Master LN
Master LN
Master LN
Nizar Touzi
nizar.touzi@polytechnique.edu
3
4
Conditional Expectation
and Linear Parabolic
PDEs
where T > 0 is a given maturity date. Here, b and σ are F⊗B(Rn )-progressively
measurable functions from [0, T ] × Ω × Rn to Rn and MR (n, d), respectively.
In particular, for every fixed x ∈ Rn , the processes {bt (x), σt (x), t ∈ [0, T ]} are
F−progressively measurable.
7
8 CHAPTER 1. CONDITIONAL EXPECTATION AND LINEAR PDEs
Let us mention that there is a notion of weak solutions which relaxes some
conditions from the above definition in order to allow for more general stochas-
tic differential equations. Weak solutions, as opposed to strong solutions, are
defined on some probabilistic structure (which becomes part of the solution),
and not necessarily on (Ω, F, F, P, W ). Thus, for a weak solution we search for a
probability structure (Ω̃, F̃, F̃, P̃, W̃ ) and a process X̃ such that the requirement
of the above definition holds true. Obviously, any strong solution is a weak
solution, but the opposite claim is false.
The main existence and uniqueness result is the following.
Theorem 1.2. Let X0 ∈ L2 be a r.v. independent of W . Assume that the
processes b. (0) and σ. (0) are in H2 , and that for some K > 0:
|bt (x) − bt (y)| + |σt (x) − σt (y)| ≤ K|x − y| for all t ∈ [0, T ], x, y ∈ Rn .
Then, for all T > 0, there exists a unique strong solution of (1.1) in H2 . More-
over,
2
≤ C 1 + E|X0 |2 eCT ,
E sup |Xt | (1.2)
t≤T
Clearly , the norms k.kH2 and k.kH2c on the Hilbert space H2 are equivalent.
Consider the map U on H2 by:
Z t Z t
U (X)t := X0 + bs (Xs )ds + σs (Xs )dWs , 0 ≤ t ≤ T.
0 0
By the Lipschitz property of b and σ in the x−variable and the fact that
b. (0), σ. (0) ∈ H2 , it follows that this map is well defined on H2 . In order
to prove existence and uniqueness of a solution for (1.1), we shall prove that
U (X) ∈ H2 for all X ∈ H2 and that U is a contracting mapping with respect to
the norm k.kH2c for a convenient choice of the constant c > 0.
1- We first prove that U (X) ∈ H2 for all X ∈ H2 . To see this, we decompose:
" Z Z
T t 2 #
2 2
kU (X)kH2 ≤ 3T kX0 kL2 + 3T E
bs (Xs )ds dt
0 0
" Z Z
T t 2 #
+3E
σs (Xs )dWs dt
0 0
1.1. Stochastic differential equations 9
"Z
T Z t 2 # "Z
T
#
2 2
E
bs (Xs )ds dt ≤ KT E (1 + |bt (0)| + |Xs | )ds < ∞,
0 0 0
"Z
T Z t 2 # " Z t 2 #
E σs (Xs )dWs dt ≤ T E max σs (Xs )dWs dt
0
0 t≤T 0
"Z #
T
≤ 4T E |σs (Xs )|2 ds
0
"Z #
T
2 2
≤ 4T KE (1 + |σs (0)| + |Xs | )ds < ∞.
0
2- To see that U is a contracting mapping for the norm k.kH2c , for some convenient
choice of c > 0, we consider two process X, Y ∈ H2 with X0 = Y0 , and we
estimate that:
2
E |U (X)t − U (Y )t |
Z t 2 Z t 2
≤ 2E
(bs (Xs ) − bs (Ys )) ds + 2E
(σs (Xs ) − σs (Ys )) dWs
0 0
Z t 2 Z t
2
= 2E (bs (Xs ) − bs (Ys )) ds + 2E |σs (Xs ) − σs (Ys )| ds
0 0
Z t Z t
2 2
= 2tE |bs (Xs ) − bs (Ys )| ds + 2E |σs (Xs ) − σs (Ys )| ds
0 0
Z t
2
≤ 2(T + 1)K E |Xs − Ys | ds.
0
2K(T + 1)
Hence, kU (X) − U (Y )kc ≤ kX − Y kc , and therefore U is a contract-
c
ing mapping for sufficiently large c.
Step 2 We next prove the estimate (1.2). We shall alleviate the notation writ-
10 CHAPTER 1. CONDITIONAL EXPECTATION AND LINEAR PDEs
where we used the Doob’s maximal inequality. Since b and σ are Lipschitz-
continuous in x, uniformly in t and ω, this provides:
Z t
2 2 2
E sup |Xu | ≤ C(K, T ) 1 + E|X0 | + E sup |Xu | ds
u≤t 0 u≤s
Exercise 1.3. In the context of this section, assume that the coefficients µ
and σ are locally Lipschitz and linearly growing in x, uniformly in (t, ω). By a
localization argument, prove that strong existence and uniqueness holds for the
stochastic differential equation (1.1).
In addition to the estimate (1.2) of Theorem 1.2, we have the following flow
continuity results of the solution of the SDE.
Theorem 1.4. Let the conditions of Theorem 1.2 hold true, and consider some
(t, x) ∈ [0, T ) × Rn with t ≤ t0 ≤ T .
