก = ๐œ‡ [๐‘ก, ๐‘‡ ] and ๐‘ = 0.
If
๐‘†
is continuous,
Theorem 3.4. Assume ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for some ๐‘0 โˆˆ (0, 1). As ๐‘ โ†’ 0,
(i) in the case with intermediate consumption,
๐œ…
ห† ๐‘ก (๐‘) โ†’
๐ท๐‘ก
๐œ‚๐‘ก
uniformly in ๐‘ก, ๐‘ƒ -a.s.
(ii) if ๐‘† is continuous,
๐œ‹
ห† (๐‘) โ†’ ๐œ† +
๐‘๐œ‚
๐œ‚โˆ’
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ )
and the corresponding wealth processes converge in the semimartingale
topology.
Remark 3.5. If we consider the limit ๐‘ โ†’ 0โˆ’, we need not a priori assume
that
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž
for some
๐‘ 0 > 0.
Without that condition, the assertions
of Theorem 3.4 remain valid if (i) is replaced by the weaker statement that
lim๐‘โ†’0โˆ’ ๐œ…
ห† ๐‘ก (๐‘) โ†’ ๐ท๐‘ก /๐œ‚๐‘ก ๐‘ƒ -a.s.
for all
๐‘ก.
If
๐”ฝ
is continuous, (i) remains valid
without changes. In particular, these convergences hold even if
Literature.
In the following discussion we assume
is part of the folklore that the
๐œ‹
ห† (๐‘)
by formally setting
log-optimal
๐‘ = 0.
๐ท โ‰ก1
๐‘ขlog (๐‘ฅ0 ) = โˆž.
for simplicity. It
strategy can be obtained from
Initiated by Jouini and Napp [36], a re-
cent branch of the literature studies the stability of the utility maximization
problem under perturbations of the utility function (with respect to pointwise convergence) and other ingredients of the problem. To the best of our
knowledge, intermediate consumption was not considered so far and the previous results for continuous time concern continuous semimartingale models.
We note that
log(๐‘ฅ) = lim๐‘โ†’0 (๐‘ˆ (๐‘) (๐‘ฅ) โˆ’ ๐‘โˆ’1 )
and here the additive con-
stant does not inuence the optimal strategy, i.e., we have pointwise conver-
๐‘ˆ (๐‘) . Now Larsen [51, Theorem 2.2]
ห†๐‘‡ for ๐‘ˆ (๐‘) converges in probabilwealth ๐‘‹
gence of utility functions equivalent to
implies that the optimal terminal
ity to the
log-optimal one and that the value functions at time zero converge
pointwise (in the continuous case without consumption). We use the specic
form of our utility functions and obtain a stronger result. Finally, we can
mention that on the dual side and for
the continuity of
๐‘ž -optimal
๐‘ โ†’ 0โˆ’,
the convergence is related to
measures as mentioned after Remark 3.3.
V.4 Tools and Ideas for the Proofs
For general
of
๐ท
๐ท
๐‘,
and
it seems dicult to determine the precise inuence
๐œ‹
ห† (๐‘).
on the optimal trading strategy
We can read Theorem 3.4(ii) as
a partial result on the excess hedging demand
๐œ‹
ห† (๐‘, 1)
99
๐œ‹
ห† (๐‘) โˆ’ ๐œ‹
ห† (๐‘, 1)
๐ท โ‰ก 1.
due to
๐ท;
here
denotes the optimal strategy for the case
Corollary 3.6. Suppose that the conditions of Theorem 3.4(ii) hold. Then
๐œ‹
ห† (๐‘) โˆ’ ๐œ‹
ห† (๐‘, 1) โ†’ ๐‘ ๐œ‚ /๐œ‚โˆ’ in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) as ๐‘ โ†’ 0; i.e., the asymptotic excess
hedging demand due to ๐ท is given by ๐‘ ๐œ‚ /๐œ‚โˆ’ .
The stability theory mentioned above considers also perturbations of the
probability measure
๐‘ƒ
(see Kardaras and ยšitkovi¢ [45]) and our corollary
can be related as follows. In the special case when
under
๐‘ƒ
๐ท
๐‘ˆ (๐‘)
is a martingale,
corresponds to the standard power utility function optimized under
the measure
demand due
๐‘‘๐‘ƒหœ = (๐ท๐‘‡ /๐ท0 ) ๐‘‘๐‘ƒ (see Remark II.2.2). The excess hedging
หœ.
to ๐ท then represents the inuence of the subjective beliefs ๐‘ƒ
V.4 Tools and Ideas for the Proofs
In this section we introduce our main tools and then present the basic ideas
how to apply them for the proofs of the theorems.
V.4.1 Opportunity Processes
We x
๐‘
and assume
๐‘ข๐‘ (๐‘ฅ0 ) < โˆž
throughout this section. We rst discuss
๐ฟ = ๐ฟ(๐‘) as introduced
๐ฟ๐‘‡ = ๐ท๐‘‡ and that
๐‘ข๐‘ (๐‘ฅ0 ) = ๐ฟ0 ๐‘1 ๐‘ฅ๐‘0 is the value function from (2.5). Moreover, ๐ฟ has the following properties by Lemma II.3.5 and the bounds (2.4) for ๐ท .
the properties of the (primal) opportunity process
in (3.1).
Directly from that equation we have that
Lemma 4.1. The opportunity process satises ๐ฟ, ๐ฟโˆ’ > 0.
(i) If ๐‘ โˆˆ (0, 1), ๐ฟ is a supermartingale satisfying
)โˆ’๐‘ [
๐ฟ๐‘ก โ‰ฅ ๐œ‡ [๐‘ก, ๐‘‡ ]
๐ธ
(
โˆ˜
โˆซ
๐‘ก
๐‘‡
]
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก โ‰ฅ ๐‘˜1 .
(ii) If ๐‘ < 0, ๐ฟ is a bounded semimartingale satisfying
)โˆ’๐‘ [
๐ธ
0 < ๐ฟ๐‘ก โ‰ค ๐œ‡ [๐‘ก, ๐‘‡ ]
(
โˆ˜
โˆซ
๐‘ก
๐‘‡
]
(
)1โˆ’๐‘
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก โ‰ค ๐‘˜2 ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]
.
If in addition there is no intermediate consumption, then ๐ฟ is a submartingale.
In particular,
๐ฟ
๐›ฝ :=
is always a special semimartingale. We denote by
1
> 0,
1โˆ’๐‘
๐‘ž :=
๐‘
โˆˆ (โˆ’โˆž, 0) โˆช (0, 1)
๐‘โˆ’1
(4.1)
V Risk Aversion Asymptotics
100
the relative risk tolerance and the exponent conjugate to
These constants are of course redundant given
๐‘,
๐‘,
respectively.
but turn out to simplify
the notation.
In the case with intermediate consumption, the opportunity process and
the optimal consumption are related by
๐‘ห†๐‘ก =
( ๐ท )๐›ฝ
๐‘ก
๐ฟ๐‘ก
ห†๐‘ก
๐‘‹
and hence
๐œ…
ห†๐‘ก =
( ๐ท )๐›ฝ
๐‘ก
(4.2)
๐ฟ๐‘ก
according to Theorem II.5.1. Next, we introduce the convex-dual analogue
of
๐ฟ;
cf. Section II.4 for the following notions and results. The
dual problem
is
inf ๐ธ
๐‘Œ โˆˆY
[โˆซ
๐‘‡
]
๐‘ˆ๐‘กโˆ— (๐‘Œ๐‘ก ) ๐œ‡โˆ˜ (๐‘‘๐‘ก) ,
(4.3)
0
{
}
๐‘ˆ๐‘กโˆ— (๐‘ฆ) = sup๐‘ฅ>0 ๐‘ˆ๐‘ก (๐‘ฅ) โˆ’ ๐‘ฅ๐‘ฆ = โˆ’ 1๐‘ž ๐‘ฆ ๐‘ž ๐ท๐‘ก๐›ฝ is the conjugate of ๐‘ˆ๐‘ก .
Only three properties of the domain Y = Y (๐‘) are relevant for us. First,
each element ๐‘Œ โˆˆ Y is a positive càdlàg supermartingale. Second, the set Y
๐‘โˆ’1
,
depends on ๐‘ only by a normalization: with the constant ๐‘ฆ0 (๐‘) := ๐ฟ0 (๐‘)๐‘ฅ0
โ€ฒ
โˆ’1
the set Y := ๐‘ฆ0 (๐‘)
Y (๐‘) does not depend on ๐‘. As the elements of Y will
where
occur only in terms of certain fractions, the constant plays no role. Third,
the
๐‘„ โˆˆ M is contained in Y (modulo scaling).
โˆ—
The dual opportunity process ๐ฟ is the analogue of ๐ฟ for the dual problem
๐‘ƒ -density
process of any
and can be dened by
โŽง
]
[โˆซ
๏ฃด
โŽจess sup๐‘Œ โˆˆY ๐ธ ๐‘‡ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘ก
๐ฟโˆ—๐‘ก :=
]
[โˆซ
๏ฃด
๐‘‡
โŽฉ
ess inf ๐‘Œ โˆˆY ๐ธ ๐‘ก ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
Here the extremum is attained at the minimizer
denote by
๐‘Œห† = ๐‘Œห† (๐‘).
if
๐‘<0,
if
๐‘ โˆˆ (0, 1).
(4.4)
๐‘Œ โˆˆY
for (4.3), which we
Finally, we shall use that the primal and the dual
opportunity process are related by the power
๐ฟโˆ— = ๐ฟ๐›ฝ .
(4.5)
V.4.2 Bellman BSDE
We continue with a xed
๐‘ such that ๐‘ข๐‘ (๐‘ฅ0 ) < โˆž.
We recall (Chapter III) the
Bellman BSDE, which in the present chapter will be used only for continuous
๐‘†.
๐ฟ = ๐ฟ0 + ๐‘ ๐ฟ โˆ™ ๐‘€ + ๐‘ ๐ฟ + ๐ด๐ฟ be the
๐ฟ
๐ฟ
respect to ๐‘€ . Then the triplet (๐ฟ, ๐‘ , ๐‘ )
In this case, recall (2.6) and let
KW decomposition
1
1 of
๐ฟ
In this chapter, we write
with
๐‘๐ฟ
line with the BSDE literature.
instead of
๐œ‘๐ฟ
(as in Chapter III), since this is more in
V.4 Tools and Ideas for the Proofs
101
satises the Bellman BSDE
๐‘‘๐ฟ๐‘ก =
(
(
๐‘ž
๐‘๐ฟ )
๐‘ ๐ฟ )โŠค
๐‘‘โŸจ๐‘€ โŸฉ๐‘ก ๐œ†๐‘ก + ๐‘ก
โˆ’ ๐‘๐‘ˆ๐‘กโˆ— (๐ฟ๐‘กโˆ’ ) ๐œ‡(๐‘‘๐‘ก)
๐ฟ๐‘กโˆ’ ๐œ†๐‘ก + ๐‘ก
2
๐ฟ๐‘กโˆ’
๐ฟ๐‘กโˆ’
+ ๐‘๐‘ก๐ฟ ๐‘‘๐‘€๐‘ก + ๐‘‘๐‘๐‘ก๐ฟ ;
(4.6)
๐ฟ๐‘‡ = ๐ท๐‘‡ .
Put dierently, the nite variation part of
โˆซ
๐‘ž
=
2
๐ด๐ฟ
๐‘ก
๐‘ก
๐ฟ๐‘ โˆ’
(
0
๐ฟ
satises
(
๐‘ ๐ฟ )โŠค
๐‘๐ฟ )
๐œ†๐‘  + ๐‘ 
๐‘‘โŸจ๐‘€ โŸฉ๐‘  ๐œ†๐‘  + ๐‘ 
โˆ’๐‘
๐ฟ๐‘ โˆ’
๐ฟ๐‘ โˆ’
๐‘ก
โˆซ
๐‘ˆ๐‘ โˆ— (๐ฟ๐‘ โˆ’ ) ๐œ‡(๐‘‘๐‘ ).
0
(4.7)
โˆ—
Here ๐‘ˆ is dened as after (4.3). Moreover, the optimal trading strategy
๐œ‹
ห†
can be described by
(
๐‘๐ฟ )
๐œ‹
ห†๐‘ก = ๐›ฝ ๐œ†๐‘ก + ๐‘ก .
๐ฟ๐‘กโˆ’
(4.8)
See Corollary III.3.12 for these results. Finally, still under the assumption
of continuity, the solution to the dual problem (4.3) is given by the local
martingale
(
1
๐‘Œห† = ๐‘ฆ0 โ„ฐ โˆ’ ๐œ† โˆ™ ๐‘€ +
๐ฟโˆ’
with the constant
Remark 4.2.
๐‘ฆ0 = ๐‘ขโ€ฒ๐‘ (๐‘ฅ0 ) = ๐ฟ0 ๐‘ฅ๐‘โˆ’1
0
Continuity of
local martingale
๐‘๐ฟ
๐‘†
โˆ™
)
๐‘๐ฟ ,
(4.9)
(cf. Remark III.5.18).
does not imply that
๐ฟ
is continuous; the
may still have jumps (see also Remark III.3.13(i)). If
the ltration
๐”ฝ is continuous (i.e., all ๐”ฝ-martingales are continuous), it clearly
follows that ๐ฟ and ๐‘† are continuous. The most important example with this
property is the Brownian ltration.
V.4.3 The Strategy for the Proofs
We can now summarize the basic scheme that is common for the proofs of
the three theorems.
The rst step is to prove the
process
๐ฟ
pointwise convergence
or of the dual opportunity process
๐ฟโˆ— ;
of the opportunity
the choice of the process
depends on the theorem. The convergence of the optimal propensity to consume
of
๐ฟ
๐œ…
ห†
then follows in view of the feedback formula (4.2). The denitions
and
๐ฟโˆ—
via the value processes lend themselves to control-theoretic ar-
guments, and of course Jensen's inequality will be the basic tool to derive
๐ฟโˆ— = ๐ฟ๐›ฝ from (4.5), it is essentially equivaโˆ—
lent whether one works with ๐ฟ or ๐ฟ , as long as ๐‘ is xed. However, the dual
estimates. In view of the relation
problem has the advantage of being dened over a set of supermartingales,
which are easier to handle than consumption and wealth processes. This is
particularly useful when passing to the limit.
V Risk Aversion Asymptotics
102
The second step is the convergence of the trading strategy
its formula (4.8) contains the integrand
of
๐ฟ
๐‘€.
with respect to
๐‘๐ฟ
๐œ‹
ห†.
Note that
from the KW decomposition
Therefore, the convergence of
convergence of the martingale part ๐‘€ ๐ฟ (resp. ๐‘€ ๐ฟ
โˆ—
๐œ‹
ห†
is related to the
). In general, the pointwise
convergence of a semimartingale is not enough to deduce the convergence
of its martingale part; this requires some control over the semimartingale
decomposition. In our case, this control is given by the Bellman BSDE (4.6),
which can be seen as a description for the dependence of the nite variation
part
๐ด๐ฟ
on the martingale part
๐‘€ ๐ฟ.
As we use the BSDE to show the
๐ฟ
convergence of ๐‘€ , we benet from techniques from the theory of quadratic
BSDEs. However, we cannot apply standard results from that theory since
our assumptions are not strong enough.
In general, our approach is to extract as much information as possible
by basic control arguments and convex analysis
before
tackling the BSDE,
rather than to rely exclusively on (typically delicate) BSDE arguments. For
instance, we use the BSDE only after establishing the pointwise convergence
of its left hand side, i.e., the opportunity process. This essentially eliminates
the need for an
a priori
estimate or a comparison principle and constitutes
a key reason for the generality of our results.
Our procedure shares basic
features of the viscosity approach to Markovian control problems, where one
also works directly with the value function before tackling the HamiltonJacobi-Bellman equation.
V.5 Auxiliary Results
We start by collecting inequalities for the dependence of the opportunity
processes on
๐‘.
The precise formulations are motivated by the applications
in the proofs of the previous theorems, but the comparison results are also
of independent interest.
V.5.1 Comparison Results
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for a given exponent ๐‘0 .
For convenience, we recall the quantities ๐›ฝ = 1/(1 โˆ’ ๐‘) > 0 and ๐‘ž = ๐‘/(๐‘ โˆ’ 1)
dened in (4.1). It is useful to note that ๐‘ž โˆˆ (โˆ’โˆž, 0) for ๐‘ โˆˆ (0, 1) and vice
versa. When there is a second exponent ๐‘0 under consideration, ๐›ฝ0 and ๐‘ž0
have the obvious denition. We also recall from (2.4) the bounds ๐‘˜1 and ๐‘˜2
for ๐ท .
We assume in the entire section that
Proposition 5.1. Let 0 < ๐‘ < ๐‘0 < 1. For each ๐‘ก โˆˆ [0, ๐‘‡ ],
๐ฟโˆ—๐‘ก (๐‘)
โ‰ค ๐ธ
๐ฟ๐‘ก (๐‘) โ‰ค
(
[โˆซ
๐‘ก
โˆ˜
๐‘‡
]1โˆ’๐‘ž/๐‘ž0 (
)๐‘ž/๐‘ž0
๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘˜1๐›ฝโˆ’๐›ฝ0 ๐ฟโˆ—๐‘ก (๐‘0 )
,
๐‘˜2 ๐œ‡ [๐‘ก, ๐‘‡ ]
)1โˆ’๐‘/๐‘0
๐ฟ๐‘ก (๐‘0 )๐‘/๐‘0 .
(5.1)
(5.2)
V.5 Auxiliary Results
103
If ๐‘ < ๐‘0 < 0, the converse inequalities hold, if in (5.1) ๐‘˜1 is replaced by ๐‘˜2 .
If ๐‘ < 0 < ๐‘0 < 1, the converse inequalities hold, if in (5.2) ๐‘˜2 is replaced by
๐‘˜1 .
Proof. We x ๐‘ก and begin with (5.1). To unify the proofs, we rst argue a
๐‘‹ = (๐‘‹๐‘  )๐‘ โˆˆ[๐‘ก,๐‘‡ ] > 0 is optional and ๐›ผ โˆˆ (0, 1), then
]๐›ผ
]
]1โˆ’๐›ผ [ โˆซ ๐‘‡
[โˆซ ๐‘‡
[โˆซ ๐‘‡
๐›ฝ โˆ˜
๐›ฝ ๐›ผ โˆ˜
๐ท๐‘ ๐›ฝ ๐‘‹๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก .
๐ธ
๐ท๐‘  ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐ท๐‘  ๐‘‹๐‘  ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก โ‰ค ๐ธ
๐ธ
Jensen inequality: if
๐‘ก
๐‘ก
๐‘ก
(5.3)
To see this, introduce the probability space
[
๐œˆ(๐ผ × ๐บ) := ๐ธ ๐œ‰ โˆ’1
โˆซ
(
[๐‘ก, ๐‘‡ ]×ฮฉ, โ„ฌ([๐‘ก, ๐‘‡ ])โŠ—โ„ฑ, ๐œˆ
]
1๐บ ๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ ) ,
)
, where
๐บ โˆˆ โ„ฑ, ๐ผ โˆˆ โ„ฌ([๐‘ก, ๐‘‡ ]),
๐ผ
๐œ‰ := ๐ธ[
with the normalizing factor
โˆซ๐‘‡
๐‘ก
๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ )โˆฃโ„ฑ๐‘ก ].
On this space,
๐‘‹
is a
random variable and we have the conditional Jensen inequality
[ ]
[ ]๐›ผ
๐ธ ๐œˆ ๐‘‹ ๐›ผ [๐‘ก, ๐‘‡ ] × โ„ฑ๐‘ก โ‰ค ๐ธ ๐œˆ ๐‘‹ [๐‘ก, ๐‘‡ ] × โ„ฑ๐‘ก
๐œŽ -eld [๐‘ก, ๐‘‡ ] × โ„ฑ๐‘ก := {[๐‘ก, ๐‘‡ ] × ๐ด : ๐ด โˆˆ โ„ฑ๐‘ก }. But this inequality
0
0
coincides with (5.3) if we identify ๐ฟ ([๐‘ก, ๐‘‡ ] × ฮฉ, [๐‘ก, ๐‘‡ ] × โ„ฑ๐‘ก ) and ๐ฟ (ฮฉ, โ„ฑ๐‘ก ) by
for the
using that an element of the rst space is necessarily constant in its time
variable.
0 < ๐‘ โ‰ค ๐‘0 < 1 and let ๐‘Œห† := ๐‘Œห† (๐‘0 ) be the solution of the dual
problem for ๐‘0 . Using (4.4) and then (5.3) with ๐›ผ := ๐‘ž/๐‘ž0 โˆˆ (0, 1) and
(
)๐›ผ
๐‘‹๐‘ ๐›ผ := (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž0 = (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž ,
]
[โˆซ ๐‘‡
(
)๐‘ž
โˆ—
๐ฟ๐‘ก (๐‘) โ‰ค ๐ธ
๐ท๐‘ ๐›ฝ ๐‘Œห†๐‘  /๐‘Œห†๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘ก
]1โˆ’๐‘ž/๐‘ž0 [ โˆซ ๐‘‡
]๐‘ž/๐‘ž0
[โˆซ ๐‘‡
๐›ฝ โˆ˜
โ‰ค๐ธ
๐ท๐‘  ๐œ‡ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐ธ
๐ท๐‘ ๐›ฝ (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž0 ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
.
Let
๐‘ก
Now
๐‘ก
๐ท๐‘ ๐›ฝ โ‰ค ๐‘˜1๐›ฝโˆ’๐›ฝ0 ๐ท๐‘ ๐›ฝ0
๐›ฝ โˆ’ ๐›ฝ0 < 0,
since
which completes the proof of the
๐‘ < 0, the inmum in (4.4) is
either > 1 or < 0, reversing the
rst claim in view of (4.4). In the cases with
replaced by a supremum and
๐›ผ = ๐‘ž/๐‘ž0
is
direction of Jensen's inequality.
We turn to (5.2).
Let
0 < ๐‘ โ‰ค ๐‘0 < 1
and
ห† = ๐‘‹(๐‘)
ห† , ๐‘ห† = ๐‘ห†(๐‘).
๐‘‹
Using (3.1) and the usual Jensen inequality twice,
๐‘‡
[โˆซ
]
๐ท๐‘  ๐‘ห†๐‘๐‘  0 ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘ก
]๐‘0 /๐‘
[โˆซ ๐‘‡
โˆ˜
1โˆ’๐‘0 /๐‘
โ‰ฅ ๐œ‡ [๐‘ก, ๐‘‡ ]
๐ธ
๐ท๐‘ ๐‘/๐‘0 ๐‘ห†๐‘๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
ห† ๐‘0 โ‰ฅ ๐ธ
๐ฟ๐‘ก (๐‘0 )๐‘‹
๐‘ก
๐‘ก
(
โˆ˜
โ‰ฅ ๐‘˜2 ๐œ‡ [๐‘ก, ๐‘‡ ]
)1โˆ’๐‘0 /๐‘ (
ห†๐‘
๐ฟ๐‘ก (๐‘)๐‘‹
๐‘ก
)๐‘0 /๐‘
and the claim follows. The other cases are similar.
V Risk Aversion Asymptotics
104
๐ฟ(๐‘)
๐‘0 .
A useful consequence is that
the possibly critical exponent
Corollary 5.2.
gains moments as
๐‘
moves away from
(i) Let 0 < ๐‘ < ๐‘0 < 1. Then
๐ฟ(๐‘) โ‰ค ๐ถ๐ฟ(๐‘0 )
(5.4)
with a constant ๐ถ independent of ๐‘0 and ๐‘. In the case without intermediate consumption we can take ๐ถ = 1.