(i) There is a constant C such that:
t,x0 2 0
≤ CeCt |x − x0 |2 .
t,x
E sup Xs − Xs |
(1.3)
t≤s≤t0
R t0
(ii) Assume further that B := supt<t0 ≤T (t0 − t)−1 E t |br (0)|2 + |σr (0)|2 dr <
0
Proof. (i) To simplify the notations, we set Xs := Xst,x and Xs0 := Xst,x for all
s ∈ [t, T ]. We also denote δx := x − x0 , δX := X − X 0 , δb := b(X) − b(X 0 ) and
1.1. Stochastic differential equations 11
Then, it follows from the Doob maximal inequality and the Lipschitz property
of the coefficients b and σ that:
Z s Z s
0 2 2
2 2
h(t ) := E sup |δXs | ≤ 3 |δx| + (s − t) E δbu du + 4
E δσu du
t≤s≤t0 t t
Z s
≤ 3 |δx|2 + K 2 (t0 + 4) E|δXu |2 du
t
Z s
≤ 3 |δx|2 + K 2 (t0 + 4) h(u)du .
t
Observe that
t0 t0
!
Z 2 Z 2
2
E|Xt0 − x| ≤ 2 E br (Xr )dr + E σr (Xr )dr
t t
0
t0
!
Z t Z
2 2
≤ 2 T E|br (Xr )| dr + E|σr (Xr )| dr
t t
Z t0
K 2 E|Xr − x|2 + |x|2 + E|br (0)|2 dr
≤ 6(T + 1)
t
Z t0
≤ 6(T + 1) (t0 − t)(|x|2 + B) + K 2 E|Xr − x|2 dr .
t
By the Gronwall inequality, this shows that
0
E|Xt0 − x|2 ≤ C(|x|2 + B)|t0 − t|eC(t −t) .
Plugging this estimate in (1.5), we see that:
Z u
2 0 C 0 2
h(u) ≤ 3 C(|x| + B)|t − t|e (t − t) + K (T + 4) h(r)dr , (1.6)
t0
where µ and σ satisfy the required condition for existence and uniqueness of a
strong solution.
For a function f : Rn −→ R, we define the function Af by
t,x
E[f (Xt+h )] − f (x)
Af (t, x) = lim if the limit exists.
h→0 h
Clearly, Af is well-defined for all bounded C 2 − function with bounded deriva-
tives and
∂2f
1
Af (t, x) = µ(t, x) · f (t, x) + Tr σσ T (t, x) , (1.7)
2 ∂x∂xT
1.3. Connection with PDE 13
Theorem 1.7. Let the coefficients µ, σ be continuous and satisfy (1.9). Assume
further that the function k is uniformly bounded from below, and f has quadratic
growth in x uniformly in t. Let v be a C 1,2 [0, T ), Rd ∩C 0 [0, T ) × Rd solution
of (1.8) with quadratic growth in x uniformly in t. Then
"Z #
T
t,x t,x t,x t,x
v(t, x) = E βs f (s, Xs )ds + βT g XT , t ≤ T, x ∈ Rd ,
t
Rs Rs Rs t,x
where Xst,x := x+ t
µ(u, Xut,x )du+ t
σ(u, Xut,x )dWu and βst,x := e− t
k(u,Xu )du
for t ≤ s ≤ T .
Proof. We first introduce the sequence of stopping times
1
∧ ∧ inf s > t : Xst,x − x ≥ n ,
τn := T−
n
and we oberve that τn −→ T P−a.s. Since v is smooth, it follows from Itô’s
formula that for t ≤ s < T :
t,x t,x t,x ∂v
+ Av s, Xst,x ds
d βs v s, Xs = βs −kv +
∂t
∂v
+βst,x s, Xst,x · σ s, Xst,x dWs
∂x
t,x t,x ∂v t,x
t,x
= βs −f (s, Xs )ds + s, Xs · σ s, Xs dWs ,
∂x
E βτt,x v τn , Xτt,x
n n
− v(t, x)
Z τn
∂v
βst,x −f (s, Xs )ds + s, Xst,x · σ s, Xst,x dWs .
= E
t ∂x
Now observe that the integrands in the stochastic integral is bounded by def-
inition of the stopping time τn , the smoothness of v, and the continuity of σ.
Then the stochastic integral has zero mean, and we deduce that
Z τn
t,x t,x t,x t,x
v(t, x) = E βs f s, Xs ds + βτn v τn , Xτn . (1.10)
t
Since τn −→ T and the Brownian motion has continuous sample paths P−a.s.
it follows from the continuity of v that, P−a.s.
Z τn
βst,x f s, Xst,x ds + βτt,x v τn , Xτt,x
n n
t Z T
n→∞
βst,x f s, Xst,x ds + βTt,x v T, XTt,x
−→ (1.11)
t Z
T
βst,x f s, Xst,x ds + βTt,x g XTt,x
=
t
1.3. Connection with PDE 15
By the estimate stated in the existence and uniqueness theorem 1.2, the latter
bound is integrable, and we deduce from the dominated convergence theorem
that the convergence in (1.11) holds in L1 (P), proving the required result by
taking limits in (1.10). ♦
By the Law of Large Numbers, it follows that v̂k (t, x) −→ v(t, x) P−a.s. More-
over the error estimate is provided by the Central Limit Theorem:
√ k→∞
k (v̂k (t, x) − v(t, x)) −→ N 0, Var g XTt,x
in distribution,
RT RT
for 1 ≤ i ≤ d, where µ, σ are F−adapted processes with 0 |µit |dt+ 0 |σti,j |2 dt <
∞ for all i, j = 1, . . . , d. It is convenient to use the matrix notations to represent
the dynamics of the price vector S = (S 1 , . . . , S d ):
λt := σt−1 (µt − rt 1) , 0 ≤ t ≤ T,
called the risk premium process. Here 1 is the vector of ones in Rd . We shall
frequently make use of the discounted processes
Z t
St
S̃t := 0 = St exp − ru du ,
St 0
Using the above matrix notations, the dynamics of the process S̃ are given by
dS̃t = S̃t ? (µt − rt 1)dt + σt dWt = S̃t ? σt (λt dt + dWt ) .
and the discounted wealth process induced by an initial capital X0 and a port-
folio strategy π can be written in
Z t
X̃tπ = X̃0 + π̃u · σu dBu , for 0 ≤ t ≤ T. (1.15)
0
The purpose of this section is to show that the financial market described
above contains no arbitrage opportunities. Our first observation is that, by the
1.3. Connection with PDE 19
For this reason, Q is called a risk neutral measure, or an equivalent local mar-
tingale measure, for the price process S.
We also observe that the discounted wealth process satisfies:
X̃ π is a Q−local martingale for every π ∈ A, (1.17)
as a stochastic integral with respect to the Q−Brownian motion B.
Theorem 1.12. The continuous-time financial market described above contains
no arbitrage opportunities, i.e. for every π ∈ A:
X0 = 0 and XTπ ≥ 0 P − a.s. =⇒ XTπ = 0 P − a.s.
Proof. For π ∈ A, the discounted wealth process X̃ π is a Q−local martingale
π
bounded from below h byi a Q−martingale. Then X̃ is a Q−super-martingale.
In particular, EQ X̃Tπ ≤ X̃0 = X0 . Recall that Q is equivalent to P and S 0
is strictly positive. Then, this inequality shows that, whenever X0π = 0 and
XTπ ≥ 0 P−a.s. (or equivalently Q−a.s.), we have X̃Tπ = 0 Q−a.s. and therefore
XTπ = 0 P−a.s. ♦
where the last equality follows from the Markov property of the process S.
Assuming further that g has linear growth, it follows that V has linear growth
in s uniformly in t. Since V is defined by a conditional expectation, it is expected
to satisfy the linear PDE:
1
−∂t V − rs ? DV − Tr (s ? σ)2 D2 V − rV
= 0. (1.19)
2
More precisely, if V ∈ C 1,2 (R+ , Rd ), the V is a classical solution of (1.19) and
satisfies the final condition V (T, .) = g. Coversely, if the PDE (1.19) combined
with the final condition v(T, .) = g has a classical solution v with linear growth,
then v coincides with the derivative security price V .
22 CHAPTER 1. CONDITIONAL EXPECTATION AND LINEAR PDEs
Chapter 2
Stochastic Control
and Dynamic Programming
The set S is called the parabolic interior of the state space. We will denote by
S̄ := cl(S) its closure, i.e. S̄ = [0, T ] × Rn for finite T , and S̄ = S for T = ∞.
b : (t, x, u) ∈ S × U −→ b(t, x, u) ∈ Rn
and
23
24 CHAPTER 2. STOCHASTIC CONTROL, DYNAMIC PROGRAMMING
for some constant K independent of (t, x, y, u). For each control process ν ∈ U,
we consider the controlled stochastic differential equation :
If the above equation has a unique solution X, for a given initial data, then
the process X is called the controlled process, as its dynamics is driven by the
action of the control process ν.
We shall be working with the following subclass of control processes :
U0 := U ∩ H2 , (2.4)
which guarantees the existence of a controlled process on the time interval [0, T 0 ]
for each given initial condition and control. The following result is an immediate
consequence of Theorem 1.2.
Theorem 2.1. Let ν ∈ U0 be a control process, and ξ ∈ L2 (P) be an F0 −measurable
random variable. Then, there exists a unique F−adapted process X ν satisfying
(6.3) together with the initial condition X0ν = ξ. Moreover for every T > 0,
there is a constant C > 0 such that
E sup |Xsν |2 < C(1 + E[|ξ|2 ])eCt for all t ∈ cl([0, T )). (2.5)
0≤s≤t
f, k : [0, T ) × Rn × U −→ R and g : Rn −→ R
for some constant K independent of (t, x, u). We define the cost function J on
[0, T ] × Rn × U by :
"Z #
T
J(t, x, ν) := E β ν (t, s)f (s, Xst,x,ν , νs )ds + β ν (t, T )g(XTt,x,ν )1T <∞ ,
t
and {Xst,x,ν , s ≥ t} is the solution of (6.3) with control process ν and initial
condition Xtt,x,ν = x.
Admissible control processes. In the finite horizon case T < ∞, the quadratic
growth condition on f and g together with the bound on k − ensure that J(t, x, ν)
is well-defined for all control process ν ∈ U0 . We then define the set of admissible
controls in this case by U0 .
More attention is needed for the infinite horizon case. In particular, the
discount term k needs to play a role to ensure the finiteness of the integral. In
this setting the largest set of admissible control processes is given by
∞ ν
n Z o
β (t, s) 1+|Xst,x,ν |2 +|νs )| ds < ∞ for all x when T = ∞.
U0 := ν ∈ U : E
0
The stochastic control problem. The purpose of this section is to study the min-
imization problem
V (t, x) := sup J(t, x, ν) for (t, x) ∈ S.
ν∈U0
Our main concern is to describe the local behavior of the value function V
by means of the so-called dynamic programming equation, or Hamilton-Jacobi-
Bellman equation. We continue with some remarks.
Remark 2.2. (i) If V (t, x) = J(t, x, ν̂t,x ), we call ν̂t,x an optimal control for
the problem V (t, x).
(ii) The following are some interesting subsets of controls :
- a process ν ∈ U0 which is adapted to the natural filtration FX of the
associated state process is called feedback control,
- a process ν ∈ U0 which can be written in the form νs = ũ(s, Xs ) for some
measurable map ũ from [0, T ] × Rn into U , is called Markovian control;
notice that any Markovian control is a feedback control,
- the deterministic processes of U0 are called open loop controls.