(ii) Let ๐‘Ÿ โ‰ฅ 1 and 0 < ๐‘ โ‰ค ๐‘0 /๐‘Ÿ. Then
[
]
๐ธ (๐ฟ๐œ (๐‘))๐‘Ÿ โ‰ค ๐ถ๐‘Ÿ
for all stopping times ๐œ , with a constant ๐ถ๐‘Ÿ independent of ๐‘0 , ๐‘, ๐œ . In
particular, ๐ฟ(๐‘) is of class (D) for all ๐‘ โˆˆ (0, ๐‘0 ).
Proof. (i) Denote ๐ฟ = ๐ฟ(๐‘0 ). By Lemma 4.1, ๐ฟ/๐‘˜1 โ‰ฅ 1, hence ๐ฟ๐‘/๐‘0 =
๐‘/๐‘
๐‘/๐‘0
(๐ฟ/๐‘˜1 )๐‘/๐‘0 โ‰ค ๐‘˜1 0 (๐ฟ/๐‘˜1 ) as ๐‘/๐‘0 โˆˆ (0, 1). Proposition 5.1 yields the
( โˆ˜
)1โˆ’๐‘/๐‘0
result with ๐ถ = ๐œ‡ [0, ๐‘‡ ]๐‘˜2 /๐‘˜1
; note that ๐ถ โ‰ค 1 โˆจ (1 + ๐‘‡ )๐‘˜2 /๐‘˜1 . In
the absence of intermediate consumption we may assume ๐‘˜1 = ๐‘˜2 = 1 by the
subsequent Remark 5.3 and then ๐ถ = 1.
(ii) Let ๐‘Ÿ โ‰ฅ 1, 0 < ๐‘ โ‰ค ๐‘0 /๐‘Ÿ , and ๐ฟ = ๐ฟ(๐‘0 ). Proposition 5.1 shows
(
)๐‘Ÿ(1โˆ’๐‘/๐‘0 ) ๐‘Ÿ๐‘/๐‘0 (
)๐‘Ÿ ๐‘Ÿ๐‘/๐‘
๐ฟ๐‘ก (๐‘)๐‘Ÿ โ‰ค ๐‘˜2 ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]
๐ฟ๐‘ก
โ‰ค (1 โˆจ ๐‘˜2 )(1 + ๐‘‡ ) ๐ฟ๐‘ก 0 .
๐‘˜1
๐‘Ÿ๐‘/๐‘0 โˆˆ (0, 1), thus ๐ฟ๐‘Ÿ๐‘/๐‘0
๐‘Ÿ๐‘/๐‘
๐‘Ÿ๐‘/๐‘
๐ธ[๐ฟ๐œ 0 ] โ‰ค ๐ฟ0 0 โ‰ค 1 โˆจ ๐‘˜2 .
Note
Remark 5.3.
๐ท โ‰ก 1
is a supermartingale by Lemma 4.1 and
In the case without intermediate consumption we may assume
Indeed, ๐ท reduces to the random
๐ท๐‘‡ and can be absorbed into the measure ๐‘ƒ as follows. Under the
หœ with ๐‘ƒ -density process ๐œ‰๐‘ก = ๐ธ[๐ท๐‘‡ โˆฃโ„ฑ๐‘ก ]/๐ธ[๐ท๐‘‡ ], the opportunity
measure ๐‘ƒ
หœ (๐‘ฅ) = 1 ๐‘ฅ๐‘ is ๐ฟ
หœ = ๐ฟ/๐œ‰ by Remark II.3.2. If
process for the utility function ๐‘ˆ
๐‘
หœ
หœ 0 ) and then the
Corollary 5.2(i) is proved for ๐ท โ‰ก 1, we conclude ๐ฟ(๐‘)
โ‰ค ๐ฟ(๐‘
inequality for ๐ฟ follows.
in the proof of Corollary 5.2(i).
variable
Inequality (5.4) is stated for reference as it has a simple form; however,
note that it was deduced using the very poor estimate
the pure investment case, we have
๐ถ=1
๐‘Ž๐‘ โ‰ฅ ๐‘Ž for ๐‘Ž, ๐‘ โ‰ฅ 1.
In
and so (5.4) is a direct comparison
result. Intermediate consumption destroys this monotonicity property: (5.4)
fails for
๐ถ = 1
๐ท โ‰ก1
๐‘ = 0.1
in that case, e.g., if
a standard Brownian motion, and
by explicit calculation.
and
and
๐‘…๐‘ก = ๐‘ก + ๐‘Š๐‘ก , where ๐‘Š is
๐‘0 = 0.2, as can be seen
This is not surprising from a BSDE perspective,
because the driver of (4.6) is not monotone with respect to
of the
๐‘‘๐œ‡-term.
๐‘ in the presence
In the pure investment case, the driver is monotone and so
the comparison result can be expected, even for the entire parameter range.
This is conrmed by the next result; note that the inequality is
to (5.2) for the considered parameters.
converse
V.5 Auxiliary Results
105
Proposition 5.4. Let ๐‘ < ๐‘0 < 0, then
๐ฟ๐‘ก (๐‘) โ‰ค
)๐‘ โˆ’๐‘
๐‘˜2 ( โˆ˜
๐œ‡ [๐‘ก, ๐‘‡ ] 0 ๐ฟ๐‘ก (๐‘0 ).
๐‘˜1
In the case without intermediate consumption, ๐ฟ(๐‘) โ‰ค ๐ฟ(๐‘0 ).
The proof is based on the following auxiliary statement.
Lemma 5.5. Let ๐‘Œ
>0
be a supermartingale. For xed 0 โ‰ค ๐‘ก โ‰ค ๐‘  โ‰ค ๐‘‡ ,
1
( [
]) 1โˆ’๐‘ž
๐‘ž 7โ†’ ๐œ™(๐‘ž) := ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
๐œ™ : (0, 1) โ†’ โ„+ ,
is a monotone decreasing function
we have
(
[ ๐‘ƒ -a.s. If ๐‘Œ is a martingale,
])
๐œ™(1) := lim๐‘žโ†’1โˆ’ ๐œ™(๐‘ž) = exp โˆ’ ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก ) log(๐‘Œ๐‘  /๐‘Œ๐‘ก )โ„ฑ๐‘ก ๐‘ƒ -a.s., where the
conditional expectation has values in โ„ โˆช {+โˆž}.
Lemma 5.5 can be obtained using Jensen's inequality and a suitable
change of measure; see Lemma II.4.10 for details.
Proof of Proposition 5.4.
denote
๐‘Œห† := ๐‘Œห† (๐‘).
โˆซ
๐‘‡
Let
0 < ๐‘ž0 < ๐‘ž < 1
be the dual exponents and
By Lemma 5.5 and Jensen's inequality for
โˆซ
]
[
๐ธ (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ ) โ‰ค
1โˆ’๐‘ž
1โˆ’๐‘ž0
โˆˆ (0, 1),
1โˆ’๐‘ž
( [
]) 1โˆ’๐‘ž
0
๐ธ (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž0 โ„ฑ๐‘ก
๐œ‡โˆ˜ (๐‘‘๐‘ )
๐‘‡
๐‘ก
๐‘ก
( 1โˆ’๐‘ž )( โˆซ
1โˆ’ 1โˆ’๐‘ž
โˆ˜
0
โ‰ค ๐œ‡ [๐‘ก, ๐‘‡ ]
๐‘‡
) 1โˆ’๐‘ž
1โˆ’๐‘ž0
] โˆ˜
๐‘ž0 ห†
ห†
.
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ„ฑ๐‘ก ๐œ‡ (๐‘‘๐‘ )
[
๐‘ก
Using (2.4) and (4.4) twice, we conclude that
๐ฟโˆ—๐‘ก (๐‘) โ‰ค ๐‘˜2๐›ฝ
โˆซ
๐‘‡
]
[
๐ธ (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ )
๐‘ก
1โˆ’๐‘ž
1โˆ’๐‘ž
โˆ’๐›ฝ0 1โˆ’๐‘ž
1โˆ’ 1โˆ’๐‘ž
0 โˆ˜
0
๐‘˜2๐›ฝ ๐‘˜1
๐œ‡ [๐‘ก, ๐‘‡ ]
)( โˆซ
1โˆ’๐‘ž
1โˆ’๐‘ž
โˆ’๐›ฝ0 1โˆ’๐‘ž
1โˆ’ 1โˆ’๐‘ž
0 โˆ˜
0
๐‘˜2๐›ฝ ๐‘˜1
๐œ‡ [๐‘ก, ๐‘‡ ]
)
(
โ‰ค
(
โ‰ค
Now (4.5) and
๐›ฝ = 1โˆ’๐‘ž
๐‘‡
๐ธ
[
๐ท๐‘ ๐›ฝ0 (๐‘Œห†๐‘  /๐‘Œห†๐‘ก )๐‘ž0 โ„ฑ๐‘ก
]
) 1โˆ’๐‘ž
1โˆ’๐‘ž0
๐œ‡ (๐‘‘๐‘ )
โˆ˜
๐‘ก
1โˆ’๐‘ž
๐ฟโˆ—๐‘ก (๐‘0 ) 1โˆ’๐‘ž0 .
yield the rst result. In the case without inter-
mediate consumption, we may assume
๐ทโ‰ก1
and hence
๐‘˜1 = ๐‘˜2 = 1,
as in
Remark 5.3.
Remark 5.6.
Our argument for Proposition 5.4 extends to
๐‘ = โˆ’โˆž
(cf.
Lemma 6.7 below). The proposition generalizes [54, Proposition 2.2], where
the result is proved for the case without intermediate consumption and under
the additional condition that
๐‘ž0 -optimal
๐‘Œห† (๐‘0 ) is a martingale (or equivalently, that the
equivalent martingale measure exists).
V Risk Aversion Asymptotics
106
Propositions 5.1 and 5.4 combine to the following continuity property of
๐‘ 7โ†’ ๐ฟ(๐‘)
at interior points of
(โˆ’โˆž, 0).
We will not pursue this further as
we are interested mainly in the boundary points of this interval.
Corollary 5.7. Assume ๐ท โ‰ก 1 and let ๐ถ๐‘ก := ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]. If ๐‘ โ‰ค ๐‘0 < 0,
1โˆ’๐‘/๐‘0
๐ถ๐‘ก
1โˆ’๐‘0 /๐‘+๐‘0 โˆ’๐‘
๐ฟ(๐‘0 )๐‘/๐‘0 โ‰ค ๐ฟ(๐‘) โ‰ค ๐ถ๐‘ก๐‘0 โˆ’๐‘ ๐ฟ(๐‘0 ) โ‰ค ๐ถ๐‘ก
๐ฟ(๐‘)๐‘0 /๐‘ .
In particular, ๐‘ 7โ†’ ๐ฟ๐‘ก (๐‘) is continuous on (โˆ’โˆž, 0) uniformly in ๐‘ก, ๐‘ƒ -a.s.
Remark 5.8. The optimal propensity to consume ๐œ…ห†(๐‘) is not monotone with
respect to
๐‘ก + ๐‘Š๐‘ก ,
๐‘
๐ท โ‰ก 1 and ๐‘…๐‘ก =
๐‘ โˆˆ {โˆ’1/2, โˆ’1, โˆ’2}.
in general. For instance, monotonicity fails for
where
๐‘Š
is a standard Brownian motion, and
One can note that
๐‘
determines both the risk aversion and the elasticity of
intertemporal substitution (see, e.g., Gollier [27, ยŸ15]). As with any timeadditive utility specication, it is not possible in our setting to study the
dependence on each of these quantities in an isolated way.
V.5.2 ๐ต๐‘€ ๐‘‚ Estimate
In this section we give
๐ต๐‘€ ๐‘‚
estimates for the martingale part of
๐ฟ.
The
following lemma is well known; we state the proof since the argument will
be used also later on.
Lemma 5.9. Let ๐‘‹ be a submartingale satisfying 0 โ‰ค ๐‘‹ โ‰ค ๐›ผ for some
constant ๐›ผ > 0. Then for all stopping times 0 โ‰ค ๐œŽ โ‰ค ๐œ โ‰ค ๐‘‡ ,
]
]
[
[
๐ธ [๐‘‹]๐œ โˆ’ [๐‘‹]๐œŽ โ„ฑ๐œŽ โ‰ค ๐ธ ๐‘‹๐œ2 โˆ’ ๐‘‹๐œŽ2 โ„ฑ๐œŽ .
Proof.
๐‘‹๐‘ก2
=
๐‘‹ โˆซ= ๐‘‹0 + ๐‘€ ๐‘‹ + ๐ด๐‘‹ be the Doob-Meyer
decomposition.
โˆซ๐œ
๐‘ก
)
+
[๐‘‹]
and
2
๐‘‹
๐‘‘๐ด๐‘‹
+ 2 0 ๐‘‹๐‘ โˆ’ (๐‘‘๐‘€๐‘ ๐‘‹ + ๐‘‘๐ด๐‘‹
๐‘ โˆ’
๐‘ก
๐‘  โ‰ฅ 0,
๐‘ 
๐œŽ
Let
๐‘‹02
[๐‘‹]๐œ โˆ’ [๐‘‹]๐œŽ โ‰ค ๐‘‹๐œ2 โˆ’ ๐‘‹๐œŽ2 โˆ’ 2
โˆซ
As
๐œ
๐‘‹๐‘ โˆ’ ๐‘‘๐‘€๐‘ ๐‘‹ .
๐œŽ
๐‘‹โˆ’ โˆ™ ๐‘€ ๐‘‹ is a
๐‘‹
๐‘‹
1
martingale. Indeed, ๐‘‹ is bounded and sup๐‘ก โˆฃ๐‘€๐‘ก โˆฃ โ‰ค 2๐›ผ + ๐ด๐‘‡ โˆˆ ๐ฟ , so the
๐‘‹ 1/2 โˆˆ ๐ฟ1 , hence [๐‘‹ โˆ™ ๐‘€ ๐‘‹ ]1/2 โˆˆ ๐ฟ1 ,
BDG inequalities [18, VII.92] show [๐‘€ ]๐‘‡
โˆ’
๐‘‡
๐‘‹
1
which by the BDG inequalities implies that sup๐‘ก โˆฃ๐‘‹โˆ’ โˆ™ ๐‘€๐‘ก โˆฃ โˆˆ ๐ฟ .
The claim follows by taking conditional expectations because
We wish to apply Lemma 5.9 to
๐ฟ(๐‘)
in the case
๐‘ < 0.
However,
the submartingale property fails in general for the case with intermediate
consumption (cf. Lemma 4.1). We introduce instead a closely related process
having this property.
V.5 Auxiliary Results
107
Lemma 5.10. Let ๐‘ < 0 and consider the case with intermediate consumption. Then
โˆซ ๐‘ก
(
)
๐ต๐‘ก :=
1+๐‘‡ โˆ’๐‘ก
1+๐‘‡
๐‘
๐ฟ๐‘ก +
1
(1 + ๐‘‡ )๐‘
๐ท๐‘  ๐‘‘๐‘ 
0
is a submartingale satisfying 0 < ๐ต๐‘ก โ‰ค ๐‘˜2 (1 + ๐‘‡ )1โˆ’๐‘ .
Proof.
Choose
(๐œ‹, ๐‘) โ‰ก (0, ๐‘ฅ0 /(1 + ๐‘‡ ))
in Proposition II.3.4 to see that
๐ต
is
a submartingale. The bound follows from Lemma 4.1.
We are now in the position to exploit Lemma 5.9.
Lemma 5.11. (i) Let ๐‘1 < 0. There exists a constant ๐ถ = ๐ถ(๐‘1 ) such
that โˆฅ๐‘€ ๐ฟ(๐‘) โˆฅ๐ต๐‘€ ๐‘‚ โ‰ค ๐ถ for all ๐‘ โˆˆ (๐‘1 , 0). In the case without intermediate consumption one can take ๐‘1 = โˆ’โˆž.
(ii) Assume ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for some ๐‘0 โˆˆ (0, 1) and let ๐œŽ be a stopping
time such that ๐ฟ(๐‘0 )๐œŽ โ‰ค ๐›ผ for a constant ๐›ผ > 0. Then there exists
๐ถ โ€ฒ = ๐ถ โ€ฒ (๐›ผ) such that โˆฅ(๐‘€ ๐ฟ(๐‘) )๐œŽ โˆฅ๐ต๐‘€ ๐‘‚ โ‰ค ๐ถ โ€ฒ for all ๐‘ โˆˆ (0, ๐‘0 ].
Proof.
(i) Let
๐‘1 < ๐‘ < 0 and let ๐œ be a stopping time.
]
[
๐ธ [๐ฟ(๐‘)]๐‘‡ โˆ’ [๐ฟ(๐‘)]๐œ โ„ฑ๐œ โ‰ค ๐ถ.
In the case without intermediate consumption,
martingale with
๐ฟ โ‰ค ๐‘˜2
We rst show that
๐ฟ = ๐ฟ(๐‘)
(5.5)
is a positive sub-
(Lemma 4.1), so Lemma 5.9 implies (5.5) with
โˆ’๐‘ก ๐‘
๐ถ = ๐‘˜22 . In the other case, dene ๐ต as in Lemma 5.10 and ๐‘“ (๐‘ก) := ( 1+๐‘‡
1+๐‘‡ ) .
โˆซ ๐‘ก โˆ’2
(๐‘ ) ๐‘‘[๐ต]๐‘  and ๐‘“ โˆ’2 (๐‘ ) โ‰ค 1 as ๐‘“ is increasing with
Then [๐ฟ]๐‘ก โˆ’ [๐ฟ]0 =
0 ๐‘“
โˆซ๐‘‡
๐‘“ (0) = 1. Thus [๐ฟ]๐‘‡ โˆ’ [๐ฟ]๐œ = ๐œ ๐‘“ โˆ’2 (๐‘ ) ๐‘‘[๐ต]๐‘  โ‰ค [๐ต]๐‘‡ โˆ’ [๐ต]๐œ . Now (5.5)
1โˆ’๐‘ and Lemma 5.9 imply
follows since ๐ต โ‰ค ๐‘˜2 (1 + ๐‘‡ )
]
[
๐ธ [๐ต]๐‘‡ โˆ’ [๐ต]๐œ โ„ฑ๐œ โ‰ค ๐‘˜22 (1 + ๐‘‡ )2โˆ’2๐‘ โ‰ค ๐‘˜22 (1 + ๐‘‡ )2โˆ’2๐‘1 =: ๐ถ(๐‘1 ).
[๐ฟ] = ๐ฟ20 + [๐‘€ ๐ฟ ] + [๐ด๐ฟ ] + 2[๐‘€ ๐ฟ , ๐ด๐ฟ ]. Since ๐ด๐ฟ is predictable,
๐‘ := 2[๐‘€ ๐ฟ , ๐ด๐ฟ ] is a local martingale with some localizing sequence (๐œŽ๐‘› ).
๐ฟ
๐ฟ
Moreover, [๐‘€ ]๐‘ก โˆ’ [๐‘€ ]๐‘  = [๐ฟ]๐‘ก โˆ’ [๐ฟ]๐‘  โˆ’ ([๐ด]๐‘ก โˆ’ [๐ด]๐‘  ) โˆ’ (๐‘๐‘ก โˆ’ ๐‘๐‘  ) and (5.5)
We have
imply
[
]
๐ธ [๐‘€ ๐ฟ ]๐‘‡ โˆง๐œŽ๐‘› โˆ’ [๐‘€ ๐ฟ ]๐œ โˆง๐œŽ๐‘› โ„ฑ๐œ โˆง๐œŽ๐‘› โ‰ค ๐ถ.
Choosing
๐œ = 0
and
๐‘› โ†’ โˆž
we see that
[๐‘€ ]๐‘‡ โˆˆ ๐ฟ1 (๐‘ƒ )
and thus Hunt's
lemma [18, V.45] shows the a.s.-convergence in this inequality; i.e., we have
]
[
๐ธ [๐‘€ ๐ฟ ]๐‘‡ โˆ’ [๐‘€ ๐ฟ ]๐œ โ„ฑ๐œ โ‰ค ๐ถ . If ๐ฟ is bounded by ๐›ผ, the jumps
bounded by 2๐›ผ (cf. [34, I.4.24]), therefore
]
[
sup ๐ธ [๐‘€ ๐ฟ ]๐‘‡ โˆ’ [๐‘€ ๐ฟ ]๐œ โˆ’ โ„ฑ๐œ โ‰ค ๐ถ + 4๐›ผ2 .
of
๐‘€๐ฟ
are
๐œ
By Lemma 4.1 we can take
intermediate consumption.
๐›ผ = ๐‘˜2 (1 + ๐‘‡ )1โˆ’๐‘1 ,
and
๐›ผ = ๐‘˜2
when there is no
V Risk Aversion Asymptotics
108
Let 0 < ๐‘ โ‰ค ๐‘0 < 1. The assumption and Corollary 5.2(i) show
๐ฟ(๐‘)๐œŽ โ‰ค ๐ถ๐›ผ for a constant ๐ถ๐›ผ independent of ๐‘ and ๐‘0 . We ap๐œŽ
ply Lemma 5.9 to the nonnegative process ๐‘‹(๐‘) := ๐ถ๐›ผ โˆ’ ๐ฟ(๐‘) , which is
]
[
๐œŽ
๐œŽ
a submartingale by Lemma 4.1, and obtain ๐ธ [๐ฟ(๐‘) ]๐‘‡ โˆ’ [๐ฟ(๐‘) ]๐œ โ„ฑ๐œ
=
]
[
2
๐ธ [๐‘‹(๐‘)]๐‘‡ โˆ’ [๐‘‹(๐‘)]๐œ โ„ฑ๐œ โ‰ค ๐ถ๐›ผ . Now the rest of the proof is as in (i).
(ii)
that
Corollary 5.12. Let ๐‘† be continuous and assume that either ๐‘ โˆˆ (0, 1)
and ๐ฟ is bounded or that ๐‘ < 0 and ๐ฟ is bounded away from zero. Then
๐œ† โˆ™ ๐‘€ โˆˆ ๐ต๐‘€ ๐‘‚, where ๐œ† and ๐‘€ are dened by (2.6).
Proof.
In both cases, the assumed bound and Lemma 4.1 imply that
is bounded away from zero and innity.
in (4.6), we obtain a constant
[โˆซ
๐‘‡
๐ธ
๐ฟโˆ’
๐‘ก
(
๐ถ>0
such that
]
(
๐‘ ๐ฟ )โŠค
๐‘ ๐ฟ )
๐œ†+
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† +
โ„ฑ๐‘ก โ‰ค ๐ถ,
๐ฟโˆ’
๐ฟโˆ’ Moreover, we have
๐‘€ ๐ฟ โˆˆ ๐ต๐‘€ ๐‘‚
๐ฟ
Taking conditional expectations
0 โ‰ค ๐‘ก โ‰ค ๐‘‡.
๐ฟ
โˆซ๐‘‡
๐ธ[ ๐‘ก ๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†โˆฃโ„ฑ๐‘ก ] โ‰ค
โ€ฒ
constant ๐ถ > 0.
by Lemma 5.11. Using the bounds for
and the Cauchy-Schwarz inequality, it follows that
๐ถ โ€ฒ (1 + โˆฅ๐‘ ๐ฟ โˆ™ ๐‘€ โˆฅ๐ต๐‘€ ๐‘‚ ) โ‰ค ๐ถ โ€ฒ (1 + โˆฅ๐‘€ ๐ฟ โˆฅ๐ต๐‘€ ๐‘‚ )
We remark that uniform bounds for
๐ฟ
for a
(as in the condition of Corol-
lary 5.12) are equivalent to a reverse Hölder inequality
ement of the dual domain
index
๐‘ž
Y;
R๐‘ž (๐‘ƒ )
for some el-
see Proposition II.4.5 for details.
Here the
๐‘ž < 1. Therefore, our corollary complements well known
that R๐‘ž (๐‘ƒ ) with ๐‘ž > 1 implies ๐œ† โˆ™ ๐‘€ โˆˆ ๐ต๐‘€ ๐‘‚ (in a suitable
satises
results stating
setting); see, e.g., Delbaen et al. [16, Theorems A,B].