(iii) Suppose that T < ∞, and let (Y, Z) be the controlled processes defined
by
dYs = Zs f (s, Xs , νs )ds and dZs = −Zs k(s, Xs , νs )ds ,
and define the augmented state process X̄ := (X, Y, Z). Then, the above
value function V can be written in the form :
V (t, x) = V̄ (t, x, 0, 1) ,
where x̄ = (x, y, z) is some initial data for the augmented state process X̄,
V̄ (t, x̄) := Et,x̄ ḡ(X̄T ) and ḡ(x, y, z) := y + g(x)z .
Hence the stochastic control problem V can be reduced without loss of
generality to the case where f = k ≡ 0. We shall appeal to this reduced
form whenever convenient for the exposition.
26 CHAPTER 2. STOCHASTIC CONTROL, DYNAMIC PROGRAMMING
(iv) For notational simplicity we consider the case T < ∞ and f = k = 0. The
previous remark shows how to immediately adapt the following argument
so that the present remark holds true without the restriction f = k = 0.
The extension to the infinite horizon case is also immediate.
Consider the value function
sup E g(XTt,x,ν ) ,
Ṽ (t, x) := (2.6)
ν∈Ut
Ut := {ν ∈ U0 : ν independent of Ft } . (2.7)
Ṽ = V, (2.8)
so that both problems are indeed equivalent. To see this, fix (t, x) ∈ S and
ν ∈ U0 . Then, ν can be written as a measurable function of the canonical
process ν((ωs )0≤s≤t , (ωs − ωt )t≤s≤T ), where, for fixed (ωs )0≤s≤t , the map
ν(ωs )0≤s≤t : (ωs − ωt )t≤s≤T 7→ ν((ωs )0≤s≤t , (ωs − ωt )t≤s≤T ) can be viewed
as a control independent on Ft . Using the independence of the increments
of the Brownian motion, together with Fubini’s Lemma, it thus follows
that
Z h
t,x,ν(ωs )0≤s≤t i
J(t, x; ν) = E g(XT ) dP((ωs )0≤s≤t )
Z
≤ Ṽ (t, x)dP((ωs )0≤s≤t ) = Ṽ (t, x).
for all (t, x) ∈ S̄. We also recall the subset of controls Ut introduced in (2.7)
above.
2.2. Dynamic programming principle 27
Theorem 2.3. Assume that V is locally bounded and fix (t, x) ∈ S. Let {θν , ν ∈
Ut } be a family of finite stopping times independent of Ft with values in [t, T ].
Then:
"Z ν #
θ
ν t,x,ν ν ν ν t,x,ν
V (t, x) ≥ sup E β (t, s)f (s, Xs , νs )ds + β (t, θ )V∗ (θ , Xθν ) .
ν∈Ut t
Observe that the supremum is now taken over the subset U of the finite
dimensional space Rk . Hence, the dynamic programming principle allows
to reduce the initial maximization problem, over the subset U of the in-
finite dimensional set of Rk −valued processes, into a finite dimensional
maximization problem. However, we are still facing an infinite dimen-
sional problem since the dynamic programming principle relates the value
function at time t to the value function at time t + 1.
(iii) In the context of the above discrete-time framework with finite horizon
T < ∞, notice that the dynamic programming principle suggests the fol-
lowing backward algorithm to compute V as well as the associated optimal
strategy (when it exists). Since V (T, ·) = g is known, the above dynamic
programming principle can be applied recursively in order to deduce the
value function V (t, x) for every t.
(iv) In the continuous time setting, there is no obvious counterpart to the
above backward algorithm. But, as the stopping time θ approaches t,
the above dynamic programming principle implies a special local behavior
for the value function V . When V is known to be smooth, this will be
obtained by means of Itô’s formula.
28 CHAPTER 2. STOCHASTIC CONTROL, DYNAMIC PROGRAMMING
(vi) Once the local behavior of the value function is characterized, we are
faced to the important uniqueness issue, which implies that V is com-
pletely characterized by its local behavior together with some convenient
boundary condition.
J(θ, Xθ , ν ε ) ≥ V (θ, Xθ ) − ε.
Clearly, one can choose ν ε = µ on the stochastic interval [t, θ]. Then
"Z #
θ
ε ε
V (t, x) ≥ J(t, x, ν ) = Et,x β(t, s)f (s, Xs , µs )ds + β(t, θ)J(θ, Xθ , ν )
t
"Z #
θ
≥ Et,x β(t, s)f (s, Xs , µs )ds + β(t, θ)V (θ, Xθ ) − ε Et,x [β(t, θ)] .
t
right-hand side of the classical dynamic programming principle (2.9) is not even
known to be well-defined.
The formulation of Theorem 2.3 avoids this measurability problem since
V∗ and V ∗ are lower- and upper-semicontinuous, respectively, and therefore
measurable. In addition, it allows to avoid the typically heavy technicalities
related to measurable selection arguments needed for the proof of the classical
(2.9) after a convenient relaxation of the control problem, see e.g. El Karoui
and Jeanblanc [5].
Proof of Theorem 2.3 For simplicity, we consider the finite horizon case
T < ∞, so that, without loss of generality, we assume f = k = 0, See Remark
2.2 (iii). The extension to the infinite horizon framework is immediate.
1. Let ν ∈ Ut be arbitrary and set θ := θν . Then:
where ν̃ω is obtained from ν by freezing its trajectory up to the stopping time
θ. Since, by definition, J(θ(ω), Xθt,x,ν (ω); ν̃ω ) ≤ V ∗ (θ(ω), Xθt,x,ν (ω)), it follows
from the tower property of conditional expectations that
With this construction, it follows from (2.10), (2.11), together with the fact that
V ≥ ϕ, that the countable family (Ai )i≥0 satisfies
n
X
t,x,ν
νsε,n := 1[t,θ] (s)νs + 1(θ,T ] (s) νs 1(An )c (θ, Xθ ) + 1Ai (θ, Xθt,x,ν )νsi,ε .
i=1
n
X
J(θ, Xθt,x,ν , ν i,ε )1Ai θ, Xθt,x,ν
+
i=1
n
X
ϕ(θ, Xθt,x,ν − 3ε 1Ai θ, Xθt,x,ν
≥
i=0
where the last equality follows from the left-hand side of (2.12)
and from the
monotone convergence theorem, due to the fact that either E ϕ(θ, Xθt,x,ν )+ <
∞ or E ϕ(θ, Xθt,x,ν )− < ∞. By the arbitrariness of ν ∈ Ut and ε > 0, this
shows that:
V (t, x) ≥ sup E ϕ(θ, Xθt,x,ν ) .
(2.13)
ν∈Ut
3. It remains to deduce the first inequality of Theorem 2.3 from (2.13). Fix
r > 0. It follows from standard arguments, see e.g. Lemma 3.5 in [12], that
we can find a sequence of continuous functions (ϕn )n such that ϕn ≤ V∗ ≤ V
for all n ≥ 1 and such that ϕn converges pointwise to V∗ on [0, T ] × Br (0).
Set φN := minn≥N ϕn for N ≥ 1 and observe that the sequence (φN )N is non-
decreasing and converges pointwise to V∗ on [0, T ] × Br (0). By (2.13) and the
monotone convergence Theorem, we then obtain:
V (t, x) ≥ lim E φN (θν , Xt,x
ν
(θν )) = E V∗ (θν , Xt,x
ν
(θν )) .
N →∞
for every s ≥ t and smooth function ϕ ∈ C 1,2 ([t, s], Rn ) and each admissible
control process ν ∈ U0 .
32 CHAPTER 2. STOCHASTIC CONTROL, DYNAMIC PROGRAMMING
Proposition 2.4. Assume the value function V ∈ C 1,2 ([0, T ), Rn ), and let the
coefficients k(·, ·, u) and f (·, ·, u) be continuous in (t, x) for all fixed u ∈ U .
Then, for all (t, x) ∈ S:
−∂t V (t, x) − H t, x, V (t, x), DV (t, x), D2 V (t, x) ≥ 0.
(2.14)
Proof. Let (t, x) ∈ S and u ∈ U be fixed and consider the constant control
process ν = u, together with the associated state process X with initial data
Xt = x. For all h > 0, Define the stopping time :
θh := inf {s > t : (s − t, Xs − x) 6∈ [0, h) × αB} ,
where α > 0 is some given constant, and B denotes the unit ball of Rn . Notice
that θh −→ t, P−a.s. when h & 0, and θh = h for h ≤ h̄(ω) sufficiently small.
1. From the first inequality of the dynamic programming principle, it follows
that :
" #
Z θh
0 ≤ Et,x β(0, t)V (t, x) − β(0, θh )V (θh , Xθh ) − β(0, r)f (r, Xr , u)dr
t
"Z #
θh
·
= −Et,x β(0, r)(∂t V + L V + f )(r, Xr , u)dr
t
"Z #
θh
−Et,x β(0, r)DV (r, Xr ) · σ(r, Xr , u)dWr ,
t
the last equality follows from Itô’s formula and uses the crucial smoothness
assumption on V .
2. Observe that β(0, r)DV (r, Xr ) · σ(r, Xr , u) is bounded on the stochastic
interval [t, θh ]. Therefore, the second expectation on the right hand-side of the
last inequality vanishes, and we obtain :
" Z #
1 θh ·
−Et,x β(0, r)(∂t V + L V + f )(r, Xr , u)dr ≥ 0
h t
We now send h to zero. The a.s. convergence of the random value inside the
expectation is easily obtained by the mean value Theorem; recall that θh = h
Rθ
for sufficiently small h > 0. Since the random variable h−1 t h β(0, r)(L· V +
f )(r, Xr , u)dr is essentially bounded, uniformly in h, on the stochastic interval
[t, θh ], it follows from the dominated convergence theorem that :
−∂t V (t, x) − Lu V (t, x) − f (t, x, u) ≥ 0.
By the arbitrariness of u ∈ U , this provides the required claim. ♦
We next wish to show that V satisfies the nonlinear partial differential equa-
tion (2.15) with equality. This is a more technical result which can be proved by
different methods. We shall report a proof, based on a contradiction argument,
which provides more intuition on this result, although it might be slightly longer
than the usual proof reported in standard textbooks.
2.3. Dynamic programming equation 33
Proposition 2.5. Assume the value function V ∈ C 1,2 ([0, T ), Rn ), and let the
function H be upper semicontinuous, and kk + k∞ < ∞. Then, for all (t, x) ∈
S:
ϕ(t, x) := V (t, x) + ε |t − t0 |2 + |x − x0 |4 .