V.6 The Limit ๐‘ โ†’ โˆ’โˆž
The rst goal of this section is to prove Theorem 3.1. Recall that the consumption strategy is related to the opportunity processes via (4.2) and (4.5).
From these relations and the intuition mentioned before Theorem 3.1, we
๐ฟโˆ—๐‘ก = ๐ฟ๐›ฝ๐‘ก converges to ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] as
๐›ฝ = 1/(1 โˆ’ ๐‘) โ†’ 0, this implies that
expect that the dual opportunity process
๐‘ โ†’ โˆ’โˆž. Noting that the exponent
๐ฟ๐‘ก (๐‘) โ†’ โˆž for all ๐‘ก < ๐‘‡ , in the case with intermediate consumption. Thereโˆ—
fore, we shall work here with ๐ฟ rather than ๐ฟ. In the pure investment case,
the situation is dierent as then ๐ฟ โ‰ค ๐‘˜2 (Lemma 4.1). There, the limit of ๐ฟ
yields additional information; this is examined in Section V.6.1 below.
Proposition 6.1. For each ๐‘ก โˆˆ [0, ๐‘‡ ],
lim ๐ฟโˆ—๐‘ก (๐‘) = ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]
๐‘โ†’โˆ’โˆž
๐‘ƒ -a.s.
and in ๐ฟ๐‘Ÿ (๐‘ƒ ), ๐‘Ÿ โˆˆ [1, โˆž),
with a uniform bound. If ๐”ฝ is continuous, the convergences are uniform in ๐‘ก.
V.6 The Limit ๐‘ โ†’ โˆ’โˆž
Remark 6.2.
109
We will use later that the same convergences hold if
๐‘ก
is
replaced by a stopping time, which is an immediate consequence in view
of the uniform bound. Of course, we mean by uniform bound that there
๐ถ > 0,
exists a constant
independent of
๐‘
and
๐‘ก,
0 โ‰ค ๐ฟโˆ—๐‘ก (๐‘) โ‰ค ๐ถ .
such that
Analogous terminology will be used in the sequel.
Proof.
We consider
0 > ๐‘ โ†’ โˆ’โˆž
and note that
๐‘ž โ†’ 1โˆ’
and
๐›ฝ โ†’ 0+.
From
Lemma 4.1 we have
0 โ‰ค ๐ฟโˆ—๐‘ก (๐‘) = ๐ฟ๐›ฝ๐‘ก (๐‘) โ‰ค ๐‘˜2๐›ฝ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] โ†’ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ],
(6.1)
๐‘ก. To obtain a lower bound, we consider the density process ๐‘Œ
๐‘„ โˆˆ M , which exists by assumption (2.1). From (4.4) we have
โˆซ ๐‘‡
]
[
๐ฟโˆ—๐‘ก (๐‘) โ‰ฅ ๐‘˜1๐›ฝ
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ ).
uniformly in
of some
๐‘ก
๐‘  โ‰ฅ ๐‘ก, clearly (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ†’ ๐‘Œ๐‘  /๐‘Œ๐‘ก ๐‘ƒ -a.s. as ๐‘ž โ†’ 1, and noting
๐‘ž
1
bound 0 โ‰ค (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ‰ค 1 + ๐‘Œ๐‘  /๐‘Œ๐‘ก โˆˆ ๐ฟ (๐‘ƒ ) we conclude by dominated
For xed
the
convergence that
Since
]
]
[
[
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โ†’ ๐ธ ๐‘Œ๐‘  /๐‘Œ๐‘ก โ„ฑ๐‘ก โ‰ก 1 ๐‘ƒ -a.s., for all ๐‘  โ‰ฅ ๐‘ก.
]
[
๐‘Œ ๐‘ž is a supermartingale, 0 โ‰ค ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โ‰ค 1. Hence, for
each
๐‘ก,
dominated convergence shows
โˆซ
๐‘‡
]
[
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ ) โ†’ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] ๐‘ƒ -a.s.
๐‘ก
This ends the proof of the rst claim. The convergence in
๐ฟ๐‘Ÿ (๐‘ƒ )
follows by
the bound (6.1).
๐”ฝ is continuous; then the martingale ๐‘Œ is
(๐‘ , ๐œ”) โˆˆ [0, ๐‘‡ ] × ฮฉ we consider (a xed version of )
]1/๐‘ž
[
๐‘“๐‘ž (๐‘ก) := ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
(๐œ”), ๐‘ก โˆˆ [0, ๐‘ ].
Assume that
xed
continuous. For
๐‘ก and increasing in ๐‘ž by Jensen's inequality,
๐‘“๐‘ž โ†’ 1 uniformly[ in ๐‘ก on the
]compact
๐‘ž
๐‘ž
[0, ๐‘ ], by Dini's lemma. The same holds for ๐‘“๐‘ž (๐‘ก) = ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ„ฑ๐‘ก (๐œ”).
โ€ฒ
Fix ๐œ” โˆˆ ฮฉ and let ๐œ€, ๐œ€ > 0. By Egorov's theorem there exist a measurable
โˆ˜
set ๐ผ = ๐ผ(๐œ”) โŠ† [0, ๐‘‡ ] and ๐›ฟ = ๐›ฟ(๐œ”) โˆˆ (0, 1) such that ๐œ‡ ([0, ๐‘‡ ] โˆ– ๐ผ) < ๐œ€ and
[
]
sup๐‘กโˆˆ[0,๐‘ ] โˆฃ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1โˆฃ < ๐œ€โ€ฒ for all ๐‘ž > 1 โˆ’ ๐›ฟ and all ๐‘  โˆˆ ๐ผ . For
๐‘ž > 1 โˆ’ ๐›ฟ and ๐‘ก โˆˆ [0, ๐‘‡ ] we have
โˆซ ๐‘‡
[
]
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1 ๐œ‡โˆ˜ (๐‘‘๐‘ )
๐‘ก โˆซ
โˆซ
[
]
[
]
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1 ๐œ‡โˆ˜ (๐‘‘๐‘ )
โ‰ค ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1 ๐œ‡โˆ˜ (๐‘‘๐‘ ) +
These functions are continuous in
and converge to
๐ผ
โ€ฒ
1
for each
โ‰ค ๐œ€ (1 + ๐‘‡ ) + ๐œ€.
๐‘ก.
Hence
[๐‘ก,๐‘‡ ]โˆ–๐ผ
V Risk Aversion Asymptotics
110
We have shown that
sup๐‘กโˆˆ[0,๐‘‡ ] โˆฃ๐ฟโˆ—๐‘ก (๐‘)โˆ’๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]โˆฃ โ†’ 0 ๐‘ƒ -a.s., and also in ๐ฟ๐‘Ÿ (๐‘ƒ )
by dominated convergence and the uniform bound resulting from (6.1) in
view of
๐‘˜2๐›ฝ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] โ‰ค (1 โˆจ ๐‘˜2 )(1 + ๐‘‡ ).
Under additional continuity assumptions, we will prove that the martingale part of
๐ฟโˆ—
converges to zero in
2 .
โ„‹๐‘™๐‘œ๐‘
We rst need some preparations.
๐‘, it follows from Lemma 4.1 that ๐ฟโˆ— has a canonical decomposition
โˆ—
โˆ—
๐ฟโˆ— = ๐ฟโˆ—0 + ๐‘€ ๐ฟ + ๐ด๐ฟ . When ๐‘† is continuous, we denote the KW decompoโˆ—
โˆ—
๐ฟโˆ— โˆ™ ๐‘€ + ๐‘ ๐ฟโˆ— + ๐ด๐ฟโˆ— . If in addition
sition with respect to ๐‘€ by ๐ฟ = ๐ฟ0 + ๐‘
๐ฟ is continuous, we obtain from ๐ฟโˆ— = ๐ฟ๐›ฝ and (4.7) by Itô's formula that
For each
โˆ—
โˆ—
โˆ—
๐‘€ ๐ฟ = ๐›ฝ๐ฟ๐›ฝโˆ’1 โˆ™ ๐‘€ ๐ฟ ; ๐‘ ๐ฟ /๐ฟโˆ— = ๐›ฝ๐‘ ๐ฟ /๐ฟ; ๐‘ ๐ฟ = ๐›ฝ๐ฟ๐›ฝโˆ’1 โˆ™ ๐‘ ๐ฟ ; (6.2)
โˆซ
โˆซ
โˆซ
(
( โˆ— )โˆ’1
โˆ— )โŠค
โˆ—
โˆ—
๐›ฝ๐œ†๐ฟโˆ— + 2๐‘ ๐ฟ
๐ฟ
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + ๐‘2
๐‘‘โŸจ๐‘ ๐ฟ โŸฉ โˆ’ ๐ท๐›ฝ ๐‘‘๐œ‡.
๐ด๐ฟ = 2๐‘ž
Here
๐‘‘๐œ‡
is a shorthand for
๐œ‡(๐‘‘๐‘ ).
Lemma 6.3. Let ๐‘0 < 0. There exists a localizing sequence (๐œŽ๐‘› ) such that
(
)๐œŽ
๐ฟโˆ— (๐‘) โˆ’๐‘› > 1/๐‘›
simultaneously for all ๐‘ โˆˆ (โˆ’โˆž, ๐‘0 ];
and moreover, if ๐‘† and ๐ฟ(๐‘) are continuous, (๐‘€ ๐ฟ
โˆ— (๐‘)
Proof.
Fix
for all
๐‘ โ‰ค ๐‘0 .
)๐œŽ๐‘› โˆˆ ๐ต๐‘€ ๐‘‚
for ๐‘ โ‰ค ๐‘0 .
๐‘0 < 0 (and corresponding ๐‘ž0 ) and a sequence ๐œ€๐‘› โ†“ 0 in (0, 1). Set
๐œŽ๐‘› = inf{๐‘ก โ‰ฅ 0 : ๐ฟโˆ—๐‘ก (๐‘0 ) โ‰ค ๐œ€๐‘› } โˆง ๐‘‡ . Then ๐œŽ๐‘› โ†’ ๐‘‡ stationarily because each
โˆ—
path of ๐ฟ (๐‘0 ) is bounded away from zero (Lemma 4.1). Proposition 5.1 im( โˆ—
)๐‘ž/๐‘ž0
โˆ—
plies that there is a constant ๐›ผ = ๐›ผ(๐‘0 ) > 0 such that ๐ฟ๐‘ก (๐‘) โ‰ฅ ๐›ผ ๐ฟ๐‘ก (๐‘0 )
It follows that
1/๐‘ž0
๐ฟโˆ—(๐œŽ๐‘› โˆง๐‘ก)โˆ’ (๐‘) โ‰ฅ ๐›ผ๐œ€๐‘›
for all
๐‘ โ‰ค ๐‘0
and we
have proved the rst claim.
Fix ๐‘ โˆˆ (โˆ’โˆž, ๐‘0 ], let ๐‘† and ๐ฟ = ๐ฟ(๐‘) be continuous and recall that
โˆ—
๐‘€ ๐ฟ = ๐›ฝ๐ฟ๐›ฝโˆ’1 โˆ™ ๐‘€ ๐ฟ from (6.2). Noting that ๐›ฝ โˆ’ 1 < 0, we have just shown
๐›ฝโˆ’1 is bounded on [0, ๐œŽ ]. Since ๐‘€ ๐ฟ โˆˆ ๐ต๐‘€ ๐‘‚ by
that the integrand ๐›ฝ๐ฟ
๐‘›
๐ฟโˆ— )๐œŽ๐‘› โˆˆ ๐ต๐‘€ ๐‘‚ .
Lemma 5.11(i), we conclude that (๐‘€
Proposition 6.4. Assume that ๐‘† and ๐ฟ(๐‘) are continuous for all ๐‘ < 0. As
๐‘ โ†’ โˆ’โˆž,
โˆ— (๐‘)
๐‘๐ฟ
Proof.
โ†’0
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) and ๐‘ ๐ฟ
โˆ— (๐‘)
โ†’0
2
in โ„‹๐‘™๐‘œ๐‘
.
๐‘0 < 0 and consider ๐‘ โˆˆ (โˆ’โˆž, ๐‘0 ]. Using Lemma 6.3, we
โˆ—
๐‘€ ๐ฟ (๐‘) โˆˆ โ„‹2 and ๐œ† โˆˆ ๐ฟ2 (๐‘€ ). Dene the
processes ๐‘‹ = ๐‘‹(๐‘) by
We x some
may assume by localization that
continuous
๐‘‹๐‘ก (๐‘) := ๐‘˜2๐›ฝ ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ] โˆ’ ๐ฟโˆ—๐‘ก (๐‘),
0 โ‰ค ๐‘‹(๐‘) โ‰ค (1โˆจ๐‘˜2 )(1+๐‘‡ ) by (6.1). Fix ๐‘. We shall apply Itô's formula
ฮฆ(๐‘‹), where
ฮฆ(๐‘ฅ) := exp(๐‘ฅ) โˆ’ ๐‘ฅ.
then
to
V.6 The Limit ๐‘ โ†’ โˆ’โˆž
For
๐‘ฅ โ‰ฅ 0, ฮฆ
111
satises
ฮฆ(๐‘ฅ) โ‰ฅ 1, ฮฆโ€ฒ (0) = 0, ฮฆโ€ฒ (๐‘ฅ) โ‰ฅ 0, ฮฆโ€ฒโ€ฒ (๐‘ฅ) โ‰ฅ 1, ฮฆโ€ฒโ€ฒ (๐‘ฅ) โˆ’ ฮฆโ€ฒ (๐‘ฅ) = 1.
โˆซ
โˆซ๐‘‡ โ€ฒ
1 ๐‘‡ โ€ฒโ€ฒ
We have
( ๐‘‘๐‘€ ๐‘‹ + ๐‘‘๐ด๐‘‹ ). As
2 0 ฮฆ (๐‘‹) ๐‘‘โŸจ๐‘‹โŸฉ = ฮฆ(๐‘‹๐‘‡ ) โˆ’ ฮฆ(๐‘‹0 ) โˆ’ 0 ฮฆ (๐‘‹)
โˆ—
โ€ฒ
๐‘‹
๐ฟ
ฮฆ (๐‘‹) is like ๐‘‹ uniformly bounded and ๐‘€ = โˆ’๐‘€ โˆˆ โ„‹2 , the stochastic
๐‘‹ is a true martingale and
integral wrt. ๐‘€
]
]
[โˆซ ๐‘‡
[โˆซ ๐‘‡
[
]
โ€ฒโ€ฒ
ฮฆโ€ฒ (๐‘‹) ๐‘‘๐ด๐‘‹ .
ฮฆ (๐‘‹) ๐‘‘โŸจ๐‘‹โŸฉ = 2๐ธ ฮฆ(๐‘‹๐‘‡ ) โˆ’ ฮฆ(๐‘‹0 ) โˆ’ 2๐ธ
๐ธ
0
0
โˆ—
๐‘‘๐ด๐‘‹ = โˆ’๐‘˜2๐›ฝ ๐‘‘๐œ‡ โˆ’ ๐‘‘๐ด๐ฟ , so that (6.2) yields
(
( )โˆ’1
(
)
โˆ— )โŠค
โˆ—
= โˆ’๐‘ž ๐›ฝ๐œ†๐ฟโˆ— + 2๐‘ ๐ฟ
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† โˆ’ ๐‘ ๐ฟโˆ—
๐‘‘โŸจ๐‘ ๐ฟ โŸฉ + 2 ๐ท๐›ฝ โˆ’ ๐‘˜2๐›ฝ ๐‘‘๐œ‡.
Note that
2 ๐‘‘๐ด๐‘‹
Letting
๐ฟโˆ—
๐‘ โ†’ โˆ’โˆž,
๐‘ž โ†’ 1โˆ’
๐‘,
we have
and
๐›ฝ โ†’ 0+.
Hence, using that
๐‘‹
and
are bounded uniformly in
โˆ’๐‘ž๐›ฝ๐ธ
[โˆซ
๐‘‡
]
ฮฆโ€ฒ (๐‘‹)(๐œ†๐ฟโˆ— )โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† โ†’ 0,
0
[โˆซ
๐‘‡
(
) ]
ฮฆโ€ฒ (๐‘‹) ๐ท๐›ฝ โˆ’ ๐‘˜2๐›ฝ ๐‘‘๐œ‡ โ†’ 0,
[ 0
]
๐ธ ฮฆ(๐‘‹๐‘‡ ) โˆ’ ฮฆ(๐‘‹0 ) โ†’ 0,
๐ธ
where the last convergence is due to Proposition 6.1 (and the subsequent
remark). If
๐‘‡
[โˆซ
๐ธ
0
๐‘œ
denotes the sum of these three expectations tending to zero,
]
ฮฆ (๐‘‹) ๐‘‘โŸจ๐‘‹โŸฉ
[โˆซ ๐‘‡
{ ( โˆ—)
}]
( )โˆ’1
โˆ—
โŠค
=๐ธ
ฮฆโ€ฒ (๐‘‹) 2๐‘ž ๐‘ ๐ฟ
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + ๐‘ ๐ฟโˆ—
๐‘‘โŸจ๐‘ ๐ฟ โŸฉ + ๐‘œ.
โ€ฒโ€ฒ
0
( โˆ— )โŠค
โˆ—
โˆ—
๐‘‘โŸจ๐‘‹โŸฉ = ๐‘‘โŸจ๐ฟโˆ— โŸฉ = ๐‘ ๐ฟ
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ + ๐‘‘โŸจ๐‘ ๐ฟ โŸฉ. For the right hand side,
โ€ฒ
we use ฮฆ (๐‘‹) โ‰ฅ 0 and โˆฃ๐‘žโˆฃ < 1 and the Cauchy-Schwarz inequality to obtain
[โˆซ ๐‘‡
{( โˆ— )
}]
โˆ—
โˆ—
โŠค
๐ธ
ฮฆโ€ฒโ€ฒ (๐‘‹) ๐‘ ๐ฟ
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ + ๐‘‘โŸจ๐‘ ๐ฟ โŸฉ
0
[โˆซ ๐‘‡
}]
{( โˆ— )
( โˆ— )โˆ’1
๐ฟโˆ—
โŠค
๐ฟโˆ—
โ€ฒ
๐ฟ โŠค
โ‰ค๐ธ
ฮฆ (๐‘‹) ๐‘
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ + ๐œ† ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + ๐‘ ๐ฟ
๐‘‘โŸจ๐‘ โŸฉ + ๐‘œ.
Note
0
We bring the terms with
[โˆซ
๐‘‡
โˆ—
๐‘๐ฟ
and
โˆ—
๐‘๐ฟ
to the left hand side, then
]
}( ๐ฟโˆ— )โŠค
๐ฟโˆ—
๐ธ
ฮฆ (๐‘‹) โˆ’ ฮฆ (๐‘‹) ๐‘
๐‘‘โŸจ๐‘€ โŸฉ ๐‘
0
[โˆซ ๐‘‡
]
[โˆซ ๐‘‡
]
{ โ€ฒโ€ฒ
( โˆ— )โˆ’1 }
โ€ฒ
๐ฟโˆ—
โ€ฒ
โŠค
+๐ธ
ฮฆ (๐‘‹) โˆ’ ๐‘ฮฆ (๐‘‹) ๐ฟ
๐‘‘โŸจ๐‘ โŸฉ โ‰ค ๐ธ
ฮฆ (๐‘‹)๐œ† ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + ๐‘œ.
{
0
โ€ฒโ€ฒ
โ€ฒ
0
V Risk Aversion Asymptotics
112
As
ฮฆโ€ฒ (0) = 0,
lim๐‘โ†’โˆ’โˆž ฮฆโ€ฒ (๐‘‹๐‘ก ) โ†’ 0 ๐‘ƒ -a.s. for all ๐‘ก, with a uniform
2
๐ฟ (๐‘€ ) implies that the right hand side converges to
we have
๐œ† โˆˆ
โ€ฒโ€ฒ
โ€ฒ
recall ฮฆ โˆ’ ฮฆ โ‰ก 1
bound, hence
zero.
We
and
( )โˆ’1
ฮฆโ€ฒโ€ฒ (๐‘‹) โˆ’ ๐‘ฮฆโ€ฒ (๐‘‹) ๐ฟโˆ—
โ‰ฅ ฮฆโ€ฒโ€ฒ (0) = 1.
Thus both expectations on the left hand side are nonnegative and we can
conclude that they converge to zero; therefore,
and
โˆ—
๐ธ[โŸจ๐‘ ๐ฟ โŸฉ๐‘‡ ] โ†’ 0.
Proof of Theorem 3.1.
๐‘
[โˆซ๐‘‡
0
โˆ—
โˆ—
(๐‘ ๐ฟ )โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ
]
โ†’0
In view of (4.2), part (i) follows from Proposition 6.1;
note that the convergence in
uniformly in
๐ธ
โ„›๐‘Ÿ๐‘™๐‘œ๐‘
is immediate as
by Lemma 6.3 and (4.2).
and (6.2) that
๐œ…
ห† (๐‘)
is locally bounded
For part (ii), recall from (4.8)
โˆ—
๐œ‹
ห† = ๐›ฝ(๐œ† + ๐‘ ๐ฟ /๐ฟ) = ๐›ฝ๐œ† + ๐‘ ๐ฟ /๐ฟโˆ—
๐›ฝ๐œ† โ†’ 0 in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ). By Lemma 6.3, 1/๐ฟโˆ—
is locally bounded uniformly in ๐‘, hence ๐œ‹
ห† (๐‘) โ†’ 0 in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) follows from
Proposition 6.4. As ๐œ…
ห† (๐‘) is locally bounded uniformly in ๐‘, Corollary 8.4(i)
ห† .
from the Appendix yields the convergence of the wealth processes ๐‘‹(๐‘)
for each
๐‘.
As
๐›ฝ โ†’ 0,
clearly
V.6.1 Convergence to the Exponential Problem
In this section, we prove Theorem 3.2 and establish the convergence of the
corresponding opportunity processes. We assume that there is no intermediate consumption, that
contingent claim
๐ต
๐‘†
is locally bounded and satises (3.3), and that the
๐ต later). Hence
๐œ—ห† โˆˆ ฮ˜ for (3.2). It is
capital ๐‘ฅ0 . If ๐œ— โˆˆ ฮ˜,
is bounded (we will choose a specic
there exists an (essentially unique) optimal strategy
๐œ—ห† does not depend on the initial
we denote by ๐บ(๐œ—) = ๐œ— โˆ™ ๐‘… the corresponding gains process and dene
หœ = ๐บ๐‘ก (๐œ—)}. We consider the value process (from
ฮ˜(๐œ—, ๐‘ก) = {๐œ—หœ โˆˆ ฮ˜ : ๐บ๐‘ก (๐œ—)
easy to see that
initial wealth zero) of (3.2),
[
(
) ]
หœ โ„ฑ๐‘ก ,
๐‘‰๐‘ก (๐œ—) := ess sup๐œ—โˆˆฮ˜(๐œ—,๐‘ก)
๐ธ โˆ’ exp ๐ต โˆ’ ๐บ๐‘‡ (๐œ—)
หœ
0 โ‰ค ๐‘ก โ‰ค ๐‘‡.
๐œ—1 , ๐œ—2 โˆˆ ฮ˜ โ‡’ ๐œ—1 1[0,๐‘ก] + ๐œ—2 1(๐‘ก,๐‘‡ ] โˆˆ ฮ˜. With
หœ = ๐บ๐‘ก (๐œ—) + ๐บ๐‘ก,๐‘‡ (๐œ—1
หœ (๐‘ก,๐‘‡ ] ) for ๐œ—หœ โˆˆ ฮ˜(๐œ—, ๐‘ก).