Then
where B denotes the unit ball centered at x0 . We next observe that the param-
eter γ defined by the following is positive:
+
k∞
−γeηkk := max (V − ϕ) < 0. (2.17)
∂Nη
and observe that, by continuity of the state process, (θ, Xθ ) ∈ ∂Nη , so that :
+
k∞
(V − ϕ)(θ, Xθ ) ≤ −γeηkk
where the ”dW ” integral term has zero mean, as its integrand is bounded on the
stochastic interval [t0 , θ]. Observe also that (∂t ϕ + Lνr ϕ)(r, Xr ) + f (r, Xr , νr ) ≤
h(r, Xr ) ≤ 0 on the stochastic interval [t0 , θ]. We therefore deduce that :
hZ θ i
V (t0 , x0 ) ≥ γ + Et0 ,x0 β(t0 , r)f (r, Xr , νr )dr + β(t0 , θ)V (θ, Xθ ) .
t0
which is the required contradiction of the second part of the dynamic program-
ming principle, and thus completes the proof. ♦
≤ Const |x − x0 |,
where we used the Lipschitz-continuity of g together with the flow estimates
of Theorem 1.4, and the fact that the coefficients b and σ are Lipschitz in x
uniformly in (t, u). This compltes the proof of the Lipschitz property of the
value function V .
(ii) To prove the Hölder continuity in t, we shall use the dynamic programming
principle.
(ii-1) We first make the following important observation. A careful review
of the proof of Theorem 2.3 reveals that, whenever the stopping times θν are
constant (i.e. deterministic), the dynamic programming principle holds true
with the semicontinuous envelopes taken only with respect to the x−variable.
Since V was shown to be continuous in the first part of this proof, we deduce
that:
V (t, x) = sup E V t0 , Xtt,x,ν
0 (2.19)
ν∈U0
dXt = νt dWt ,
1. If V is C 1,2 ([0, T ), R), then it follows from Proposition 2.4 that V satisfies
1
−∂t V − u2 D2 V ≥ 0 for all u ∈ R,
2
and all (t, x) ∈ [0, T ) × R. By sending u to infinity, it follows that
where g conc is the concave envelope of g, i.e. the smallest concave majorant of
g. Notice that g conc < ∞ as g is bounded from above by a line.
V (t, x) := sup Et,x [g(XTν )] ≤ sup Et,x [g conc (XTν )] = g conc (x),
ν∈U0 ν∈U0
V ∈ C 1,2 ([0, T ), R)
=⇒ V (t, x) = g conc (x) for all (t, x) ∈ [0, T ) × R.
Now recall that this implication holds for any arbitrary non-negative lower semi-
continuous function g. We then obtain a contradiction whenever the function
g conc is not C 2 (R). Hence
As in the previous chapter, we assume here that the filtration F is defined as the
P−augmentation of the canonical filtration of the Brownian motion W defined
on the probability space (Ω, F, P).
Our objective is to derive similar results, as those obtained in the previous
chapter for standard stochastic control problems, in the context of optimal stop-
ping problems. We will then first start by the formulation of optimal stopping
problems, then the corresponding dynamic programming principle, and dynamic
programming equation.
39
40 CHAPTER 3. OPTIMAL STOPPING, DYNAMIC PROGRAMMING
is well-defined for all (t, x) ∈ S and τ ∈ T[t,T ] . Here, X t,x denotes the unique
strong solution of (3.1) with initial condition Xtt,x = x.
The optimal stopping problem is now defined by:
is called the stopping region and is of particular interest: whenever the state is
in this region, it is optimal to stop immediately. Its complement S c is called
the continuation region.
Remark 3.1. As in the previous chapter, we could have considered an appear-
ently more general criterion
Z τ
t,x
V (t, x) := sup E β(t, s)f (s, Xs )ds + β(t, τ )g Xτ 1τ <∞ ,
τ ∈T[t,T ] t
with
Rs
β(t, s) := e− t
k(s,Xs )ds
for 0 ≤ t ≤ s < T.
we see immediately that we may reduce this problem to the context of (3.4).
Remark 3.2. Consider the subset of stopping rules:
t
T[t,T ] := τ ∈ T[t,T ] : τ independent of Ft . (3.6)
By a similar argument as in Remark 2.2 (iv), we can see that the maximization
in the optimal stopping problem (3.4) can be restricted to this subset, i.e.
for all (t, x) ∈ S and τ ∈ T[t,T ] . In particular, the proof in the latter reference
does not require any heavy measurable selection, and is essentially based on the
supermartingale nature of the so-called Snell envelope process. Moreover, we
observe that it does not require any Markov property of the underlying state
process.
We report here a different proof in the sprit of the weak dynamic program-
ming principle for stochastic control problems proved in the previous chapter.
The subsequent argument is specific to our Markovian framework and, in this
sense, is weaker than the classical dynamic programming principle. However,
the combination of the arguments of this chapter with those of the previous
chapter allow to derive a dynamic programming principle for mixed stochastic
control and stopping problem.
t
The following claim will be making using of the subset T[t,T ] , introduced
in (3.6), of all stopping times in T[t,T ] which are independent of Ft , and the
notations:
for all (t, x) ∈ S̄. We recall that V∗ and V ∗ are the lower and upper semicon-
tinuous envelopes of V , and that V∗ = V ∗ = V whenever V is continuous.
t
Theorem 3.3. Assume that V is locally bounded. For (t, x) ∈ S, let θ ∈ T̄[t,T ]
be a stopping time such that Xθt,x is bounded. Then:
Proof. Inequality (3.9) follows immediately from the tower property and the
fact that J ≤ V ∗ .
We next prove inequality (3.10) with V∗ replaced by an arbitrary function
which implies (3.10) by the same argument as in Step 3 of the proof of Theorem
2.3.
42 CHAPTER 3. OPTIMAL STOPPING, DYNAMIC PROGRAMMING
Arguying as in Step 2 of the proof of Theorem 2.3, we first observe that, for
every ε > 0, we can find a countable family Āi ⊂ (ti − ri , ti ] × Ai ⊂ S, together
with a sequence of stopping times τ i,ε in T[ttii,T ] , i ≥ 1, satisfying Ā0 = {T } × Rd
and
n
!