๐บ๐‘‡ (๐œ—)
Note the concatenation property
๐บ๐‘ก,๐‘‡ (๐œ—) :=
โˆซ๐‘‡
๐‘ก
๐œ— ๐‘‘๐‘…,
we have
Therefore, if we dene the exponential opportunity process
[
(
) ]
หœ โ„ฑ๐‘ก ,
๐ฟexp
:= ess inf ๐œ—โˆˆฮ˜
๐ธ exp ๐ต โˆ’ ๐บ๐‘ก,๐‘‡ (๐œ—)
หœ
๐‘ก
0 โ‰ค ๐‘ก โ‰ค ๐‘‡,
(6.3)
then using standard properties of the essential inmum one can check that
๐‘‰๐‘ก (๐œ—) = โˆ’ exp(โˆ’๐บ๐‘ก (๐œ—)) ๐ฟexp
๐‘ก .
๐ฟexp is a reduced form of the value process, analogous to ๐ฟ(๐‘) for power
exp
utility. We also note that ๐ฟ๐‘‡
= exp(๐ต).
Thus
Lemma 6.5. The exponential opportunity process ๐ฟexp is a submartingale
satisfying ๐ฟexp โ‰ค โˆฅ exp(๐ต)โˆฅ๐ฟโˆž (๐‘ƒ ) and ๐ฟexp , ๐ฟexp
โˆ’ > 0.
V.6 The Limit ๐‘ โ†’ โˆ’โˆž
Proof.
113
The martingale optimality principle of dynamic programming, which
๐‘‰ (๐œ—) is a supermartingale for every ๐œ— โˆˆ ฮ˜ such that ๐ธ[๐‘‰โ‹… (๐œ—)] > โˆ’โˆž and a marexp , we
tingale if and only if ๐œ— is optimal. As ๐‘‰ (๐œ—) = โˆ’ exp(โˆ’๐บ(๐œ—)) ๐ฟ
obtain the submartingale property by the choice ๐œ— โ‰ก 0. It follows that
โˆž
โˆž
๐ฟexp โ‰ค โˆฅ๐ฟexp
๐‘‡ โˆฅ๐ฟ = โˆฅ exp(๐ต)โˆฅ๐ฟ .
ห†
The optimal strategy ๐œ— is optimal for all the conditional problems (6.3),
[
(
) ]
exp
ห† โ„ฑ๐‘ก > 0. Thus ๐œ‰ := exp(โˆ’๐บ(๐œ—))
ห† ๐ฟexp
hence ๐ฟ๐‘ก
= ๐ธ exp ๐ต โˆ’ ๐บ๐‘ก,๐‘‡ (๐œ—)
is proved here exactly as in Proposition II.6.2, yields that
is a positive martingale, by the optimality principle. In particular, we have
๐‘ƒ [inf 0โ‰ค๐‘กโ‰ค๐‘‡ ๐œ‰๐‘ก > 0] = 1,
and now the same property for
๐ฟexp
follows.
๐‘† is continuous and denote the KW decomposition of ๐ฟexp
exp = ๐ฟexp + ๐‘ ๐ฟexp โˆ™ ๐‘€ + ๐‘ ๐ฟexp + ๐ด๐ฟexp . Then the
with respect to ๐‘€ by ๐ฟ
0 exp )
( exp ๐ฟexp
triplet (โ„“, ๐‘ง, ๐‘›) := ๐ฟ
,๐‘
, ๐‘๐ฟ
satises the BSDE
Assume that
๐‘‘โ„“๐‘ก =
(
(
๐‘ง ๐‘ก )โŠค
๐‘ง๐‘ก )
1
โ„“๐‘กโˆ’ ๐œ†๐‘ก +
+ ๐‘ง๐‘ก ๐‘‘๐‘€๐‘ก + ๐‘‘๐‘›๐‘ก
๐‘‘โŸจ๐‘€ โŸฉ๐‘ก ๐œ†๐‘ก +
2
โ„“๐‘กโˆ’
โ„“๐‘กโˆ’
with terminal condition
โ„“๐‘‡ = exp(๐ต),
and the optimal strategy
(6.4)
๐œ—ห† is
exp
๐‘๐ฟ
๐œ—ห† = ๐œ† + exp .
๐ฟโˆ’
(6.5)
This can be derived directly by dynamic programming or inferred, e.g., from
Frei and Schweizer [23, Proposition 1]. We will actually reprove the BSDE
later, but present it already at this stage for the following motivation.
We observe that (6.4)
coincides with the BSDE
, except that ๐‘ž is
(4.6)
replaced by 1 and the terminal condition is exp(๐ต) instead of ๐ท๐‘‡ . From now
we assume exp(๐ต) = ๐ท๐‘‡ , then one can guess that the solutions ๐ฟ(๐‘)
on
should converge to
๐ฟexp
as
๐‘ž โ†’ 1โˆ’,
or equivalently
๐‘ โ†’ โˆ’โˆž.
Theorem 6.6. Let ๐‘† be continuous.
(i) As ๐‘ โ†“ โˆ’โˆž, ๐ฟ๐‘ก (๐‘) โ†“ ๐ฟexp
๐‘ƒ -a.s. for all ๐‘ก, with a uniform bound.
๐‘ก
(ii) If ๐ฟ(๐‘) is continuous for each ๐‘ < 0, then ๐ฟexp is also continuous and
the convergence ๐ฟ(๐‘) โ†“ ๐ฟexp is uniform in ๐‘ก, ๐‘ƒ -a.s. Moreover,
(1 โˆ’ ๐‘) ๐œ‹
ห† (๐‘) โ†’ ๐œ—ห†
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ).
We note that (ii) is also a statement about the rate of convergence for
๐œ‹
ห† (๐‘) โ†’ 0 in Theorem 3.1(ii) for the case without intermediate consumption.
The proof occupies most of the remainder of the section. Part (i) follows from
the next two lemmata; recall that the monotonicity of
๐‘ 7โ†’ ๐ฟ๐‘ก (๐‘) was already
established in Proposition 5.4 while the uniform bound is from Lemma 4.1.
Lemma 6.7. We have ๐ฟ(๐‘) โ‰ฅ ๐ฟexp for all ๐‘ < 0.
V Risk Aversion Asymptotics
114
Proof.
๐ต = 0 by a change of measure
๐‘‘๐‘ƒ (๐ต) = (๐‘’๐ต /๐ธ[๐‘’๐ต ]) ๐‘‘๐‘ƒ . Let ๐‘„๐ธ โˆˆ M ๐‘’๐‘›๐‘ก be the measure with
minimal entropy ๐ป(๐‘„โˆฃ๐‘ƒ ); see, e.g., [37, Theorem 3.5]. Let ๐‘Œ be its ๐‘ƒ -density
from
As is well-known, we may assume that
๐‘ƒ
to
process, then
๐‘„
โˆ’ log(๐ฟexp
๐‘ก )=๐ธ
๐ธ
[
]
]
[
log(๐‘Œ๐‘‡ /๐‘Œ๐‘ก )โ„ฑ๐‘ก = ๐ธ (๐‘Œ๐‘‡ /๐‘Œ๐‘ก ) log(๐‘Œ๐‘‡ /๐‘Œ๐‘ก )โ„ฑ๐‘ก .
(6.6)
This is merely a dynamic version of the well-known duality relation stated,
e.g., in [37, Theorem 2.1] and one can retrieve this version, e.g., from [23,
Eq. (8),(10)]. Using the decreasing function
๐œ™
from Lemma 5.5,
(
])
[
๐ฟexp
=
exp
โˆ’
๐ธ
(๐‘Œ
/๐‘Œ
)
log(๐‘Œ
/๐‘Œ
)
= ๐œ™(1)
๐‘ก
๐‘ก โ„ฑ๐‘ก
๐‘‡
๐‘‡
๐‘ก
]1/๐›ฝ
[
โ‰ค ๐ฟโˆ— (๐‘)1/๐›ฝ = ๐ฟ(๐‘),
โ‰ค ๐œ™(๐‘ž) = ๐ธ (๐‘Œ๐‘‡ /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
where (4.4) was used for the second inequality.
Lemma 6.8. Let ๐‘† be continuous. Then lim sup๐‘โ†’โˆ’โˆž ๐ฟ๐‘ก (๐‘) โ‰ค ๐ฟexp
๐‘ก .
Proof. Fix ๐‘ก โˆˆ [0, ๐‘‡ ]. We denote โ„ฐ๐‘ก๐‘‡ (๐‘‹) := โ„ฐ(๐‘‹)๐‘‡ /โ„ฐ(๐‘‹)๐‘ก and similarly
๐‘‹๐‘ก๐‘‡ := ๐‘‹๐‘‡ โˆ’ ๐‘‹๐‘ก .
(i) Let ๐œ— โˆˆ ๐ฟ(๐‘…) be such that โˆฃ๐œ— โˆ™ ๐‘…โˆฃ + โŸจ๐œ— โˆ™ ๐‘…โŸฉ is bounded by a constant.
Noting that ๐ฟ(๐‘…) โŠ† ๐’œ because ๐‘… is continuous, we have from (3.1) that
]
]
[
[
๐‘
๐‘
(โˆฃ๐‘โˆฃโˆ’1 ๐œ— โˆ™ ๐‘…)โ„ฑ๐‘ก
(๐œ‹ โˆ™ ๐‘…)โ„ฑ๐‘ก โ‰ค ๐ธ ๐ท๐‘‡ โ„ฐ๐‘ก๐‘‡
๐ฟ๐‘ก (๐‘) = ess inf ๐œ‹โˆˆ๐’œ ๐ธ ๐ท๐‘‡ โ„ฐ๐‘ก๐‘‡
) ]
[
(
= ๐ธ ๐ท๐‘‡ exp โˆ’ (๐œ— โˆ™ ๐‘…)๐‘ก๐‘‡ + 1 โŸจ๐œ— โˆ™ ๐‘…โŸฉ๐‘ก๐‘‡ โ„ฑ๐‘ก .
2โˆฃ๐‘โˆฃ
The expression under the last conditional expectation is bounded uniformly
[
(
) ]
๐‘, so the last line converges to ๐ธ exp ๐ต โˆ’ (๐œ— โˆ™ ๐‘…)๐‘ก๐‘‡ โ„ฑ๐‘ก ๐‘ƒ -a.s.
๐‘ โ†’ โˆ’โˆž; recall ๐ท๐‘‡ = exp(๐ต). We have shown
[
(
) ]
lim sup ๐ฟ๐‘ก (๐‘) โ‰ค ๐ธ exp ๐ต โˆ’ (๐œ— โˆ™ ๐‘…)๐‘ก๐‘‡ โ„ฑ๐‘ก ๐‘ƒ -a.s.
in
when
(6.7)
๐‘โ†’โˆ’โˆž
๐œ— โˆˆ ๐ฟ(๐‘…) be such that exp(โˆ’๐œ— โˆ™ ๐‘…) is of class (D). Dening
times ๐œ๐‘› = inf{๐‘  > 0 : โˆฃ๐œ— โˆ™ ๐‘…๐‘  โˆฃ + โŸจ๐œ— โˆ™ ๐‘…โŸฉ๐‘  โ‰ฅ ๐‘›}, we have
[
(
) ]
lim sup ๐ฟ๐‘ก (๐‘) โ‰ค ๐ธ exp ๐ต โˆ’ (๐œ— โˆ™ ๐‘…)๐œ๐‘ก๐‘‡๐‘› โ„ฑ๐‘ก ๐‘ƒ -a.s.
(ii) Let
stopping
the
๐‘โ†’โˆ’โˆž
for each
๐‘›,
and also
(iii)
[ ๐œ—1((0,๐œ๐‘› ] . Using the
) class
] (D) property, the
๐ธ exp ๐ต โˆ’ (๐œ— โˆ™ ๐‘…)๐‘ก๐‘‡ โ„ฑ๐‘ก in ๐ฟ1 (๐‘ƒ ) as ๐‘› โ†’ โˆž,
by step (i) applied to
right hand side converges to
๐‘ƒ -a.s.
along a subsequence. Hence (6.7) again holds.
The previous step has a trivial extension: Let
๐‘”๐‘ก๐‘‡ โˆˆ ๐ฟ0 (โ„ฑ๐‘‡ )
๐‘”๐‘ก๐‘‡ โ‰ค (๐œ— โˆ™ ๐‘…)๐‘ก๐‘‡ for some ๐œ— as in (ii).
]
[
lim sup ๐ฟ๐‘ก (๐‘) โ‰ค ๐ธ exp(๐ต โˆ’ ๐‘”๐‘ก๐‘‡ )โ„ฑ๐‘ก ๐‘ƒ -a.s.
random variable such that
๐‘โ†’โˆ’โˆž
Then
be a
V.6 The Limit ๐‘ โ†’ โˆ’โˆž
(iv)
Let
๐‘›
sequence ๐‘”๐‘ก๐‘‡
๐œ—ห† โˆˆ ฮ˜ be the optimal strategy. We claim that there
โˆˆ ๐ฟ0 (โ„ฑ๐‘‡ ) of random variables as in (iii) such that
(
)
(
)
๐‘›
ห† in ๐ฟ1 (๐‘ƒ ).
exp ๐ต โˆ’ ๐‘”๐‘ก๐‘‡
โ†’ exp ๐ต โˆ’ ๐บ๐‘ก,๐‘‡ (๐œ—)
Indeed, we may assume
๐ต = 0,
as in the previous proof.
115
exists a
Then our claim
follows by the construction of Schachermayer [69, Theorem 2.2] applied to
[๐‘ก, ๐‘‡ ]; recall[the denitions
[69, Eq. (4),(5)]. We conclude
(
) ]
ห† โ„ฑ๐‘ก = ๐ฟexp ๐‘ƒ -a.s. by the
that lim sup๐‘โ†’โˆ’โˆž ๐ฟ๐‘ก (๐‘) โ‰ค ๐ธ exp ๐ต โˆ’ ๐บ๐‘ก,๐‘‡ (๐œ—)
๐‘ก
๐ฟ1 (๐‘ƒ )-continuity of the conditional expectation.
the time interval
ห† exp is a martingale, hence of class (D).
Remark 6.9. Recall that exp(โˆ’๐บ(๐œ—))๐ฟ
If
๐ฟexp
is uniformly bounded away from zero, it follows that
ห†
exp(โˆ’๐บ(๐œ—))
is
already of class (D) and the last two steps in the previous proof are unnecessary. This situation occurs precisely when the right hand side of (6.6) is
bounded uniformly in ๐‘ก. In standard terminology, the latter condition states
that the reverse Hölder inequality R๐ฟ log(๐ฟ) (๐‘ƒ ) is satised by the density
process of the minimal entropy martingale measure.
Lemma 6.10. Let ๐‘† be continuous and assume that ๐ฟ(๐‘) is continuous for
all ๐‘ < 0. Then ๐ฟexp is continuous and ๐ฟ๐‘ก (๐‘) โ†’ ๐ฟexp
uniformly in ๐‘ก, ๐‘ƒ -a.s.
๐‘ก
2 .
Moreover, ๐‘ ๐ฟ(๐‘) โ†’ ๐‘ exp in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) and ๐‘ (๐‘) โ†’ ๐‘ exp in โ„‹๐‘™๐‘œ๐‘
We have already identied the monotone limit
๐ฟexp
= lim ๐ฟ๐‘ก (๐‘).
๐‘ก
Hence,
by uniqueness of the KW decomposition, the above lemma follows from the
subsequent one, which we state separately to clarify the argument. The most
important input from the control problems is that by stopping, we can bound
๐ฟ(๐‘)
away from zero simultaneously for all
๐‘
(cf. Lemma 6.3).
Lemma 6.11. Let
( ๐‘† be continuous
) and assume that ๐ฟ(๐‘) is continuous for
๐ฟ(๐‘)
หœ ๐‘,
หœ ๐‘
หœ ) of the
all ๐‘ < 0. Then ๐ฟ(๐‘), ๐‘ , ๐‘ (๐‘) converge to a solution (๐ฟ,
BSDE (6.4) as ๐‘ โ†’ โˆ’โˆž: ๐ฟหœ is continuous and ๐ฟ๐‘ก (๐‘) โ†’ ๐ฟหœ๐‘ก uniformly in ๐‘ก,
หœ in โ„‹2 .
หœ in ๐ฟ2 (๐‘€ ) and ๐‘ (๐‘) โ†’ ๐‘
๐‘ƒ -a.s.; while ๐‘ ๐ฟ(๐‘) โ†’ ๐‘
๐‘™๐‘œ๐‘
๐‘™๐‘œ๐‘
Proof.
For notational simplicity, we write the proof for the one-dimensional
= 1). We x a sequence ๐‘๐‘› โ†“ โˆ’โˆž and corresponding ๐‘ž๐‘› โ†‘ 1. As
หœ ๐‘ก := lim๐‘› ๐ฟ๐‘ก (๐‘๐‘› ) exists.
๐‘ 7โ†’ ๐ฟ๐‘ก (๐‘) is monotone and positive, the ๐‘ƒ -a.s. limit ๐ฟ
๐ฟ(๐‘
)
2
๐‘›
The sequence ๐‘€
of martingales is bounded in the Hilbert space โ„‹
๐ฟ(๐‘
)
๐‘›
by Lemma 5.11(i). Hence it has a subsequence, still denoted by ๐‘€
,
2
2
หœ
which converges to some ๐‘€ โˆˆ โ„‹ in the weak topology of โ„‹ . If we denote
หœ , we have by orthogonality that
หœ = ๐‘หœ โˆ™ ๐‘€ + ๐‘
the KW decomposition by ๐‘€
๐ฟ(๐‘
)
2
๐ฟ(๐‘
)
หœ weakly in ๐ฟ (๐‘€ ) and ๐‘ ๐‘› โ†’ ๐‘
หœ weakly in โ„‹2 . We shall use
๐‘ ๐‘› โ†’๐‘
case (๐‘‘
the BSDE to deduce the strong convergence.
๐‘๐‘› and in (6.4)
1 (
๐‘ง )2
๐‘“ (๐‘ก, ๐‘™, ๐‘ง) := ๐‘™ ๐œ†๐‘ก +
2
๐‘™
The drivers in the BSDE (4.6) corresponding to
๐‘“ ๐‘› (๐‘ก, ๐‘™, ๐‘ง) := ๐‘ž๐‘› ๐‘“ (๐‘ก, ๐‘™, ๐‘ง),
are
V Risk Aversion Asymptotics
116
(๐‘ก, ๐‘™, ๐‘ง) โˆˆ [0, ๐‘‡ ] × (0, โˆž) × โ„.
(๐‘™๐‘š , ๐‘ง๐‘š ) โ†’ (๐‘™, ๐‘ง) โˆˆ (0, โˆž) × โ„, we
for
For xed
๐‘ก
and any convergent sequence
have
๐‘“ ๐‘š (๐‘ก, ๐‘™๐‘š , ๐‘ง๐‘š ) โ†’ ๐‘“ (๐‘ก, ๐‘™, ๐‘ง) ๐‘ƒ -a.s.
By Lemmata 6.7 and 6.5 we can nd a localizing sequence
1/๐‘˜ < ๐ฟ(๐‘)๐œ๐‘˜ โ‰ค ๐‘˜2
for all
(๐œ๐‘˜ )
such that
๐‘ < 0,
where the upper bound is from Lemma 4.1. For the processes from (2.6) we
๐œ†๐œ๐‘˜ โˆˆ ๐ฟ2 (๐‘€ ) and ๐‘€ ๐œ๐‘˜ โˆˆ โ„‹2 for each ๐‘˜ .
๐‘›
๐‘›
๐ฟ(๐‘๐‘› ) , ๐‘ ๐‘› = ๐‘ ๐ฟ(๐‘๐‘› ) , and
To relax the notation, let ๐ฟ = ๐ฟ(๐‘๐‘› ), ๐‘ = ๐‘
๐‘›
๐ฟ(๐‘
)
๐‘›
๐‘›
๐‘›
โˆ™
๐‘€ =๐‘€
= ๐‘ ๐‘€ + ๐‘ . The purpose of the localization is that (๐‘“ ๐‘› )
๐‘›
๐‘› ๐œ
are uniformly quadratic in the relevant domain: As (๐ฟ , ๐‘ ) ๐‘˜ takes values
in [1/๐‘˜, ๐‘˜2 ] × โ„ and
โˆฃ๐‘“ ๐‘› (๐‘ก, ๐‘™, ๐‘ง)โˆฃ โ‰ค ๐‘™๐œ†2๐‘ก + ๐œ†๐‘ก ๐‘ง + ๐‘ง 2 /๐‘™ โ‰ค (1 + ๐‘™)๐œ†2๐‘ก + (1 + 1/๐‘™)๐‘ง 2 ,
may assume that
we have for all
๐‘š, ๐‘› โˆˆ โ„•
that
๐‘›
โˆฃ๐‘“ ๐‘š (๐‘ก, ๐ฟ๐‘›๐‘ก , ๐‘๐‘ก๐‘› )โˆฃ๐œ๐‘˜ โ‰ค ๐œ‰๐‘ก + ๐ถ๐‘˜ (๐‘๐‘กโˆง๐œ
)2 ,
๐‘˜
( )2
๐œ‰ := (1 + ๐‘˜2 ) ๐œ†๐œ๐‘˜ โˆˆ ๐ฟ1๐œ๐‘˜ (๐‘€ ),
where
(6.8)
๐ถ๐‘˜ := 1 + ๐‘˜.
๐ฟ๐‘Ÿ๐œ (๐‘€ ) := {๐ป โˆˆ ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) : ๐ป1[0,๐œ ] โˆˆ ๐ฟ๐‘Ÿ (๐‘€ )} for a stopping time ๐œ and
๐‘Ÿ โ‰ฅ 1. Similarly, we set โ„‹๐œ2 = {๐‘‹ โˆˆ ๐’ฎ : ๐‘‹ ๐œ โˆˆ โ„‹2 }. Now the following can
Here
be shown using a technique of Kobylanski [48].
Lemma 6.12. For xed ๐‘˜,
(i) ๐‘ ๐‘› โ†’ ๐‘หœ in ๐ฟ2๐œ๐‘˜ (๐‘€ ) and ๐‘ ๐‘› โ†’ ๐‘หœ in โ„‹๐œ2๐‘˜ ,
(ii) sup๐‘กโ‰ค๐‘‡ โˆฃ๐ฟ๐‘›๐‘กโˆง๐œ๐‘˜ โˆ’ ๐ฟหœ๐‘กโˆง๐œ๐‘˜ โˆฃ โ†’ 0 ๐‘ƒ -a.s.
๐‘˜ , it follows
หœ
โˆ’ ๐ฟ๐‘ก โˆฃ โ†’ 0 ๐‘ƒ -a.s.
หœ ๐‘,
หœ ๐‘
หœ ) satises the
(๐ฟ,
The proof is deferred to Appendix V.9. Since (ii) holds for all
that
หœ
๐ฟ
๐‘›
is continuous. Now Dini's lemma shows sup๐‘กโ‰ค๐‘‡ โˆฃ๐ฟ๐‘ก
as claimed.
Lemma 6.12 also implies that the limit
[0, ๐œ๐‘˜ ] for all ๐‘˜ , hence on [0, ๐‘‡ ].
๐ฟ๐‘›๐‘‡ = ๐ท๐‘‡ = exp(๐ต) for all ๐‘›.
BSDE (6.4) on
satised as
The terminal condition is
To end the proof, note that the convergences hold for the original sequence
(๐‘๐‘› ),
rather than just for a subsequence, since
and since our choice of
(๐œ๐‘˜ )
๐‘ 7โ†’ ๐ฟ(๐‘) is monotone
does not depend on the subsequence.
We can now nish the proof of Theorem 6.6 (and Theorem 3.2).
Proof of Theorem 6.6.
Part (i) was already proved. For (ii), uniform conver-
gence and continuity were shown in Lemma 6.10. In view of (4.8) and (6.5),
the claim for the strategies is that
exp
๐‘ ๐ฟ(๐‘)
๐‘๐ฟ
(1 โˆ’ ๐‘) ๐œ‹
ห† (๐‘) = ๐œ† +
โ†’ ๐œ† + exp = ๐œ—ห†
๐ฟ(๐‘)
๐ฟ
in
๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ).
V.7 The Limit ๐‘ โ†’ 0
117
๐ฟ(๐‘) +
(๐ฟ(๐‘))โˆ’1 + ๐ฟexp + (๐ฟexp )โˆ’1 is bounded uniformly in ๐‘, and, by Lemma 6.10,
๐ฟ(๐‘) โˆ™ ๐‘€ โ†’ ๐‘ exp โˆ™ ๐‘€ in โ„‹2 . We have
that ๐‘
๐ฟ(๐‘)
๐ฟexp
๐‘
๐ฟ(๐‘) โˆ™ ๐‘€ โˆ’ ๐‘๐ฟexp โˆ™ ๐‘€ 2
โ„‹
(
) exp
exp )
1 ( ๐ฟ(๐‘)
1
๐ฟ
1
โˆ™ ๐‘€
โˆ™ ๐‘€
โ‰ค ๐ฟ(๐‘) ๐‘
๐‘
.
โˆ’ ๐‘๐ฟ
+
โˆ’
๐ฟexp
๐ฟ(๐‘)
2
2
By a localization as in the previous proof, we may assume that
โ„‹
โ„‹
Clearly the rst norm converges to zero. Noting that
๐ต๐‘€ ๐‘‚)
๐‘
๐ฟexp
โˆ™
๐‘€ โˆˆ โ„‹2
(even
due to Lemma 5.9, the second norm tends to zero by dominated
convergence for stochastic integrals.
The last result of this section concerns the convergence of the (normalized) solution
after Remark 3.3.
๐‘Œห† (๐‘)
of the dual problem (4.3); see also the comment
We recall the assumption (3.3) and that there is no
intermediate consumption.
measure which minimizes the relative entropy
๐‘‘๐‘ƒ (๐ต) := (๐‘’๐ต /๐ธ[๐‘’๐ต ]) ๐‘‘๐‘ƒ .
๐‘„๐ธ (๐ต) โˆˆ M be the
๐ป( โ‹… โˆฃ๐‘ƒ (๐ต)) over M , where
To state the result, let
For
๐ต = 0
this is simply the minimal entropy
martingale measure, and the existence of
๐‘„๐ธ (๐ต)
follows from the existence
of the latter by a change of measure.
Proposition 6.13. Let ๐‘† be continuous and assume that ๐ฟ(๐‘) is continuous
for all ๐‘ < 0. Then ๐‘Œห† (๐‘)/๐‘Œห†0 (๐‘) converges in the semimartingale topology to
the density process of ๐‘„๐ธ (๐ต) as ๐‘ โ†’ โˆ’โˆž.
Proof. We deduce from Lemma 6.10 that ๐ฟโˆ’1 โˆ™ ๐‘ โ†’ (๐ฟexp )โˆ’1 โˆ™ ๐‘ exp in
๐‘Œห† /๐‘Œห†0 = โ„ฐ(โˆ’๐œ† โˆ™ ๐‘€ + ๐ฟโˆ’1 โˆ™ ๐‘ ) by (4.9),
(
)
ห† /๐‘Œห†0 โ†’ โ„ฐ โˆ’๐œ† โˆ™ ๐‘€ +(๐ฟexp )โˆ’1 โˆ™ ๐‘ exp in the semiLemma 8.2(ii) shows that ๐‘Œ
๐ธ
martingale topology. The right hand side is the density process of ๐‘„ (๐ต);
2 ,
โ„‹๐‘™๐‘œ๐‘
as in the previous proof. Since
this follows, e.g., from [23, Proposition 1].
V.7 The Limit ๐‘ โ†’ 0
In this section we prove Theorem 3.4, some renements of that result, as well
as the corresponding convergence for the opportunity processes and the dual
problem.
Due to substantial technical dierences, we consider separately
and from above. Recall the semimartingale
below
[ โˆซ ๐‘‡ ๐‘ โ†’ โˆ˜0 from
]
๐ท
๐œ‡
(๐‘‘๐‘ )
โ„ฑ
with
canonical decomposition
๐‘ 
๐‘ก
๐‘ก
] โˆซ ๐‘ก
[โˆซ ๐‘‡
๐œ‚๐‘ก = (๐œ‚0 + ๐‘€๐‘ก๐œ‚ ) + ๐ด๐œ‚๐‘ก = ๐ธ
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก โˆ’
๐ท๐‘  ๐œ‡(๐‘‘๐‘ ).
(7.1)
the limits
๐œ‚๐‘ก = ๐ธ
0
Clearly
๐œ‚
0
is a supermartingale with continuous nite variation part, and a
martingale in the case without intermediate consumption (๐œ‡
= 0).
From (2.4)
we have the uniform bounds
0 < ๐‘˜1 โ‰ค ๐œ‚ โ‰ค (1 + ๐‘‡ )๐‘˜2 .
(7.2)
V Risk Aversion Asymptotics
118
V.7.1 The Limit ๐‘ โ†’ 0โˆ’
We start with the convergence of the opportunity processes.
Proposition 7.1. As ๐‘ โ†’ 0โˆ’,
(i) for each ๐‘ก โˆˆ [0, ๐‘‡ ], ๐ฟโˆ—๐‘ก (๐‘) โ†’ ๐œ‚๐‘ก ๐‘ƒ -a.s. and in ๐ฟ๐‘Ÿ (๐‘ƒ ) for ๐‘Ÿ โˆˆ [1, โˆž),
with a uniform bound.
(ii) if ๐”ฝ is continuous, then ๐ฟโˆ—๐‘ก (๐‘) โ†’ ๐œ‚๐‘ก uniformly in ๐‘ก, ๐‘ƒ -a.s.; and in โ„›๐‘Ÿ
for ๐‘Ÿ โˆˆ [1, โˆž).
(iii) if ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for some ๐‘0 โˆˆ (0, 1), then ๐ฟโˆ—๐‘ก (๐‘) โ†’ ๐œ‚๐‘ก uniformly in ๐‘ก,
๐‘ƒ -a.s.; in โ„›๐‘Ÿ for ๐‘Ÿ โˆˆ [1, โˆž); and prelocally in โ„›โˆž .
The same assertions hold for ๐ฟโˆ— replaced by ๐ฟ.
Proof.
We note that
๐ฟ = (๐ฟโˆ— )1/๐›ฝ ,
๐‘ โ†’ 0โˆ’
๐‘ž โ†’ 0+ and ๐›ฝ โ†’ 1โˆ’. In view
๐ฟโˆ— . From Lemma 4.1,
implies
of
it suces to prove the claims for
0 โ‰ค ๐ฟโˆ—๐‘ก (๐‘) โ‰ค ๐œ‡โˆ˜ [๐‘ก, ๐‘‡ ]โˆ’๐›ฝ๐‘ ๐ธ
[โˆซ
๐‘ก
๐‘‡
]๐›ฝ
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก โ†’ ๐œ‚๐‘ก
To obtain a lower bound, we consider the density process
in
๐‘Œ
โ„›โˆž .
of some
(7.3)
๐‘„ โˆˆ M.
(i) Using (4.4) we obtain
๐ฟโˆ—๐‘ก (๐‘)
๐‘‡
โˆซ
โ‰ฅ
]
[
๐ธ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ ).
๐‘ก
๐ท๐‘ ๐›ฝ โ†’ ๐ท๐‘ 
โ„›โˆž
(๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ†’ 1 ๐‘ƒ -a.s. for ๐‘ž โ†’ 0. We can argue
๐‘ž
1
as in Proposition 6.1: For ๐‘  โ‰ฅ ๐‘ก xed, 0 โ‰ค (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ‰ค 1 + ๐‘Œ๐‘  /๐‘Œ๐‘ก โˆˆ ๐ฟ (๐‘ƒ )
]
[ ๐›ฝ
๐‘ž
๐‘ž
yields ๐ธ ๐ท๐‘  (๐‘Œ๐‘  /๐‘Œ๐‘ก ) โ„ฑ๐‘ก โ†’ ๐ธ[๐ท๐‘  โˆฃโ„ฑ๐‘ก ] ๐‘ƒ -a.s. Since ๐‘Œ is a supermartingale,
]
[
0 โ‰ค ๐ธ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โ‰ค 1 โˆจ ๐‘˜2 , and we conclude for each ๐‘ก that
Clearly
โˆซ
in
and
๐‘‡
๐ธ
[
๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก
]
โˆ˜
๐œ‡ (๐‘‘๐‘ ) โ†’
๐‘‡
[ ]
๐ธ ๐ท๐‘  โ„ฑ๐‘ก ๐œ‡โˆ˜ (๐‘‘๐‘ ) = ๐œ‚๐‘ก ๐‘ƒ -a.s.
๐‘ก
๐‘ก
๐ฟโˆ—๐‘ก (๐‘) โ†’ ๐œ‚๐‘ก ๐‘ƒ -a.s.
Hence
โˆซ
and the convergence in
๐ฟ๐‘Ÿ (๐‘ƒ )
follows by the
bound (7.3).
(ii)
Assume that
๐”ฝ
is continuous.
Our argument will be similar to
Proposition 6.1, but the source of monotonicity is dierent.
[0, ๐‘‡ ] × ฮฉ
] 1
[
๐‘”๐‘ž (๐‘ก) := ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก 1โˆ’๐‘ž (๐œ”),
Then
Fix
(๐‘ , ๐œ”) โˆˆ
and consider
๐‘”๐‘ž (๐‘ก)
is continuous in
Dini's lemma yields
๐‘”๐‘ž โ†’ 1
๐‘ก
and decreasing in
uniformly on
[0, ๐‘ ],
๐‘ก โˆˆ [0, ๐‘ ].
๐‘ž
by virtue of Lemma 5.5.
hence
]
[
๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โ†’ 1
V.7 The Limit ๐‘ โ†’ 0
uniformly in
formly in
๐‘ก
๐‘ก.
We deduce that
119
]
[
๐ธ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก (๐œ”) โ†’ ๐ธ[๐ท๐‘  โˆฃโ„ฑ๐‘ก ](๐œ”)
uni-
since
[
]
๐ธ ๐ท๐‘ ๐›ฝ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ ๐ธ[๐ท๐‘  โˆฃโ„ฑ๐‘ก ]
] [
]
[
โ‰ค ๐ธ โˆฃ๐ท๐‘ ๐›ฝ โˆ’ ๐ท๐‘  โˆฃ(๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก + ๐ธ ๐ท๐‘  {(๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โˆ’ 1}โ„ฑ๐‘ก [
]
]
[
โ‰ค โˆฅ๐ท๐‘ ๐›ฝ โˆ’ ๐ท๐‘  โˆฅ๐ฟโˆž (๐‘ƒ ) ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก + โˆฅ๐ท๐‘  โˆฅ๐ฟโˆž (๐‘ƒ ) ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1
[
]
โ‰ค โˆฅ๐ท๐‘ ๐›ฝ โˆ’ ๐ท๐‘  โˆฅ๐ฟโˆž (๐‘ƒ ) + ๐‘˜2 ๐ธ (๐‘Œ๐‘  /๐‘Œ๐‘ก )๐‘ž โ„ฑ๐‘ก โˆ’ 1.
The rest of the argument is like the end of the proof of Proposition 6.1.
(iii) Let
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž
for some
๐‘0 โˆˆ (0, 1).
Then we can take a dierent
approach via Proposition 5.1, which shows that
๐ฟโˆ—๐‘ก (๐‘)
โ‰ฅ ๐ธ
[โˆซ
๐‘ก
๐‘‡
]1โˆ’๐‘ž/๐‘ž0 (
)๐‘ž/๐‘ž0
๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘˜1๐›ฝโˆ’๐›ฝ0 ๐ฟโˆ—๐‘ก (๐‘0 )
๐‘ < 0, where we note that ๐‘ž0 < 0. Using that almost every path of
โˆ—
๐ฟ (๐‘0 ) is bounded and bounded away from zero (Lemma 4.1), the right hand
โˆซ๐‘‡
โˆ˜
side ๐‘ƒ -a.s. tends to ๐œ‚๐‘ก = ๐ธ[
๐‘ก ๐ท๐‘  ๐œ‡ (๐‘‘๐‘ )โˆฃโ„ฑ๐‘ก ] uniformly in ๐‘ก as ๐‘ž โ†’ 0. Since
๐ฟโˆ— (๐‘0 ) is prelocally bounded, the prelocal convergence in โ„›โˆž follows in the
for all
same way.
Remark 7.2.
even in
โ„›โˆž .
One can ask when the convergence in Proposition 7.1 holds
The following statements remain valid if
๐ฟโˆ—
replaced by
๐ฟ.
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for some ๐‘0 โˆˆ (0, 1), and in addition
โˆ—
that ๐ฟ (๐‘0 ) is (locally) bounded. Then the argument for Proposiโˆ—
โˆž (โ„›โˆž ).
tion 7.1(iii) shows ๐ฟ (๐‘) โ†’ ๐œ‚ in โ„›
๐‘™๐‘œ๐‘
โˆ—
โˆž (โ„›โˆž ) implies that ๐ฟโˆ— (๐‘) is (locally)
(ii) Conversely, ๐ฟ (๐‘) โ†’ ๐œ‚ in โ„›
๐‘™๐‘œ๐‘
bounded away from zero for all ๐‘ < 0 close to zero, because ๐œ‚ โ‰ฅ ๐‘˜1 > 0.
(i) Assume again that
As we turn to the convergence of the martingale part
๐‘€ ๐ฟ(๐‘) ,
a suitable
localization will again be crucial.
Lemma 7.3. Let ๐‘1 < 0. There exists a localizing sequence (๐œŽ๐‘› ) such that
(
Proof.
)๐œŽ
๐ฟ(๐‘) โˆ’๐‘› > 1/๐‘›
simultaneously for all ๐‘ โˆˆ [๐‘1 , 0).
This follows from Proposition 5.4 and Lemma 4.1.
Next, we state a basic result (i) for the convergence of
๐‘€ ๐ฟ(๐‘)
in
2
โ„‹๐‘™๐‘œ๐‘
and
stronger convergences under additional assumptions (ii) and (iii), for which
Remark 7.2(i) gives sucient conditions.
V Risk Aversion Asymptotics
120
Proposition 7.4. Assume that ๐‘† is continuous. As ๐‘ โ†’ 0โˆ’,
2 .
(i) ๐‘€ ๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚ in โ„‹๐‘™๐‘œ๐‘
๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚ in ๐ต๐‘€ ๐‘‚ .
(ii) if ๐ฟ(๐‘) โ†’ ๐œ‚ in โ„›โˆž
๐‘™๐‘œ๐‘
๐‘™๐‘œ๐‘ , then ๐‘€
โˆž
๐ฟ(๐‘)
๐œ‚
(iii) if ๐ฟ(๐‘) โ†’ ๐œ‚ in โ„› , then ๐‘€
โ†’ ๐‘€ in ๐ต๐‘€ ๐‘‚.
Proof.
๐‘‹ = ๐‘‹(๐‘) = ๐œ‚ โˆ’ ๐ฟ(๐‘).
๐‘‹ is bounded uniformly in ๐‘
๐‘€ ๐‘‹(๐‘) โ†’ 0. Lemma 5.9 applied to
โˆฅ๐œ‚โˆฅโˆž โˆ’๐œ‚ shows that ๐‘€ ๐œ‚ โˆˆ ๐ต๐‘€ ๐‘‚. We may restrict our attention to ๐‘ in some
๐ฟ(๐‘) โˆฅ
interval [๐‘1 , 0) and Lemma 5.11 shows that sup๐‘โˆˆ[๐‘ ,0) โˆฅ๐‘€
๐ต๐‘€ ๐‘‚ < โˆž.
1
๐ฟ
๐ฟ
๐ฟ
โˆ™
Due to the orthogonality of the sum ๐‘€
= ๐‘
๐‘€ + ๐‘ , we have in
Set
Then
by Lemma 4.1 and our aim is to prove
particular that
sup โˆฅ๐‘ ๐ฟ (๐‘) โˆ™ ๐‘€ โˆฅ๐ต๐‘€ ๐‘‚ < โˆž.
(7.4)
๐‘โˆˆ[๐‘1 ,0)
Under the condition of (iii),
to zero since
๐œ‚ โ‰ฅ ๐‘˜1 > 0;
๐ฟ(๐‘)
is bounded away from zero for all
moreover,
๐œ†
โˆ™
๐‘€ โˆˆ ๐ต๐‘€ ๐‘‚
(i) and (ii) we may assume by a localization as in Lemma 7.3 that
bounded away from zero uniformly in
assume that
๐œ† โˆ™ ๐‘€ โˆˆ ๐ต๐‘€ ๐‘‚,
๐‘.
Since
๐‘€
๐‘
close
by Corollary 5.12. For
๐ฟโˆ’ (๐‘)
is
is continuous, we may also
by another localization.
Using the formula (4.7) for
๐ด๐ฟ
and the decomposition (7.1) of
๐œ‚,
the
๐‘‹ is continuous and
nite variation part ๐ด
{
}
2 ๐‘‘๐ด๐‘‹ = 2 (1 โˆ’ ๐‘)๐ท๐›ฝ ๐ฟ๐‘žโˆ’ โˆ’ ๐ท ๐‘‘๐œ‡
(7.5)
{
}
( ๐ฟ )โŠค
โˆ’ ๐‘ž ๐ฟโˆ’ ๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + 2๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ + ๐ฟโˆ’1
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ .
โˆ’ ๐‘
In particular, we note that
๐‘‹
[๐‘€ ] = [๐‘‹] โˆ’
๐‘‹02
2
=๐‘‹ โˆ’
๐‘‹02
โˆซ
โˆ’2
๐‘‹โˆ’ ๐‘‘๐‘‹.
(7.6)
๐‘‹02 โ†’ 0 and ๐ธ[๐‘‹๐‘‡2 ] โ†’ 0 by Proposition 7.1 (Remark 6.2
โˆž by assumption and under (ii)
applies). In case (iii) we have ๐‘‹ โ†’ 0 in โ„›
]
[ 2
1
2
the same holds after a localization. If we denote ๐‘œ๐‘ก := ๐ธ ๐‘‹๐‘‡ โˆ’ ๐‘‹๐‘ก โ„ฑ๐‘ก , we
1
1
โˆž in cases (ii) and
therefore have that ๐‘œ0 โ†’ 0 in case (i) and ๐‘œ โ†’ 0 in โ„›
]
[โˆซ๐‘‡
๐‘ž
2
๐›ฝ
(iii). Denote also ๐‘œ๐‘ก := 2๐ธ
๐‘ก ๐‘‹โˆ’ {(1 โˆ’ ๐‘)๐ท ๐ฟโˆ’ โˆ’ ๐ท} ๐‘‘๐œ‡ โ„ฑ๐‘ก . Recalling
๐‘ž
๐›ฝ
that ๐‘ โ†’ 0โˆ’ implies ๐‘ž โ†’ 0+ and ๐›ฝ โ†’ 1โˆ’, we have (1 โˆ’ ๐‘)๐ท ๐ฟโˆ’ โˆ’ ๐ท โ†’ 0 in
2
โˆž
โ„›โˆž and since ๐‘‹โˆ’ is bounded uniformly
follows that ๐‘œ โ†’ 0 in โ„› .
โˆซ in ๐‘, it ๐‘‹
๐‘‹
As ๐‘€
โˆˆ ๐ต๐‘€ ๐‘‚ and ๐‘‹โˆ’ is bounded, ๐‘‹โˆ’ ๐‘‘๐‘€ is a martingale and (7.6)
For case (i) we have
yields
[
]
]
[
[
๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐‘ก โ„ฑ๐‘ก = ๐ธ ๐‘‹๐‘‡2 โˆ’ ๐‘‹๐‘ก2 โ„ฑ๐‘ก โˆ’ 2๐ธ
โˆซ
๐‘ก
๐‘‡
]
๐‘‹โˆ’ ๐‘‘๐ด โ„ฑ๐‘ก .
๐‘‹
V.7 The Limit ๐‘ โ†’ 0
Using (7.5) and the denitions of
๐‘œ1
and
๐‘œ2 ,
121
we can rewrite this as
]
[
๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐‘ก โ„ฑ๐‘ก โˆ’ ๐‘œ1๐‘ก + ๐‘œ2๐‘ก
[โˆซ ๐‘‡
{
( ๐ฟ )โŠค
} ]
๐‘‹โˆ’ ๐ฟโˆ’ ๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ† + 2๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ + ๐ฟโˆ’1
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ ๐ฟ โ„ฑ๐‘ก .
= ๐‘ž๐ธ
โˆ’ ๐‘
๐‘ก
Applying the Cauchy-Schwarz inequality and using that
๐‘‹โˆ’ , ๐ฟโˆ’ , ๐ฟโˆ’1
โˆ’
are
๐‘, it follows that
]
[
๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐‘ก โ„ฑ๐‘ก โˆ’ ๐‘œ1๐‘ก + ๐‘œ2๐‘ก
]
[โˆซ ๐‘‡
๐‘‹โˆ’ (1 + ๐ฟโˆ’ )๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†โ„ฑ๐‘ก
โ‰ค ๐‘ž๐ธ
๐‘ก
]
[โˆซ ๐‘‡
( ๐ฟ )โŠค
โˆ’1
๐ฟ
๐‘‹โˆ’ (1 + ๐ฟโˆ’ ) ๐‘
+ ๐‘ž๐ธ
๐‘‘โŸจ๐‘€ โŸฉ ๐‘ โ„ฑ๐‘ก
๐‘ก
(
)
โ‰ค ๐‘ž๐ถ โˆฅ๐œ† โˆ™ ๐‘€ โˆฅ๐ต๐‘€ ๐‘‚ + โˆฅ๐‘ ๐ฟ(๐‘) โˆ™ ๐‘€ โˆฅ๐ต๐‘€ ๐‘‚ ,
bounded uniformly in
where
๐ถ > 0 is a constant independent of ๐‘ and ๐‘ก. In view of
๐‘ž๐ถ โ€ฒ with a constant ๐ถ โ€ฒ > 0 and we
]
[
๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐‘ก โ„ฑ๐‘ก โ‰ค ๐‘ž๐ถ โ€ฒ + ๐‘œ1๐‘ก โˆ’ ๐‘œ2๐‘ก .
hand side is bounded by
(7.4), the right
have
For (i) we only have to prove the convergence to zero of the left hand side
for
๐‘ก = 0 and so this ends the proof. For (ii) and (iii) we observe that
= [๐‘€ ๐‘‹ ]๐‘กโˆ’ + (ฮ”๐‘€๐‘ก๐‘‹ )2 and โˆฃฮ”๐‘€ ๐‘‹ โˆฃ = โˆฃฮ”๐‘‹โˆฃ โ‰ค 2โˆฅ๐‘‹โˆฅโ„›โˆž to obtain
]
[
sup ๐ธ [๐‘€ ๐‘‹ ]๐‘‡ โˆ’ [๐‘€ ๐‘‹ ]๐‘กโˆ’ โ„ฑ๐‘ก โ‰ค ๐‘ž๐ถ โ€ฒ + โˆฅ๐‘œ1 โˆฅโ„›โˆž + โˆฅ๐‘œ2 โˆฅโ„›โˆž + 4โˆฅ๐‘‹โˆฅ2โ„›โˆž
[๐‘€ ๐‘‹ ]๐‘ก
๐‘กโ‰ค๐‘‡
and we have seen that the right hand side tends to
0
as
๐‘ โ†’ 0โˆ’.
V.7.2 The Limit ๐‘ โ†’ 0+
๐ฟ(๐‘) for ๐‘ โ†’ 0+ is meaningless without supposing
that ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž for some ๐‘0 โˆˆ (0, 1), so we make this a standing assumption
We notice that the limit of
for the entire Section V.7.2. We begin with a result on the integrability of
the tail of the sequence.