X
τ n,ε := τ 1{τ <θ} + 1{τ ≥θ} θ, Xθt,x τ i,ε 1Āi θ, Xθt,x
T 1(Ān )c +
i=1
t
defines a stopping time in T[t,T ] . We then deduce from the tower property and
(3.11) that
V̄ (t, x) ≥ ¯ x; τ n,ε )
J(t,
E g Xτt,x 1{τ <θ} + 1{τ ≥θ} ϕ(θ, Xθt,x ) − 3ε 1Ān (θ, Xθt,x )
≥
+E 1{τ ≥θ} g(XTt,x )1(Ān )c (θ, Xθt,x ) .
t
and the result follows from the arbitrariness of ε > 0 and τ ∈ T[t,T ]. ♦
We next apply Itô’s formula, and observe that the expected value of the diffusion
term vanishes because (t, Xt ) lies in the compact subset [t0 , t0 + h] × B for
t ∈ [t0 , θh ]. Then:
" #
−1 θh
Z
t0 ,x0
E (∂t + A)V (t, Xt )dt ≥ 0.
h t0
V (t0 , x0 ) > g(x0 ) and − (∂t + A)V (t0 , x0 ) > 0 at some (t0 , x0 ) ∈ S, (3.14)
Moreover:
Next, let
t, Xtt0 ,x0 ∈
θ := inf t > t0 : 6 Nh .
t
For an arbitrary stopping rule τ ∈ T[t,T ] , we compute by Itô’s formula that:
E [V (τ ∧ θ, Xτ ∧θ ) − V (t0 , x0 )] = E [(V − ϕ) (τ ∧ θ, Xτ ∧θ )]
+E [ϕ (τ ∧ θ, Xτ ∧θ ) − ϕ(t0 , x0 )]
= E [(V − ϕ) (τ ∧ θ, Xτ ∧θ )]
"Z #
τ ∧θ
+E (∂t + A)ϕ(t, Xtt0 ,x0 )dt ,
t0
where the diffusion term has zero expectation because the process (t, Xtt0 ,x0 ) is
confined to the compact subset Nh on the stochastic interval [t0 , τ ∧ θ]. Since
−Lϕ ≥ 0 on Nh by (3.15), this provides:
E [V (τ ∧ θ, Xτ ∧θ ) − V (t0 , x0 )] ≤ E [(V − ϕ) (τ ∧ θ, Xτ ∧θ )]
≤ −γP[τ ≥ θ],
by (3.16). Then, since V ≥ g + δ on Nh by (3.15):
V (t0 , x0 ) ≥ γP[τ ≥ θ] + E g(Xτt0 ,x0 ) + δ 1{τ <θ} + V θ, Xθt0 ,x0 1{τ ≥θ}
t
By the arbitrariness of τ ∈ T[t,T ] , this provides the desired contradiction of (3.9).
♦
Proof. (i) For t ∈ [0, T ] and x, x0 ∈ Rn , it follows from the Lipschitz property
of g that:
0
|V (t, x) − V (t, x0 )| ≤ Const sup E Xτt,x − Xτt,x
τ ∈T[t,T ]
0
≤ Const E sup Xτt,x − Xτt,x
t≤s≤T
≤ Const |x − x0 |
45
where the last inequality follows from the fact that V ≥ g. Using the Lipschitz
property of g, this provides:
0 t,x t,x
t,x
0 ≤ V (t, x) − E V t , Xt0 ≤ Const E sup Xs − Xt0
t≤s≤t0
√
≤ Const (1 + |x|) t0 − t
by the flow continuity result of Theorem 1.4. Using this estimate together with
the Lipschitz property proved in (i) above, this provides:
which means that there is no subinterval of R from which the process X can
not exit.
We consider the infinite horizon optimal stopping problem:
1
Av := µv 0 + σ 2 v 00 . (3.21)
2
The ordinary differential equation
Av − βv = 0 (3.22)
Clearly ψ and φ are uniquely determined up to a positive constant, and all other
solution of (3.22) can be expressed as a linear combination of ψ and φ.
The following result follows from an immediate application of Itô’s formula.
We now show that the value function V is concave up to some change of vari-
able, and provides conditions under which V is C 1 across the exercise boundary,
i.e. the boundary between the exercise and the continuation regions. For the
next result, we observe that the fnction (ψ/φ) is continuous and strictly increas-
ing by (3.23), and therefore invertible.
For ε > 0, consider the ε−optimal stopping rules τ1 , τ2 ∈ T for the problems
V (b1 ) and V (b2 ):
E e−βτi g Xτ0,b
i
i
≥ V (bi ) − ε for i = 1, 2.
where θ denotes the shift operator on the canonical space, i.e. θt (ω)(s) =
ω(t + s). In words, the stopping rule τ ε uses the ε−optimal stopping rule τ1 if
the level b1 is reached before the level b2 , and the ε−optimal stopping rule τ2
otherwise. Then, it follows from the strong Markov property that
h ε
i
V (x) ≥ E e−βτ g Xτ0,x ε
h x
i
= E e−βHb1 E e−βτ1 g Xτ0,b
1
1
1{Hbx <Hbx }
1 2
h i
−βHbx2 1
−βτ2 0,b2
+E e E e g Xτ2 1{Hbx <Hbx }
2 1
h i
−βHbx1
≥ (V (b1 ) − ε) E e 1{Hbx <Hbx }
1 2
h i
−βHbx2
+ (V (b2 ) − ε) E e 1{Hbx <Hbx } .