Lemma 7.5. Let
that
1 โ‰ค ๐‘Ÿ < โˆž.
ess sup
๐‘กโˆˆ[0,๐‘‡ ], ๐‘โˆˆ(0,๐‘0 /๐‘Ÿ]
Proof.
There exists a localizing sequence (๐œŽ๐‘› ) such
๐ฟ๐‘กโˆง๐œŽ๐‘› (๐‘)
is in ๐ฟ๐‘Ÿ (๐‘ƒ ) for all ๐‘›.
๐‘1 = ๐‘0 /๐‘Ÿ and ๐œŽ๐‘› = inf{๐‘ก > 0 : ๐ฟ๐‘ก (๐‘1 ) > ๐‘›} โˆง ๐‘‡ , then by
๐‘Ÿ
Corollary 5.2(ii), sup๐‘ก ๐ฟ๐‘กโˆง๐œŽ๐‘› (๐‘1 ) โ‰ค ๐‘› + ฮ”๐ฟ๐œŽ๐‘› (๐‘1 ) โˆˆ ๐ฟ (๐‘ƒ ). But ๐ฟ(๐‘) โ‰ค
๐ถ๐ฟ(๐‘1 ) by Corollary 5.2(i), so (๐œŽ๐‘› ) already satises the requirement.
Let
122
V Risk Aversion Asymptotics
Proposition 7.6. As ๐‘ โ†’ 0+,
๐ฟโˆ— (๐‘) โ†’ ๐œ‚,
uniformly in ๐‘ก, ๐‘ƒ -a.s.; in โ„›๐‘Ÿ๐‘™๐‘œ๐‘ for ๐‘Ÿ โˆˆ [1, โˆž); and prelocally in โ„›โˆž . Moreover, the convergence takes place in โ„›โˆž (in โ„›โˆž
๐‘™๐‘œ๐‘ ) if and only if ๐ฟ(๐‘1 ) is
(locally) bounded for some ๐‘1 โˆˆ (0, ๐‘0 ). The same assertions hold for ๐ฟโˆ—
replaced by ๐ฟ.
Proof.
implies
claims
๐‘ โˆˆ (0, ๐‘0 ) in this proof and recall that ๐‘ โ†’ 0+
๐‘ž โ†’ 0โˆ’ and ๐›ฝ โ†’ 1โˆ’. Since ๐ฟ = (๐ฟโˆ— )1/๐›ฝ , it suces to prove the
โˆ—
for ๐ฟ . Using Lemma 4.1,
]๐›ฝ
[โˆซ ๐‘‡
โˆ—
โˆ˜
โˆ’๐›ฝ๐‘
(7.7)
๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก โ†’ ๐œ‚๐‘ก in โ„›โˆž .
๐ฟ๐‘ก (๐‘) โ‰ฅ ๐œ‡ [๐‘ก, ๐‘‡ ]
๐ธ
We consider only
๐‘ก
Conversely, by Proposition 5.1,
๐ฟโˆ—๐‘ก (๐‘)
โ‰ค ๐ธ
[โˆซ
๐‘‡
๐‘ก
]1โˆ’๐‘ž/๐‘ž0 (
)๐‘ž/๐‘ž0
๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ )โ„ฑ๐‘ก
๐‘˜1๐›ฝโˆ’๐›ฝ0 ๐ฟโˆ—๐‘ก (๐‘0 )
.
๐ฟโˆ— (๐‘0 ) is
๐‘ก as ๐‘ž โ†’ 0โˆ’.
Since almost every path of
tends to
๐œ‚๐‘ก
uniformly in
bounded, the right hand side
By localizing
๐ฟโˆ— (๐‘0 )
โ†’ ๐œ‚
uniformly in
๐‘ƒ -a.s.
to be prelocally
bounded, the same argument shows the prelocal convergence in
โˆ—
We have proved that ๐ฟ (๐‘)
(7.8)
๐‘ก, ๐‘ƒ -a.s.
โ„›โˆž .
In view of
๐‘Ÿ
Lemma 7.5, the convergence in โ„›๐‘™๐‘œ๐‘ follows by dominated convergence.
For the second claim, note that the if statement is shown exactly like
โ„›โˆž convergence and the converse holds by boundedness of ๐œ‚ .
Of course, if ๐ฟ(๐‘1 ) is (locally) bounded for some ๐‘1 โˆˆ (0, ๐‘0 ), then in fact
๐ฟ(๐‘) has this property for all ๐‘ โˆˆ (0, ๐‘1 ], by Corollary 5.2(i).
the prelocal
We turn to the convergence of the martingale part. The major diculty
will be that
๐ฟ(๐‘)
may have unbounded jumps; i.e., we have to prove the
convergence of quadratic BSDEs whose solutions are not locally bounded.
Proposition 7.7. Assume that ๐‘† is continuous. As ๐‘ โ†’ 0+,
2 .
(i) ๐‘€ ๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚ in โ„‹๐‘™๐‘œ๐‘
(ii) if there exists ๐‘1 โˆˆ (0, ๐‘0 ] such that ๐ฟ(๐‘1 ) is locally bounded, then
๐‘€ ๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚ in ๐ต๐‘€ ๐‘‚๐‘™๐‘œ๐‘ .
(iii) if there exists ๐‘1 โˆˆ (0, ๐‘0 ] such that ๐ฟ(๐‘1 ) is bounded, then ๐‘€ ๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚
in ๐ต๐‘€ ๐‘‚.
The following terminology will be useful in the proof. We say that real
numbers
(๐‘ฅ๐œ€ )
converge to
๐‘ฅ
linearly
as
๐œ€โ†’0
if
lim sup 1๐œ€ โˆฃ๐‘ฅ๐œ€ โˆ’ ๐‘ฅโˆฃ < โˆž.
๐œ€โ†’0+
V.7 The Limit ๐‘ โ†’ 0
123
Lemma 7.8. Let ๐‘ฅ๐œ€ โ†’ ๐‘ฅ linearly and ๐‘ฆ๐œ€ โ†’ ๐‘ฆ linearly. Then
(i) lim sup๐œ€โ†’0 1๐œ€ โˆฃ๐‘ฅ๐œ€ โˆ’ ๐‘ฆ๐œ€ โˆฃ < โˆž if ๐‘ฅ = ๐‘ฆ,
(ii) ๐‘ฅ๐œ€ ๐‘ฆ๐œ€ โ†’ ๐‘ฅ๐‘ฆ linearly,
(iii) if ๐‘ฅ > 0 and ๐œ‘ is a real function with ๐œ‘(0) = 1 and dierentiable at 0,
then (๐‘ฅ๐œ€ )๐œ‘(๐œ€) โ†’ ๐‘ฅ linearly.
Proof.
(i) This is immediate from the triangle inequality.
(ii) This follows from
โˆฃ๐‘ฅ๐œ€ ๐‘ฆ๐œ€ โˆ’ ๐‘ฅ๐‘ฆโˆฃ โ‰ค โˆฃ๐‘ฅ๐œ€ โˆฃโˆฃ๐‘ฆ๐œ€ โˆ’ ๐‘ฆโˆฃ + โˆฃ๐‘ฆโˆฃโˆฃ๐‘ฅ๐œ€ โˆ’ ๐‘ฅโˆฃ
because
convergent sequences are bounded.
(iii) Here we use
โˆฃ(๐‘ฅ๐œ€ )๐œ‘(๐œ€) โˆ’ ๐‘ฅโˆฃ โ‰ค โˆฃ๐‘ฅ๐œ€ โˆฃโˆฃ(๐‘ฅ๐œ€ )๐œ‘(๐œ€)โˆ’1 โˆ’ 1โˆฃ + โˆฃ๐‘ฅ๐œ€ โˆ’ ๐‘ฅโˆฃ;
๐‘ฅ๐œ€ โ†’ ๐‘ฅ linearly, the question is reduced to the
โˆ’1
๐œ‘(๐œ€)โˆ’1
boundedness of ๐œ€
โˆฃ(๐‘ฅ๐œ€ )
โˆ’ 1โˆฃ. Fix 0 < ๐›ฟ1 < ๐‘ฅ < ๐›ฟ2 and ๐œš(๐›ฟ, ๐œ€) :=
โˆฃ๐›ฟ ๐œ‘(๐œ€)โˆ’1 โˆ’ 1โˆฃ. For ๐œ€ small enough, ๐‘ฅ๐œ€ โˆˆ [๐›ฟ1 , ๐›ฟ2 ] and then
as
{๐‘ฅ๐œ€ }
is bounded and
๐œš(๐›ฟ1 , ๐œ€) โˆง ๐œš(๐›ฟ2 , ๐œ€) โ‰ค โˆฃ(๐‘ฅ๐œ€ )๐œ‘(๐œ€)โˆ’1 โˆ’ 1โˆฃ โ‰ค ๐œš(๐›ฟ1 , ๐œ€) โˆจ ๐œš(๐›ฟ2 , ๐œ€).
โˆ’1 โˆฃ๐œš(๐›ฟ, ๐œ€)โˆฃ = ๐‘‘ ๐›ฟ ๐œ‘(๐œ€) โˆฃ
= โˆฃ log(๐›ฟ)๐œ‘โ€ฒ (0)โˆฃ < โˆž.
For ๐›ฟ > 0 we have lim๐œ€ ๐œ€
๐œ€=0
๐‘‘๐œ€
Hence the upper and the lower bound above converge to 0 linearly.
Proof of Proposition 7.7.
We rst prove (ii) and (iii), i.e, we assume that
๐ฟ(๐‘) โ‰ฅ ๐‘˜1 from Lemma 4.1.
By Corollary 5.2(i) there exists a constant ๐ถ > 0 independent of ๐‘ such that
๐ฟ(๐‘) โ‰ค ๐ถ๐ฟ(๐‘1 ) for all ๐‘ โˆˆ (0, ๐‘1 ]. Hence ๐ฟ(๐‘) is bounded uniformly in
๐‘ โˆˆ (0, ๐‘1 ] in the case (iii) and for (ii) this holds after a localization. Now
๐ฟ(๐‘) โˆฅ
Lemma 5.11(ii) implies sup๐‘โˆˆ(0,๐‘ ] โˆฅ๐‘€
๐ต๐‘€ ๐‘‚ < โˆž and we can proceed
1
๐ฟ(๐‘1 ) is locally bounded (resp. bounded).
Recall
exactly as in the proof of items (ii) and (iii) of Proposition 7.4.
(i)
This case is more dicult because we have to use prelocal bounds
and Lemma 5.11(ii) does not apply.
Again, we want to imitate the proof
of Proposition 7.4(i), or more precisely, the arguments after (7.6). We note
2 -convergence those estimates are required only at
โ„‹๐‘™๐‘œ๐‘
๐ต๐‘€ ๐‘‚-norms can be replaced by โ„‹2 -norms. Inspecting
that for the claimed
๐‘ก = 0
and so the
that proof in detail, we see that we can proceed in the same way once we
establish:
โˆ™
There exists a localizing sequence
all
(๐œŽ๐‘› )
and constants
๐ถ๐‘›
such that for
๐‘›,
(b)
(๐ป1[0,๐œŽ๐‘› ] ) โˆ™ ๐‘€ ๐ฟ(๐‘) is a martingale for all ๐ป predictable and bounded, and all ๐‘ โˆˆ (0, ๐‘0 ),
(
)
sup๐‘โˆˆ(0,๐‘0 ] ๐ฟโˆ’ (๐‘) + ๐ฟโˆ’1
โˆ’ (๐‘) โ‰ค ๐ถ๐‘› on [0, ๐œŽ๐‘› ],
(c)
lim sup๐‘โ†’0+ โˆฅ๐‘ ๐ฟ(๐‘) 1[0,๐œŽ๐‘› ] โˆฅ๐ฟ2 (๐‘€ ) โ‰ค ๐ถ๐‘› .
(a)
V Risk Aversion Asymptotics
124
We may assume by localization that
๐œ† โˆ™ ๐‘€ โˆˆ โ„‹2 .
We now prove (a)-(c);
(๐œŽ๐‘› ) explicitly, we write by localization. . . as usual.
(a) Fix ๐‘ โˆˆ (0, ๐‘0 ). By Lemma 4.1 and Lemma 5.2(ii), ๐ฟ = ๐ฟ(๐‘) is
a supermartingale of class (D). Hence its Doob-Meyer decomposition ๐ฟ =
๐ฟ0 + ๐‘€ ๐ฟ + ๐ด๐ฟ is such that ๐ด๐ฟ is decreasing and nonpositive, and ๐‘€ ๐ฟ is a
instead of indicating
true martingale. Thus
0 โ‰ค ๐ธ[โˆ’๐ด๐ฟ
๐‘‡ ] = ๐ธ[๐ฟ0 โˆ’ ๐ฟ๐‘‡ ] < โˆž.
After localizing as in Lemma 7.5 (with
๐ฟ
Hence sup๐‘ก โˆฃ๐‘€๐‘ก โˆฃ
โ‰ค sup๐‘ก ๐ฟ๐‘ก โˆ’
๐ด๐ฟ
๐‘‡
โˆˆ
๐‘Ÿ = 1),
๐ฟ1 (๐‘ƒ ).
we have
sup๐‘ก ๐ฟ๐‘ก โˆˆ ๐ฟ1 (๐‘ƒ ).
Now (a) follows by the BDG
inequalities exactly as in the proof of Lemma 5.9.
(b) We have ๐ฟโˆ’ (๐‘) โ‰ฅ ๐‘˜1 by Lemma 4.1. Conversely, by Corollary 5.2(i),
๐ฟโˆ’ (๐‘) โ‰ค ๐ถ๐ฟโˆ’ (๐‘0 ) for ๐‘ โˆˆ (0, ๐‘0 ] with some universal constant ๐ถ > 0, and
๐ฟโˆ’ (๐‘0 ) is locally bounded by left-continuity.
(c) We shall use the rate of convergence obtained for ๐ฟ(๐‘) and the
๐ฟ contained in ๐ด๐ฟ via the Bellman BSDE. We may
information about ๐‘
assume by localization that (a) and (b) hold with ๐œŽ๐‘› replaced by ๐‘‡ . Thus
it suces to show that
โˆš
๐‘ ๐ฟ(๐‘) โˆš
lim sup ๐ฟโˆ’ (๐‘)๐œ† +
< โˆž.
๐ฟโˆ’ (๐‘) ๐ฟ2 (๐‘€ )
๐‘โ†’0+
Suppressing again
๐‘
in the notation, (a) and the formula (4.7) for
๐ด๐ฟ
imply
๐ธ[๐ฟ0 โˆ’ ๐ฟ๐‘‡ ] = ๐ธ[โˆ’๐ด๐ฟ
๐‘‡]
โˆซ ๐‘‡
[
] ๐‘ž [โˆซ ๐‘‡
(
(
๐‘ ๐ฟ )โŠค
๐‘ ๐ฟ )]
= ๐ธ (1 โˆ’ ๐‘)
๐ท๐›ฝ ๐ฟ๐‘žโˆ’ ๐‘‘๐œ‡ โˆ’ ๐ธ
.
๐ฟโˆ’ ๐œ† +
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† +
2
๐ฟโˆ’
๐ฟโˆ’
0
0
Recalling that
๐ฟ๐‘‡ = ๐ท๐‘‡ ,
this yields
โˆš
[โˆซ ๐‘‡
(
(
๐‘ ๐ฟ )]
๐‘๐ฟ ๐‘ ๐ฟ )โŠค
1
๐ฟโˆ’ ๐œ† + โˆš 2
= 2๐ธ
๐‘‘โŸจ๐‘€ โŸฉ ๐œ† +
๐ฟโˆ’ ๐œ† +
๐ฟโˆ’
๐ฟโˆ’
๐ฟโˆ’ ๐ฟ (๐‘€ )
0
(
โˆซ ๐‘‡
[
])
๐›ฝ ๐‘ž
1
= โˆฃ๐‘žโˆฃ ๐ธ[๐ฟ0 โˆ’ ๐ฟ๐‘‡ ] โˆ’ ๐ธ (1 โˆ’ ๐‘)
๐ท ๐ฟโˆ’ ๐‘‘๐œ‡
0
(
โˆซ ๐‘‡
[
])
๐›ฝ ๐‘ž
1
= โˆฃ๐‘žโˆฃ ๐ฟ0 โˆ’ ๐ธ ๐ท๐‘‡ + (1 โˆ’ ๐‘)
๐ท ๐ฟโˆ’ ๐‘‘๐œ‡
1
2
0
=
1
โˆฃ๐‘žโˆฃ (๐ฟ0
โˆ’ ฮ“0 ),
โˆซ๐‘‡
ฮ“0 = ฮ“0 (๐‘) = ๐ธ[๐ท๐‘‡ + (1 โˆ’ ๐‘) 0 ๐ท๐›ฝ ๐ฟ๐‘žโˆ’ ๐‘‘๐œ‡]. We know that
[โˆซ๐‘‡
]
ฮ“0 converge to ๐œ‚0 = ๐ธ 0 ๐ท๐‘  ๐œ‡โˆ˜ (๐‘‘๐‘ ) as ๐‘ โ†’ 0+ (and hence
where we have set
๐ฟ0 and
๐‘ž โ†’ 0โˆ’). However,
both
we are asking for the stronger result
1
lim sup โˆฃ๐‘žโˆฃ
โˆฃ๐ฟ0 (๐‘) โˆ’ ฮ“0 (๐‘)โˆฃ < โˆž.
๐‘โ†’0+
V.7 The Limit ๐‘ โ†’ 0
125
๐ฟ0 (๐‘) โ†’ ๐œ‚0 linearly and ฮ“0 (๐‘) โ†’ ๐œ‚0
๐‘ก = 0 read
]1/๐›ฝ+๐‘/๐‘ž0 (
)๐‘ž/๐‘ž0
1โˆ’๐›ฝ /๐›ฝ
.
๐ท๐‘ ๐›ฝ ๐œ‡โˆ˜ (๐‘‘๐‘ )
๐‘˜1 0 ๐ฟ0 (๐‘0 )
By Lemma 7.8(i), it suces to show that
linearly. Using
๐ฟโˆ— = ๐ฟ๐›ฝ ,
inequalities (7.7) and (7.8) evaluated at
๐œ‡โˆ˜ [0, ๐‘‡ ]โˆ’๐‘ ๐œ‚0 โ‰ค ๐ฟ0 (๐‘) โ‰ค ๐ธ
[โˆซ
๐‘‡
0
๐ท, items (ii) and (iii) of Lemma 7.8 yield that
๐ฟ0 (๐‘) โ†’ ๐œ‚0 linearly. The second claim, that ฮ“0 (๐‘) โ†’ ๐œ‚0 linearly, follows from
the denitions of ฮ“0 (๐‘) and ๐œ‚0 using again (2.4) and the uniform bounds for
๐ฟโˆ’ from (b). This ends the proof.
Recalling the bound (2.4) for
V.7.3 Proof of Theorem 3.4 and Other Consequences
Lemma 7.9. Assume that ๐‘† is continuous and that there exists ๐‘0 > 0 such
that ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž. As ๐‘ โ†’ 0,
๐‘ ๐ฟ(๐‘)
๐‘๐œ‚
โ†’
๐ฟโˆ’ (๐‘)
๐œ‚โˆ’
in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) and
1
๐ฟโˆ’ (๐‘)
โˆ™
๐‘ (๐‘) โ†’
1
๐œ‚โˆ’
โˆ™
๐‘๐œ‚
2
.
in โ„‹๐‘™๐‘œ๐‘
(7.9)
For a sequence ๐‘ โ†’ 0โˆ’ the convergence ๐ฟ๐‘โˆ’ (๐‘) โ†’ ๐‘๐œ‚โˆ’ in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ) holds also
without the assumption on ๐‘0 .
Proof. By localization we may assume that ๐ฟโˆ’ (๐‘) is bounded away from zero
๐ฟ(๐‘)
and innity, uniformly in
๐‘
๐œ‚
(Lemma 7.3 and Lemma 4.1 and the preceding
proof ); we also recall (7.2). We have
๐‘ ๐ฟ(๐‘)
) (
) ๐‘ ๐œ‚ ๐‘ ๐œ‚ 1 ( ๐ฟ(๐‘)
โˆ’
๐‘
โˆ’ ๐‘ ๐œ‚ + ๐œ‚โˆ’ โˆ’ ๐ฟโˆ’ (๐‘)
โ‰ค
.
๐ฟโˆ’ (๐‘) ๐œ‚โˆ’
๐ฟโˆ’ (๐‘)
๐ฟโˆ’ (๐‘)๐œ‚โˆ’
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž. The rst part of (7.9) follows from
โˆž convergences obtained in Propositions 7.4,
prelocal โ„›
Let
the
๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ )
and
7.7 and Proposi-
tions 7.1, 7.6, respectively. The proof of the second part of (7.9) is analogous.
๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž and consider a sequence
๐ฟ๐‘ก (๐‘๐‘› ) โ†’ ๐œ‚๐‘ก ๐‘ƒ -a.s. for each ๐‘ก,
rather than the convergence of ๐ฟ๐‘กโˆ’ (๐‘๐‘› ) to ๐œ‚๐‘กโˆ’ . Consider the optional set
โˆฉ
ฮ› := ๐‘› {๐ฟโˆ’ (๐‘๐‘› ) = ๐ฟ(๐‘๐‘› )} โˆฉ {๐œ‚ = ๐œ‚โˆ’ }. Because ๐ฟ(๐‘๐‘› ) and ๐œ‚ are càdlàg,
{๐‘ก : (๐œ”, ๐‘ก) โˆˆ ฮ›๐‘ } โŠ‚ [0, ๐‘‡ ] is countable ๐‘ƒ -a.s. and as ๐‘€ is continuous is follows
โˆซ๐‘‡
๐‘
that
0 1ฮ› ๐‘‘โŸจ๐‘€ โŸฉ = 0 ๐‘ƒ -a.s. Now dominated convergence for stochastic
๐œ‚
๐œ‚
integrals yields that {(๐œ‚โˆ’ โˆ’ ๐ฟโˆ’ (๐‘๐‘› ))๐‘ } โˆ™ ๐‘€ = {(๐œ‚ โˆ’ ๐ฟ(๐‘๐‘› ))1ฮ› ๐‘ } โˆ™ ๐‘€ โ†’ 0
2
in โ„‹๐‘™๐‘œ๐‘ and the rest is as before.
Now drop the assumption that
๐‘๐‘› โ†’ 0โˆ’.
Then Proposition 7.1 only yields
Proof of Theorem 3.4 and Remark 3.5.
The convergence of the optimal con-
sumption is contained in Propositions 7.1 and 7.6 by the formula (4.2). The
convergence of the portfolios follows from Lemma 7.9 in view of (4.8).
๐‘ โˆˆ (0, ๐‘0 ] we have the uniform bound ๐œ…
ห† (๐‘) โ‰ค (๐‘˜2 /๐‘˜1 )๐›ฝ0 by Lemma 4.1
(4.2); while for ๐‘ โˆˆ [๐‘1 , 0), ๐œ…
ห† (๐‘) is prelocally uniformly bounded by
For
and
Lemma 7.3 and (4.2). Hence the convergence of the wealth processes follows
from Corollary 8.4(i).
126
V Risk Aversion Asymptotics
We complement the convergence in the primal problem by a result for
the solution
๐‘Œห† (๐‘)
of the dual problem (4.3).
Proposition 7.10. Assume that ๐‘† is continuous and that there exists ๐‘0 > 0
such that ๐‘ข๐‘0 (๐‘ฅ0 ) < โˆž holds. Moreover, assume that there exists ๐‘1 โˆˆ (0, ๐‘0 ]
such that ๐ฟ(๐‘1 ) is locally bounded. As ๐‘ โ†’ 0,
๐œ‚0 (
1
๐‘Œห† (๐‘) โ†’
โ„ฐ โˆ’๐œ†โˆ™๐‘€ +
๐‘ฅ0
๐œ‚โˆ’
โˆ™
๐‘๐œ‚
)
๐‘Ÿ
in โ„‹๐‘™๐‘œ๐‘
for all ๐‘Ÿ โˆˆ [1, โˆž).
If ๐œ‚ and ๐ฟ(๐‘) are continuous for ๐‘ < 0, the convergence for a limit ๐‘ โ†’ 0โˆ’
holds in the semimartingale topology without the assumptions on ๐‘0 and ๐‘1 .
Proof.
๐ฟ(๐‘1 ) is locally bounded, then ๐ฟ(๐‘) โ†’ ๐œ‚ in โ„›โˆž
๐‘™๐‘œ๐‘ by Remark 7.2
๐ฟ(๐‘) โ†’ ๐‘€ ๐œ‚ in ๐ต๐‘€ ๐‘‚
and Proposition 7.6. Moreover, ๐‘€
๐‘™๐‘œ๐‘ by Propositions 7.4
๐ฟ(๐‘) โ†’ ๐‘ ๐œ‚ in ๐ต๐‘€ ๐‘‚
and 7.7. This implies ๐‘
๐‘™๐‘œ๐‘ by orthogonality of the KW
(i) If
decompositions. It follows that
โˆ’๐œ† โˆ™ ๐‘€ +
1
๐ฟโˆ’ (๐‘)
โˆ™
๐‘ ๐ฟ(๐‘) โ†’ โˆ’๐œ† โˆ™ ๐‘€ +
1
๐œ‚โˆ’
โˆ™
๐‘๐œ‚
in
๐ต๐‘€ ๐‘‚๐‘™๐‘œ๐‘ .
This implies that the corresponding stochastic exponentials converge in
for
๐‘Ÿ โˆˆ [1, โˆž)
(see Theorem 3.4 and Remark 3.7(2) in Protter [63]). In view
of the formula (4.9) for
(ii)
๐‘Ÿ
โ„‹๐‘™๐‘œ๐‘
๐‘Œห† (๐‘),
this ends the proof of the rst claim.
Using Lemma 7.9 and Lemma 8.2(ii), the proof of the second claim
is similar.
Note that in the standard case
sition 7.10 is
โ„ฐ(โˆ’๐œ†
โˆ™
๐‘€ ),
๐ท โ‰ก 1
the normalized limit in Propo-
i.e., the minimal martingale density (cf. [71]).
We conclude by an additional statement concerning the convergence of the
wealth processes in Theorem 3.4.
Proposition 7.11. Let the conditions of Theorem 3.4(ii) hold and assume
in addition that there exists ๐‘1 โˆˆ (0, ๐‘0 ] such that ๐ฟ(๐‘1 ) is locally bounded.
Then the convergence of the wealth processes in Theorem 3.4(ii) takes place
๐‘Ÿ for all ๐‘Ÿ โˆˆ [1, โˆž).
in โ„‹loc
Proof.
Under the additional assumption, the results of this section yield the
๐œ…
ห† (๐‘) in โ„›โˆž
ห† (๐‘) โˆ™ ๐‘€ in ๐ต๐‘€ ๐‘‚๐‘™๐‘œ๐‘
๐‘™๐‘œ๐‘ and the convergence of ๐œ‹
๐œ”
โ„‹๐‘™๐‘œ๐‘ ) by the same formulas as before. Corollary 8.4(ii) yields
convergence of
(and hence in
the claim.
V.8 Appendix A: Convergence of Stochastic Exponentials
This appendix provides some continuity results for stochastic exponentials of
continuous semimartingales in an elementary and self-contained way. They
V.8 Appendix A: Convergence of Stochastic Exponentials
127
are required for the main results of Section V.3 because our wealth processes
are exponentials.
We also use a result from the (much deeper) theory of
โ„‹๐œ” -dierentials; but this is applied only for renements of the main results.
Lemma 8.1. Let ๐‘‹ ๐‘› = ๐‘€ ๐‘› +๐ด๐‘› , ๐‘› โ‰ฅ 1 be continuous semimartingales
with
โˆ‘
๐‘›
continuous canonical โˆซdecompositions and assume that ๐‘› โˆฅ๐‘‹ โˆฅโ„‹2 < โˆž.
Then ๐‘€ ๐‘› , [๐‘€ ๐‘› ] and โˆฃ๐‘‘๐ด๐‘› โˆฃ are locally bounded uniformly in ๐‘›.
Proof.
๐‘€๐‘ก๐‘›โ˜…
๐œŽ๐‘˜ = inf{๐‘ก > 0 : sup๐‘› โˆฃ๐‘€๐‘ก๐‘› โˆฃ > ๐‘˜} โˆง ๐‘‡ . We use the notation
= sup๐‘ โ‰ค๐‘ก โˆฃ๐‘€๐‘ ๐‘› โˆฃ, then the norms โˆฅ๐‘€๐‘‡๐‘›โ˜… โˆฅ๐ฟ2 and โˆฅ๐‘€ ๐‘› โˆฅโ„‹2 are equivalent
Let
by the BDG inequalities. Now
[
]
โˆ‘
๐‘ƒ sup ๐‘€๐‘‡๐‘›โ˜… > ๐‘˜ โ‰ค ๐‘˜ โˆ’2
โˆฅ๐‘€๐‘‡๐‘›โ˜… โˆฅ๐ฟ2
๐‘›
๐‘›
[
]
โˆ‘
๐‘ƒ [๐œŽ๐‘˜ < ๐‘‡ ] โ†’ 0. Similarly, ๐‘ƒ sup๐‘› [๐‘€ ๐‘› ]๐‘‡ > ๐‘˜ โ‰ค ๐‘˜ โˆ’1 ๐‘› โˆฅ๐‘€ ๐‘› โˆฅโ„‹2
[
]
โˆซ๐‘‡
โˆ‘
๐‘ƒ sup๐‘› 0 โˆฃ๐‘‘๐ด๐‘› โˆฃ > ๐‘˜ โ‰ค ๐‘˜ โˆ’2 ๐‘› โˆฅ๐ด๐‘› โˆฅโ„‹2 yield the other claims.
shows
and
We sometimes write in
๐’ฎ 0
to indicate convergence in the semimartingale
topology.
Lemma 8.2. Let ๐‘‹ ๐‘› = ๐‘€ ๐‘› + ๐ด๐‘› , ๐‘› โ‰ฅ 1 and ๐‘‹ = ๐‘€ + ๐ด be continuous
semimartingales with continuous canonical decompositions.
โˆ‘
2 .
(i) ๐‘› โˆฅ๐‘‹ ๐‘› โˆ’ ๐‘‹โˆฅโ„‹2 < โˆž implies โ„ฐ(๐‘‹ ๐‘› ) โ†’ โ„ฐ(๐‘‹) in โ„‹๐‘™๐‘œ๐‘
2 implies โ„ฐ(๐‘‹ ๐‘› ) โ†’ โ„ฐ(๐‘‹) in ๐’ฎ 0 .
(ii) ๐‘‹ ๐‘› โ†’ ๐‘‹ in โ„‹๐‘™๐‘œ๐‘
(iii) ๐‘‹ ๐‘› โ†’ ๐‘‹ in ๐’ฎ 0 implies โ„ฐ(๐‘‹ ๐‘› ) โ†’ โ„ฐ(๐‘‹) in ๐’ฎ 0 .
Proof.
(i)
By localization we may assume that
๐‘€
and
โˆซ
โˆฃ๐‘‘๐ดโˆฃ
are bounded
๐‘›
and, by Lemma 8.1, that โˆฃ๐‘€ โˆฃ and
โˆฃ๐‘‘๐ด๐‘› โˆฃ are bounded by a constant ๐ถ
๐‘›
2
๐‘›
independent of ๐‘›. Note that ๐‘‹ โ†’ ๐‘‹ in โ„‹ ; we shall show โ„ฐ(๐‘‹ ) โ†’ โ„ฐ(๐‘‹)
2
in โ„‹ . Since this is a metric space, no loss of generality is entailed by passing
โˆซ
to a subsequence. Doing so, we have
uniformly in time,
๐‘ƒ -a.s.
๐‘€ ๐‘› โ†’ ๐‘€ , [๐‘€ ๐‘› ] โ†’ [๐‘€ ],
and
๐ด๐‘› โ†’ ๐ด
In view of the uniform bound
(
)
๐‘Œ ๐‘› := โ„ฐ(๐‘‹ ๐‘› ) = exp ๐‘‹ ๐‘› โˆ’ 21 [๐‘€ ๐‘› ] โ‰ค ๐‘’2๐ถ
we conclude that
of the stochastic
โˆฅ๐‘Œ
โˆ™
๐‘Œ ๐‘› โ†’ ๐‘Œ := โ„ฐ(๐‘‹) = exp(๐‘‹ โˆ’ 12 [๐‘€ ]) in โ„›2 . By denition
๐‘› = ๐‘Œ โˆ™ ๐‘‹ โˆ’ ๐‘Œ ๐‘› โˆ™ ๐‘‹ ๐‘› , where
exponential we have ๐‘Œ โˆ’ ๐‘Œ
๐‘‹ โˆ’ ๐‘Œ ๐‘› โˆ™ ๐‘‹ ๐‘› โˆฅโ„‹2 โ‰ค โˆฅ(๐‘Œ โˆ’ ๐‘Œ ๐‘› ) โˆ™ ๐‘‹โˆฅโ„‹2 + โˆฅ๐‘Œ ๐‘› โˆ™ (๐‘‹ โˆ’ ๐‘‹ ๐‘› )โˆฅโ„‹2 .
The rst norm tends to zero by dominated convergence for stochastic inte-
โˆฃ๐‘Œ ๐‘› โˆฃ โ‰ค ๐‘’2๐ถ and ๐‘‹ ๐‘› โ†’ ๐‘‹ in โ„‹2 .
๐‘›
Consider a subsequence of (๐‘‹ ). After passing to another subse2
(i) shows the convergence in โ„‹๐‘™๐‘œ๐‘ and Proposition 2.2 yields (ii).
grals and for the second we use that
(ii)
quence,
(iii) This follows from (ii) by using Proposition 2.2 twice.
128
V Risk Aversion Asymptotics
We return to the semimartingale
๐‘…
of asset returns, which is assumed to
be continuous in the sequel. We recall the structure condition (2.6) and dene
๐ฟ๐œ” (๐‘€ ) := {๐œ‹ โˆˆ ๐ฟ(๐‘€ ) : โˆฅ๐œ‹โˆฅ๐ฟ๐œ” (๐‘€ ) < โˆž}, where โˆฅ๐œ‹โˆฅ๐ฟ๐œ” (๐‘€ ) := โˆฅ๐œ‹
๐œ”
and โ„‹ was introduced at the end of Section V.2.2.
โˆ™
๐‘€ โˆฅโ„‹ ๐œ”
Lemma 8.3. Let ๐‘… be continuous, ๐‘Ÿ โˆˆ {2, ๐œ”}, and ๐œ‹, ๐œ‹๐‘› โˆˆ ๐ฟ๐‘Ÿ๐‘™๐‘œ๐‘ (๐‘€ ). Then
๐‘Ÿ .
๐œ‹ ๐‘› โ†’ ๐œ‹ in ๐ฟ๐‘Ÿ๐‘™๐‘œ๐‘ (๐‘€ ) if and only if ๐œ‹ ๐‘› โˆ™ ๐‘… โ†’ ๐œ‹ โˆ™ ๐‘… in โ„‹๐‘™๐‘œ๐‘
Proof.
By (2.6) we have
โˆซ
๐œ‹ โˆ™ ๐‘… = ๐œ‹ โˆ™ ๐‘€ + ๐œ‹ โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†.
Let
๐œ’ :=
โˆซ
๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†
denote the mean-variance tradeo process. The inequality
๐ธ
๐‘‡
[( โˆซ
)2 ]
[( โˆซ
โˆฃ๐œ‹ ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†โˆฃ
โ‰ค๐ธ
โŠค
๐‘‡
โŠค
๐œ‹ ๐‘‘โŸจ๐‘€ โŸฉ ๐œ‹
)( โˆซ
0
0
๐‘‡
๐œ†โŠค ๐‘‘โŸจ๐‘€ โŸฉ ๐œ†
)]
0
โˆฅ๐œ‹ โˆ™ ๐‘€ โˆฅโ„‹2 โ‰ค โˆฅ๐œ‹ โˆ™ ๐‘…โˆฅโ„‹2 โ‰ค (1 + โˆฅ๐œ’๐‘‡ โˆฅ๐ฟโˆž )โˆฅ๐œ‹ โˆ™ ๐‘€ โˆฅโ„‹2 .
bounded due to continuity, this yields the result for ๐‘Ÿ = 2.
๐‘Ÿ = ๐œ” is analogous.
implies
As
๐œ’
is locally
The proof for
Corollary 8.4. Let ๐‘… be continuous and (๐œ‹, ๐œ…), (๐œ‹๐‘› , ๐œ…๐‘› ) โˆˆ ๐’œ.
(i) Assume that ๐œ‹๐‘› โ†’ ๐œ‹ in ๐ฟ2๐‘™๐‘œ๐‘ (๐‘€ ), that (๐œ…๐‘› ) is prelocally bounded uniformly in ๐‘›, and that ๐œ…๐‘›๐‘ก โ†’ ๐œ…๐‘ก ๐‘ƒ -a.s. for each ๐‘ก โˆˆ [0, ๐‘‡ ]. Then
๐‘‹(๐œ‹ ๐‘› , ๐œ…๐‘› ) โ†’ ๐‘‹(๐œ‹, ๐œ…) in the semimartingale topology.
๐‘› ๐‘›
(ii) Assume ๐œ‹๐‘› โ†’ ๐œ‹ in ๐ฟ๐œ”๐‘™๐‘œ๐‘ (๐‘€ ) and ๐œ…๐‘› โ†’ ๐œ… in โ„›โˆž
๐‘™๐‘œ๐‘ . Then ๐‘‹(๐œ‹ , ๐œ… ) โ†’
๐‘Ÿ for all ๐‘Ÿ โˆˆ [1, โˆž).
๐‘‹(๐œ‹, ๐œ…) in โ„‹๐‘™๐‘œ๐‘
Proof.
๐œ‡(๐‘‘๐‘ )๐‘ก = ๐œ…๐‘›๐‘  โˆ™ ๐œ‡(๐‘‘๐‘ )๐‘กโˆ’ for all ๐‘ก. After
โˆซ๐‘‡ ๐‘›
localization, bounded convergence yields
0 โˆฃ๐œ…๐‘ก โˆ’ ๐œ…๐‘ก โˆฃ ๐œ‡(๐‘‘๐‘ก) โ†’ 0 ๐‘ƒ -a.s. and
2
๐‘›
โˆ™
in ๐ฟ (๐‘ƒ ). Using Lemma 8.3, we have ๐œ‹
๐‘… + ๐œ…๐‘› โˆ™ ๐œ‡(๐‘‘๐‘ก) โ†’ ๐œ‹ โˆ™ ๐‘… + ๐œ… โˆ™ ๐œ‡(๐‘‘๐‘ก)
2
in โ„‹๐‘™๐‘œ๐‘ . In view of (2.2) we conclude by Lemma 8.2(ii).
๐‘›
๐‘›
(ii) With Lemma 8.3 we obtain ๐œ‹ โˆ™ ๐‘… + ๐œ… โˆ™ ๐œ‡(๐‘‘๐‘ก) โ†’ ๐œ‹ โˆ™ ๐‘… + ๐œ… โˆ™ ๐œ‡(๐‘‘๐‘ก)
๐œ”
๐‘Ÿ
in โ„‹๐‘™๐‘œ๐‘ . Thus the stochastic exponentials converge in โ„‹๐‘™๐‘œ๐‘ for all ๐‘Ÿ โˆˆ [1, โˆž)
(i)
By continuity of
๐œ‡, ๐œ…๐‘›๐‘ 
โˆ™
(see Theorem 3.4 and Remark 3.7(2) in [63]).
V.9 Appendix B: Proof of Lemma 6.12
In this section we give the proof of Lemma 6.12. As mentioned above, the
argument is adapted from the Brownian setting of [48, Proposition 2.4].
We use the notation introduced before Lemma 6.12, in particular, re-
๐‘˜ throughout and let ๐œ := ๐œ๐‘˜ . For xed integers ๐‘š โ‰ฅ ๐‘›
๐›ฟ๐ฟ = ๐ฟ๐‘› โˆ’ ๐ฟ๐‘š , moreover, ๐›ฟ๐‘€ , ๐›ฟ๐‘ , ๐›ฟ๐‘ have the analogous
meaning. Note that ๐›ฟ๐ฟ โ‰ฅ 0 as ๐‘š โ‰ฅ ๐‘›. The technique consists in applying
Itô's formula to ฮฆ(๐›ฟ๐ฟ), where, with ๐พ := 6๐ถ๐‘˜ ,
call (6.8). We x
we abbreviate
ฮฆ(๐‘ฅ) =
)
1 ( 4๐พ๐‘ฅ
๐‘’
โˆ’ 4๐พ๐‘ฅ โˆ’ 1 .
2
8๐พ
V.9 Appendix B: Proof of Lemma 6.12
On
โ„+
129
this function satises
ฮฆ(0) = ฮฆโ€ฒ (0) = 0,
Moreover,
ฮฆโ€ฒโ€ฒ โ‰ฅ 0
ฮฆโ€ฒ โ‰ฅ 0,
ฮฆ โ‰ฅ 0,
and hence
โ„Ž(๐‘ฅ) :=
1 โ€ฒโ€ฒ
2 ฮฆ (๐‘ฅ)
1 โ€ฒโ€ฒ
2ฮฆ
โˆ’ 2๐พฮฆโ€ฒ โ‰ก 1.
โˆ’ ๐พฮฆโ€ฒ (๐‘ฅ) = 1 + ๐พฮฆโ€ฒ (๐‘ฅ)
is
nonnegative and nondecreasing.
(i) By Itô's formula we have
๐œ
[
]
๐‘š
ฮฆ(๐›ฟ๐ฟ0 ) = ฮฆ(๐›ฟ๐ฟ๐œ ) โˆ’
ฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  ) ๐‘“ ๐‘› (๐‘ , ๐ฟ๐‘›๐‘  , ๐‘๐‘ ๐‘› ) โˆ’ ๐‘“ ๐‘š (๐‘ , ๐ฟ๐‘š
๐‘  , ๐‘๐‘  ) ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
0
โˆซ ๐œ
โˆซ ๐œ
1 โ€ฒโ€ฒ
โˆ’
ฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  ) ๐‘‘๐›ฟ๐‘€๐‘  .
2 ฮฆ (๐›ฟ๐ฟ๐‘  ) ๐‘‘ โŸจ๐›ฟ๐‘€ โŸฉ๐‘  โˆ’
โˆซ
0
0
By elementary inequalities we have for all
๐‘š
and
๐‘›
that
(
)
หœ 2 + โˆฃ๐‘โˆฃ
หœ 2 ๐œ,
โˆฃ๐‘“ ๐‘› (๐‘ก, ๐ฟ๐‘› , ๐‘ ๐‘› ) โˆ’ ๐‘“ ๐‘š (๐‘ก, ๐ฟ๐‘š , ๐‘ ๐‘š )โˆฃ๐œ โ‰ค ๐œ‰ + ๐พ โˆฃ๐‘ ๐‘› โˆ’ ๐‘ ๐‘š โˆฃ2 + โˆฃ๐‘ ๐‘› โˆ’ ๐‘โˆฃ
where the index
๐‘ก
was omitted. Hence
[
๐œ
)]
(
หœ๐‘  โˆฃ2 + โˆฃ๐‘
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
ฮฆ(๐›ฟ๐ฟ0 ) โ‰ค ฮฆ(๐›ฟ๐ฟ๐œ ) +
ฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  ) ๐œ‰๐‘  + ๐พ โˆฃ๐›ฟ๐‘๐‘  โˆฃ2 + โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘
0
โˆซ ๐œ
โˆซ ๐œ
1 โ€ฒโ€ฒ
โˆ’
ฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  ) ๐‘‘๐›ฟ๐‘€๐‘  .
2 ฮฆ (๐›ฟ๐ฟ๐‘  ) ๐‘‘ โŸจ๐›ฟ๐‘€ โŸฉ๐‘  โˆ’
โˆซ
0
0
The expectation of the stochastic integral vanishes since
๐›ฟ๐ฟ
is bounded and
โ„‹2 . We deduce
๐›ฟ๐‘€ โˆˆ
โˆซ ๐œ
โˆซ ๐œ
[ 1 โ€ฒโ€ฒ
]
โ€ฒ
2
1 โ€ฒโ€ฒ
๐ธ
2 ฮฆ (๐›ฟ๐ฟ๐‘  ) โˆ’ ๐พฮฆ (๐›ฟ๐ฟ๐‘  ) โˆฃ๐›ฟ๐‘๐‘  โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘  + ๐ธ
2 ฮฆ (๐›ฟ๐ฟ๐‘  ) ๐‘‘โŸจ๐›ฟ๐‘ โŸฉ๐‘ 
0
0
โˆซ ๐œ
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  + ฮฆ(๐›ฟ๐ฟ0 )
โˆ’๐ธ
๐พฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  )โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘
0
โˆซ ๐œ
[
]
[
]
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  .
โ‰ค ๐ธ ฮฆ(๐›ฟ๐ฟ๐œ ) + ๐ธ
ฮฆโ€ฒ (๐›ฟ๐ฟ๐‘  ) ๐œ‰๐‘  + ๐พโˆฃ๐‘
(9.1)
(9.2)
(9.3)
0
We let
for all
๐‘›
หœ
๐‘š tend to innity, then ๐›ฟ๐ฟ๐‘ก = ๐ฟ๐‘›๐‘ก โˆ’ ๐ฟ๐‘š
๐‘ก converges to ๐ฟ๐‘ก โˆ’ ๐ฟ๐‘ก ๐‘ƒ -a.s.
๐‘ก and with a uniform bound, so (9.3) converges to
โˆซ ๐œ
[
]
[
]
หœ๐œ ) + ๐ธ
หœ ๐‘  ) ๐œ‰๐‘  + ๐พโˆฃ๐‘
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  ;
๐ธ ฮฆ(๐ฟ๐‘›๐œ โˆ’ ๐ฟ
ฮฆโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
0
while (9.2) converges to
โˆซ
โˆ’๐ธ
๐œ
หœ ๐‘  )โˆฃ๐‘ ๐‘› โˆ’ ๐‘
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  + ฮฆ(๐ฟ๐‘› โˆ’ ๐ฟ
หœ 0 ).
๐พฮฆโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
๐‘ 
0
0
We turn to (9.1). The continuous function
in the rst integrand.
We recall that
โ„Ž
โ„Ž(๐‘ฅ) = 12 ฮฆโ€ฒโ€ฒ (๐‘ฅ) โˆ’ ๐พฮฆโ€ฒ (๐‘ฅ)
occurs
is nonnegative and nondecreasing
V Risk Aversion Asymptotics
130
ฮฆโ€ฒโ€ฒ
๐‘š,
and note that
decreasing in
has the same properties.
๐‘› หœ
โ„Ž(๐›ฟ๐ฟ๐‘  ) = โ„Ž(๐ฟ๐‘›๐‘  โˆ’๐ฟ๐‘š
๐‘  ) โ†‘ โ„Ž(๐ฟ๐‘  โˆ’ ๐ฟ๐‘  );
Moreover, as
๐ฟ๐‘š
๐‘ก
is monotone
โ€ฒโ€ฒ
๐‘› หœ
ฮฆโ€ฒโ€ฒ (๐›ฟ๐ฟ๐‘  ) = ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’๐ฟ๐‘š
๐‘  ) โ†‘ ฮฆ (๐ฟ๐‘  โˆ’ ๐ฟ๐‘  )
๐‘ƒ -a.s. for all ๐‘ . Hence we have for any xed ๐‘š0 โ‰ค ๐‘š that
โˆซ ๐œ
โˆซ ๐œ
๐‘›
๐‘š
๐‘›
๐‘š
0
โ„Ž(๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
โ„Ž(๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
)โˆฃ๐‘
โˆ’
๐‘
โˆฃ
๐‘‘โŸจ๐‘€
โŸฉ
โ‰ฅ
๐ธ
๐ธ
๐‘ 
๐‘  )โˆฃ๐‘๐‘  โˆ’ ๐‘๐‘  โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘  ;
๐‘ 
๐‘ 
๐‘ 
0
โˆซ0 ๐œ
โˆซ ๐œ
๐‘›
๐‘š
๐‘›
๐‘š
0
ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
๐ธ
๐‘  ) ๐‘‘โŸจ๐‘ โˆ’ ๐‘ โŸฉ๐‘  .
๐‘  ) ๐‘‘โŸจ๐‘ โˆ’ ๐‘ โŸฉ๐‘  โ‰ฅ ๐ธ
0
0
The right hand sides are convex lower semicontinuous functions of
๐ฟ2 (๐‘€ )
๐‘ ๐‘š โˆˆ โ„‹2 ,
and
respectively, hence also weakly lower semicontinuous.
We conclude from the weak convergences
๐‘ ๐‘š โ†’ ๐‘หœ
and
หœ
๐‘๐‘š โ†’ ๐‘
that
๐œ
โˆซ
๐‘›
หœ๐‘š
โ„Ž(๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
๐‘  )โˆฃ๐‘๐‘  โˆ’ ๐‘๐‘  โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
โˆซ ๐œ
๐‘›
0
หœ
โ‰ฅ๐ธ
โ„Ž(๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
๐‘  )โˆฃ๐‘๐‘  โˆ’ ๐‘๐‘  โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘  ;
lim inf ๐ธ
๐‘šโ†’โˆž
๐‘๐‘š โˆˆ
0
0
๐œ
โˆซ
lim inf ๐ธ
๐‘šโ†’โˆž
๐‘›
๐‘š
ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
๐‘  ) ๐‘‘โŸจ๐‘ โˆ’ ๐‘ โŸฉ๐‘  โ‰ฅ ๐ธ
0
๐‘š0 .
for all
โˆซ
๐œ
๐‘›
0
หœ
ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ๐‘š
๐‘  ) ๐‘‘โŸจ๐‘ โˆ’ ๐‘ โŸฉ๐‘ 
0
We can now let
๐‘š0 tend toโˆซinnity, then by monotone convergence
๐œ
หœ๐‘  โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘  and the
หœ ๐‘  )โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘
๐ธ 0 โ„Ž(๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
the rst right hand side tends to
second one tends to
๐œ
โˆซ
๐ธ
หœ ๐‘  ) ๐‘‘โŸจ๐‘ ๐‘› โˆ’ ๐‘
หœ โŸฉ๐‘  โ‰ฅ 2๐ธ
ฮฆโ€ฒโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
โˆซ
๐œ
[
]
หœ โŸฉ๐‘  = 2๐ธ โŸจ๐‘ ๐‘› โˆ’ ๐‘
หœ โŸฉ๐œ ,
๐‘‘โŸจ๐‘ ๐‘› โˆ’ ๐‘
0
0
where we have used that
หœ โ‰ฅ 0
๐ฟ๐‘› โˆ’ ๐ฟ
and
ฮฆโ€ฒโ€ฒ (๐‘ฅ) = 2๐‘’4๐พ๐‘ฅ โ‰ฅ 2
for
๐‘ฅ โ‰ฅ 0.
Altogether, we have passed from (9.1)(9.3) to
โˆซ
๐ธ
๐œ
(
)
[
]
หœ ๐‘  ) โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘
หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  + ๐ธ โŸจ๐‘ ๐‘› โˆ’ ๐‘
หœ โŸฉ๐œ
โˆ’ 2๐พฮฆโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
0
โˆซ ๐œ
[
]
๐‘›
๐‘›
หœ
หœ
หœ ๐‘  ) ๐œ‰๐‘  + ๐พโˆฃ๐‘หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  .
โ‰ค ๐ธฮฆ(๐ฟ๐œ โˆ’ ๐ฟ๐œ ) โˆ’ ฮฆ(๐ฟ0 โˆ’ ๐ฟ0 ) + ๐ธ
ฮฆโ€ฒ (๐ฟ๐‘›๐‘  โˆ’ ๐ฟ
1 โ€ฒโ€ฒ
2ฮฆ
0
As
1 โ€ฒโ€ฒ
2ฮฆ
we let
๐‘›
โˆ’ 2๐พฮฆโ€ฒ โ‰ก 1,
the rst integral reduces to
๐ธ
โˆซ๐œ
0
โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  .
If
tend to innity, the right hand side converges to zero by dominated
convergence, so that we conclude
โˆซ
๐ธ
0
as claimed.
๐œ
โˆฃ๐‘๐‘ ๐‘› โˆ’ ๐‘หœ๐‘  โˆฃ2 ๐‘‘โŸจ๐‘€ โŸฉ๐‘  โ†’ 0;
[
]
หœ โŸฉ๐œ โ†’ 0
๐ธ โŸจ๐‘ ๐‘› โˆ’ ๐‘
V.9 Appendix B: Proof of Lemma 6.12
(ii) For all
โˆฃ๐ฟ๐‘›๐‘กโˆง๐œ
โˆ’
๐ฟ๐‘š
๐‘กโˆง๐œ โˆฃ
The sequence
๐‘š
๐‘›
and
131
we have
โˆซ
๐œ
๐‘š
โˆฃ๐‘“ ๐‘› (๐‘ , ๐ฟ๐‘›๐‘  , ๐‘๐‘ ๐‘› ) โˆ’ ๐‘“ ๐‘š (๐‘ , ๐ฟ๐‘š
๐‘  , ๐‘๐‘  )โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
๐‘กโˆง๐œ
๐‘›
๐‘š (9.4)
โˆ’ ๐‘€๐‘กโˆง๐œ
).
+ (๐‘€๐œ๐‘› โˆ’ ๐‘€๐œ๐‘š ) โˆ’ (๐‘€๐‘กโˆง๐œ
โ‰ค
โˆฃ๐ฟ๐‘›๐œ
โˆ’
๐ฟ๐‘š
๐œ โˆฃ
๐‘€ ๐‘š = ๐‘๐‘š
subsequence, still denoted
+
๐‘€ + ๐‘ ๐‘š is Cauchy in โ„‹๐œ2 . We pick a fast
๐‘š
๐‘š โˆ’ ๐‘€ ๐‘š+1 โˆฅ
โˆ’๐‘š .
by ๐‘€ , such that โˆฅ๐‘€
โ„‹๐œ2 โ‰ค 2
โˆ™
This implies that
๐‘€ โˆ— := sup โˆฃ๐‘€ ๐‘š โˆฃ โˆˆ โ„‹๐œ2 ;
๐‘ โˆ— := sup โˆฃ๐‘ ๐‘š โˆฃ โˆˆ ๐ฟ2๐œ (๐‘€ )
๐‘š
๐‘š
๐‘š
หœ. Therefore, lim๐‘› ๐‘“ ๐‘š (๐‘ก, ๐ฟ๐‘š
๐‘ ๐‘š converges ๐‘ƒ โŠ—โŸจ๐‘€ ๐œ โŸฉ-a.e. to ๐‘
๐‘ก , ๐‘๐‘ก ) =
หœ ๐‘ก , ๐‘หœ๐‘ก ) ๐‘ƒ โŠ—โŸจ๐‘€ ๐œ โŸฉ-a.e. Moreover, โˆฃ๐‘“ ๐‘š (๐‘ก, ๐ฟ๐‘š , ๐‘ ๐‘š )๐œ โˆฃ โ‰ค ๐œ‰๐‘ก +๐ถโˆฃ๐‘ โˆ— โˆฃ2 and this
๐‘“ (๐‘ก, ๐ฟ
๐‘ก
๐‘ก
๐‘ก
1
bound is in ๐ฟ๐œ (๐‘€ ). Passing to a subsequence if necessary, we have
โˆซ ๐œ
๐‘š
lim
โˆฃ๐‘“ ๐‘› (๐‘ , ๐ฟ๐‘›๐‘  , ๐‘๐‘ ๐‘› ) โˆ’ ๐‘“ ๐‘š (๐‘ , ๐ฟ๐‘š
๐‘  , ๐‘๐‘  )โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
๐‘šโ†’โˆž 0
โˆซ ๐œ
หœ ๐‘  , ๐‘หœ๐‘  )โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘  ๐‘ƒ -a.s.
โˆฃ๐‘“ ๐‘› (๐‘ , ๐ฟ๐‘›๐‘  , ๐‘๐‘ ๐‘› ) โˆ’ ๐‘“ (๐‘ , ๐ฟ
=
and that
0
As
]
[
๐‘š โˆ’๐‘€
หœ๐‘กโˆง๐œ โˆฃ โ†’ 0 and, after
โ„‹๐œ2 , we have ๐ธ sup๐‘กโ‰ค๐‘‡ โˆฃ๐‘€๐‘กโˆง๐œ
๐‘š
หœ๐‘กโˆง๐œ โˆฃ โ†’ 0 ๐‘ƒ -a.s. We can now take
subsequence, sup๐‘กโ‰ค๐‘‡ โˆฃ๐‘€๐‘กโˆง๐œ โˆ’ ๐‘€
หœ
๐‘€๐‘š โ†’ ๐‘€
picking a
๐‘šโ†’โˆž
in
in (9.4) to obtain
sup โˆฃ๐ฟ๐‘›๐‘กโˆง๐œ
๐‘กโ‰ค๐‘‡
หœ ๐‘กโˆง๐œ โˆฃ โ‰ค โˆฃ๐ฟ๐‘›๐œ โˆ’ ๐ฟ
หœ๐œ โˆฃ +
โˆ’๐ฟ
โˆซ
๐œ
หœ๐‘  , ๐ฟ
หœ ๐‘  )โˆฃ ๐‘‘โŸจ๐‘€ โŸฉ๐‘ 
โˆฃ๐‘“ ๐‘› (๐‘ , ๐ฟ๐‘›๐‘  , ๐‘๐‘ ๐‘› ) โˆ’ ๐‘“ (๐‘ , ๐ฟ
0
หœ๐œ ) โˆ’ (๐‘€ ๐‘› โˆ’ ๐‘€
หœ๐‘กโˆง๐œ ).
+ sup (๐‘€๐œ๐‘› โˆ’ ๐‘€
๐‘กโˆง๐œ
๐‘กโ‰ค๐‘‡
With exactly the same arguments, extracting another subsequence if necessary, the right hand side converges to zero
shown that
หœ ๐‘กโˆง๐œ โˆฃ = 0,
lim๐‘› sup๐‘กโ‰ค๐‘‡ โˆฃ๐ฟ๐‘›๐‘กโˆง๐œ โˆ’ ๐ฟ
๐‘ƒ -a.s.
as
๐‘› โ†’ โˆž.
along a subsequence.
monotonicity, we obtain the result for the whole sequence.
We have
But by
โ–ก
132
V Risk Aversion Asymptotics
Bibliography
[1] C. D. Aliprantis and K. C. Border. Innite Dimensional Analysis: A Hitchhiker's Guide. Springer, Berlin, 3rd edition, 2006.
[2] J. P. Ansel and C. Stricker. Décomposition de Kunita-Watanabe. In Séminaire
de Probabilités XXVII, volume 1557 of Lecture Notes in Math., pages 3032,
Springer, Berlin, 1993.
[3] O. E. Barndor-Nielsen and N. Shephard. Non-Gaussian Ornstein-Uhlenbeckbased models and some of their uses in nancial economics. J. R. Stat. Soc.
Ser. B Stat. Methodol., 63(2):167241, 2001.
[4] D. Becherer. Bounded solutions to backward SDE's with jumps for utility
optimization and indierence hedging. Ann. Appl. Probab., 16(4):20272054,
2006.
[5] C. Bender and C. R. Niethammer. On ๐‘ž -optimal martingale measures in
exponential Lévy models. Finance Stoch., 12(3):381410, 2008.
[6] F. E. Benth, K. H. Karlsen, and K. Reikvam. Optimal portfolio management
rules in a non-Gaussian market with durability and intertemporal substitution.
Finance Stoch., 5(4):447467, 2001.
[7] S. Biagini and M. Frittelli. Utility maximization in incomplete markets for
unbounded processes. Finance Stoch., 9(4):493517, 2005.
[8] S. Biagini and M. Frittelli. The supermartingale property of the optimal wealth
process for general semimartingales. Finance Stoch., 11(2):253266, 2007.
[9] P. Briand and Y. Hu. Quadratic BSDEs with convex generators and unbounded terminal conditions. Probab. Theory Related Fields, 141:543567,
2008.
[10] L. Carassus and M. Rásonyi. Convergence of utility indierence prices to the
superreplication price. Math. Methods Oper. Res., 64(1):145154, 2006.
[11] A. ยƒerný and J. Kallsen. On the structure of general mean-variance hedging
strategies. Ann. Probab., 35(4):14791531, 2007.
[12] T. Choulli and C. Stricker. Comparing the minimal Hellinger martingale measure of order ๐‘ž to the ๐‘ž -optimal martingale measure. Stochastic Process. Appl.,
119(4):13681385, 2009.
[13] C. Czichowsky and M. Schweizer. On the Markowitz problem under convex
cone constraints. In preparation.
[14] C. Czichowsky and M. Schweizer. Closedness in the semimartingale topology
for spaces of stochastic integrals with constrained integrands. To appear in
Sém. Probab., 2009.
[15] C. Czichowsky, N. Westray, and H. Zheng. Convergence in the semimartingale
134
BIBLIOGRAPHY
topology and constrained portfolios. To appear in Sém. Probab., 2008.
[16] F. Delbaen, P. Monat, W. Schachermayer, M. Schweizer, and C. Stricker.
Weighted norm inequalities and hedging in incomplete markets. Finance
Stoch., 1(3):181227, 1997.
[17] F. Delbaen and W. Schachermayer. The fundamental theorem of asset pricing
for unbounded stochastic processes. Math. Ann., 312:215250, 1998.
[18] C. Dellacherie and P. A. Meyer. Probabilities and Potential B. North Holland,
Amsterdam, 1982.
[19] C. Doléans-Dade and P. A. Meyer. Inégalités de normes avec poids. In Séminaire de Probabilités XIII (1977/78), volume 721 of Lecture Notes in Math.,
pages 313331. Springer, Berlin, 1979.
[20] M. Émery. Une topologie sur l'espace des semimartingales. In Séminaire
de Probabilités XIII, volume 721 of Lecture Notes in Math., pages 260280,
Springer, Berlin, 1979.
[21] W. H. Fleming and H. M. Soner. Controlled Markov Processes and Viscosity
Solutions. Springer, New York, 2nd edition, 2006.
[22] H. Föllmer and A. Schied. Stochastic Finance: An Introduction in Discrete
Time. W. de Gruyter, Berlin, 2nd edition, 2004.
[23] C. Frei and M. Schweizer. Exponential utility indierence valuation in a general semimartingale model. In F. Delbaen, M. Rásonyi, and C. Stricker, editors, Optimality and Risk - Modern Trends in Mathematical Finance. The
Kabanov Festschrift, pages 4986. Springer, 2009.
[24] T. Goll and J. Kallsen. Optimal portfolios for logarithmic utility. Stochastic
Process. Appl., 89(1):3148, 2000.
[25] T. Goll and J. Kallsen. A complete explicit solution to the log-optimal portfolio
problem. Ann. Appl. Probab., 13(2):774799, 2003.
[26] T. Goll and L. Rüschendorf. Minimax and minimal distance martingale
measures and their relationship to portfolio optimization. Finance Stoch.,
5(4):557581, 2001.
[27] C. Gollier. The Economics of Risk and Time. MIT Press, Cambridge, 2001.
[28] P. Grandits. On martingale measures for stochastic processes with independent
increments. Theory Probab. Appl., 44(1):3950, 2000.
[29] P. Grandits and T. Rheinländer. On the minimal entropy martingale measure.
Ann. Probab., 30(3):10031038, 2002.
[30] P. Grandits and C. Summer. Risk averse asymptotics and the optional decomposition. Theory Probab. Appl., 51(2):325334, 2007.
[31] M. Grasselli. A stability result for the HARA class with stochastic interest
rates. Insurance Math. Econom., 33(3):611627, 2003.
[32] J. M. Harrison and S. R. Pliska. Martingales and stochastic integrals in the
theory of continuous trading. Stochastic Process. Appl., 11(3):215260, 1981.
[33] Y. Hu, P. Imkeller, and M. Müller. Utility maximization in incomplete markets. Ann. Appl. Probab., 15(3):16911712, 2005.
[34] J. Jacod and A. N. Shiryaev. Limit Theorems for Stochastic Processes.
Springer, Berlin, 2nd edition, 2003.
[35] M. Jeanblanc, S. Klöppel, and Y. Miyahara. Minimal ๐‘“ ๐‘ž -martingale measures
for exponential Lévy processes. Ann. Appl. Probab., 17(5/6):16151638, 2007.
BIBLIOGRAPHY
135
[36] E. Jouini and C. Napp. Convergence of utility functions and convergence of
optimal strategies. Finance Stoch., 8(1):133144, 2004.
[37] Yu. Kabanov and C. Stricker. On the optimal portfolio for the exponential utility maximization: Remarks to the six-author paper. Math. Finance.,
12:125134, 2002.
[38] J. Kallsen. Optimal portfolios for exponential Lévy processes. Math. Methods
Oper. Res., 51(3):357374, 2000.
[39] J. Kallsen. ๐œŽ -localization and ๐œŽ -martingales. Theory Probab. Appl., 48(1):152
163, 2004.
[40] J. Kallsen and J. Muhle-Karbe. Utility maximization in ane stochastic
volatility models. To appear in Int. J. Theoretical Appl. Finance, 2008.
[41] I. Karatzas and C. Kardaras. The numéraire portfolio in semimartingale nancial models. Finance Stoch., 11(4):447493, 2007.
[42] I. Karatzas and S. E. Shreve. Methods of Mathematical Finance. Springer,
New York, 1998.
[43] I. Karatzas and G. ยšitkovi¢. Optimal consumption from investment and
random endowment in incomplete semimartingale markets. Ann. Probab.,
31(4):18211858, 2003.
[44] C. Kardaras. No-free-lunch equivalences for exponential Lévy models under
convex constraints on investment. Math. Finance, 19(2):161187, 2009.
[45] C. Kardaras and G. ยšitkovi¢. Stability of the utility maximization problem
with random endowment in incomplete markets. To appear in Math. Finance.
[46] N. El Karoui and M.-C. Quenez. Dynamic programming and pricing of contingent claims in an incomplete market. SIAM J. Control Optim., 33(1):2966,
1995.
[47] T. S. Kim and E. Omberg. Dynamic nonmyopic portfolio behavior. Rev. Fin.
Studies, 9(1):141161, 1996.
[48] M. Kobylanski. Backward stochastic dierential equations and partial dierential equations with quadratic growth. Ann. Probab., 28(2):558602, 2000.
[49] D. Kramkov and W. Schachermayer. The asymptotic elasticity of utility functions and optimal investment in incomplete markets. Ann. Appl. Probab.,
9(3):904950, 1999.
[50] D. Kramkov and W. Schachermayer. Necessary and sucient conditions in the
problem of optimal investment in incomplete markets. Ann. Appl. Probab.,
13(4):15041516, 2003.
[51] K. Larsen. Continuity of utility-maximization with respect to preferences.
Math. Finance, 19(2):237250, 2009.
[52] J. P. Laurent and H. Pham. Dynamic programming and mean-variance hedging. Finance Stoch., 3(1):83110, 1999.
[53] M. Mania and M. Schweizer. Dynamic exponential utility indierence valuation. Ann. Appl. Probab., 15(3):21132143, 2005.
[54] M. Mania and R. Tevzadze. A unied characterization of the ๐‘ž -optimal and
minimal entropy martingale measures. Georgian Math. J., 10(2):289310,
2003.
[55] R. C. Merton. Lifetime portfolio selection under uncertainty: the continuoustime case. Rev. Econom. Statist., 51:247257, 1969.
136
BIBLIOGRAPHY
[56] R. C. Merton. Optimum consumption and portfolio rules in a continuous-time
model. J. Econ. Theory, 3:373413, 1971.
[57] J. Mossin. Optimal multiperiod portfolio policies. J. Bus., 41(2):215229,
1968.
[58] J. Muhle-Karbe. On Utility-Based Investment, Pricing and Hedging in Incomplete Markets. PhD thesis, TU München, 2009.
[59] M. Nutz. The Bellman equation for power utility maximization with semimartingales. Preprint arXiv:0912.1883v1, 2009.
[60] M. Nutz. Power utility maximization in constrained exponential Lévy models.
To appear in Math. Finance, 2009.
[61] M. Nutz. The opportunity process for optimal consumption and investment
with power utility. Math. Financ. Econ., 3(3):139159, 2010.
[62] M. Nutz. Risk aversion asymptotics for power utility maximization. Preprint
arXiv:1003.3582v1, 2010.
[63] P. Protter. An extension of Kazamaki's results on ๐ต๐‘€ ๐‘‚ dierentials. Ann.
Probab., 8(6):11071118, 1980.
[64] R. T. Rockafellar. Integral functionals, normal integrands and measurable
selections. In Nonlinear Operators and the Calculus of Variations, volume 543
of Lecture Notes in Math., pages 157207, Springer, Berlin, 1976.
[65] P. A. Samuelson. Lifetime portfolio selection by dynamic stochastic programming. Rev. Econ. Statist., 51(3):239246, 1969.
[66] P. A. Samuelson. Why we should not make mean log of wealth big though
years to act are long. J. Banking Finance, 3:305307, 1979.
[67] M. Santacroce. On the convergence of the ๐‘-optimal martingale measures to
the minimal entropy martingale measure. Stoch. Anal. Appl., 23(1):3154,
2005.
[68] K.-I. Sato. Lévy Processes and Innitely Divisible Distributions. Cambridge
University Press, Cambridge, 1999.
[69] W. Schachermayer. Optimal investment in incomplete markets when wealth
may become negative. Ann. Appl. Probab., 11(3):694734, 2001.
[70] W. Schoutens. Lévy Processes in Finance: Pricing Financial Derivatives.
Wiley, Chichester, 2003.
[71] M. Schweizer. On the minimal martingale measure and the Föllmer-Schweizer
decomposition. Stochastic Anal. and Appl., 13:573599, 1995.
[72] S. Stoikov and T. Zariphopoulou. Dynamic asset allocation and consumption
choice in incomplete markets. Australian Econ. Pap., 44(4):414454, 2005.
[73] M. Tehranchi. Explicit solutions of some utility maximization problems in
incomplete markets. Stochastic Process. Appl., 114(1):109125, 2004.
[74] T. Zariphopoulou. A solution approach to valuation with unhedgeable risks.
Finance Stoch., 5(1):6182, 2001.
[75] G. ยšitkovi¢. A ltered version of the bipolar theorem of Brannath and
Schachermayer. J. Theoret. Probab., 15:4161, 2002.
Curriculum Vitae
Marcel Fabian Nutz
born October 2, 1982
Swiss citizen
Education
Ph.D. Studies in Mathematics, ETH Zurich
10/200709/2010
Diploma in Mathematics, ETH Zurich
10/200203/2007
(Medal of ETH, diploma with distinction)
High School Gymnasium Münchenstein (BL)
(
Matura
08/199812/2001
with distinction)
Academic Employment
Teaching assistant in Mathematics, ETH Zurich
10/200709/2010
Tutor in Mathematics, ETH Zurich
03/200402/2007