2 1
for all ε > 0, δ > 0. Multiplying by ((ψ/φ)(x0 + ε) − (ψ/φ)(x0 ))/ε, this implies
that:
g
φ (x0 + ε) − φg (x0 ) V
φ (x0 + ε) − V
φ (x0 ) ∆+ (ε)
g
φ (x0 − δ) − φg (x0 )
≤ ≤ ,
ε ε ∆− (δ) δ
(3.26)
where
ψ ψ ψ ψ
φ (x0 + ε) − φ (x0 ) φ (x0 − δ) − φ (x0 )
∆+ (ε) := and ∆− (δ) := .
ε δ
We next consider two cases:
d+ ( Vφ )
0
g
(x0 ) = (x0 ). (3.27)
dx φ
• If (ψ/φ)0 (x0 ) = 0, then, we use the fact that for every sequence εn & 0,
there is a subsequence εnk & 0 and δk & 0 such that ∆+ (εnk ) = ∆− (δk ).
Then (3.26) reduces to:
g
φ (x0 + εnk ) − φg (x0 ) V
φ (x0 + ε nk ) − V
φ (x0 )
g
φ (x0 − δk ) − φg (x0 )
≤ ≤ ,
ε nk εnk δk
and therefore
V V 0
φ (x0 + ε nk ) − φ (x0 ) g
−→ (x0 ).
εnk φ
d− ( Vφ )
0
g
(x0 ) = (x0 ),
dx φ
E g Xτt,x .
= sup
τ ∈T[t,∞)
Let us assume that V ∈ C 1,2 (S), and work towards a contradiction. We first
observe by the homogeneity of the problem that V (t, x) = V (x) is independent
of t. Moreover, it follows from Theorem 3.4 that V is concave in x and V ≥ g.
Then
V ≥ g conc , (3.28)
51
52 CHAPTER 4. THE VERIFICATION ARGUMENT
with
Rs
k(r,Xrt,x,ν ,νr )dr
β ν (t, s) := e− t .
H(t, x, r, p, γ)
1
:= sup −k(t, x, u)r + b(t, x, u) · p + Tr[σσ T (t, x, u)γ] + f (t, x, u) ,
u∈U 2
where b and σ satisfy the conditions (2.1)-(2.2), and the coefficients f and k are
measurable. From the results of the previous section, the dynamic programming
equation corresponding to the stochastic control problem (4.1) is:
The proof of the subsequent result will make use of the following linear second
order operator
for every t ≤ s and smooth function ϕ ∈ C 1,2 ([t, s], Rd ) and each admissible
control process ν ∈ U0 . The last expression is an immediate application of Itô’s
formula.
Theorem 4.1. Let T < ∞, and v ∈ C 1,2 ([0, T ), Rd ) ∩ C([0, T ] × Rd ). Assume
that kk − k∞ < ∞ and v and f have quadratic growth, i.e. there is a constant C
such that
where the last inequality uses the condition v(T, ·) ≥ g. Since the control ν ∈ U0
is arbitrary, this completes the proof of (i).
Statement (ii) is proved by repeating the above argument and observing that
the control ν̂ achieves equality at the crucial step (4.3). ♦
Remark 4.2. When U is reduced to a singleton, the optimization problem V is
degenerate. In this case, the DPE is linear, and the verification theorem reduces
to the so-called Feynman-Kac formula.
Notice that the verification theorem assumes the existence of such a solution,
and is by no means an existence result. However, it provides uniqueness in the
class of functions with quadratic growth.
We now state without proof an existence result for the DPE together with
the terminal condition V (T, ·) = g (see [8] and the references therein). The main
assumption is the so-called uniform parabolicity condition :
there is a constant c > 0 such that
(4.4)
ξ 0 σσ 0 (t, x, u) ξ ≥ c|ξ|2 for all (t, x, u) ∈ [0, T ] × Rn × U .
In the following statement, we denote by Cbk (Rn ) the space of bounded functions
whose partial derivatives of orders ≤ k exist and are bounded continuous. We
similarly denote by Cbp,k ([0, T ], Rn ) the space of bounded functions whose partial
derivatives with respect to t, of orders ≤ p, and with respect to x, of order ≤
k, exist and are bounded continuous.
Theorem 4.3. Let Condition 4.4 hold, and assume further that :
• U is compact;
• b, σ and f are in Cb1,2 ([0, T ], Rn );
• g ∈ Cb3 (Rn ).
Then the DPE (2.18) with the terminal data V (T, ·) = g has a unique solution
V ∈ Cb1,2 ([0, T ] × Rn ).
The remaining wealth (Xt − πt ) is invested in the risky asset. Therefore, the
liquidation value of a self-financing strategy satisfies
dSt dS 0
dXtπ = πt + (Xtπ − πt ) 0t
St St
= (rXt + (µ − r)πt ) dt + σπt dWt . (4.5)
∂w
(t, x) + sup Au w(t, x) = 0, (4.6)
∂t u∈R
∂w 1 ∂2w
Au w(t, x) := (rx + (µ − r)u) (t, x) + σ 2 u2 (t, x).
∂x 2 ∂x2
We next search for a solution of the dynamic programming equation of the form
v(t, x) = xγ h(t). Plugging this form of solution into the PDE (4.6), we get the
following ordinary differential equation on h :
u2
u 1
0 = h0 + γh sup r + (µ − r) + (γ − 1)σ 2 2 (4.7)
u∈R x 2 x
1
= h0 + γh sup r + (µ − r)δ + (γ − 1)σ 2 δ 2 (4.8)
δ∈R 2
1 (µ − r)2
= h0 + γh r + , (4.9)
2 (1 − γ)σ 2
Since v(T, ·) = U (x), we seek for a function h satisfying the above ordinary
differential equation together with the boundary condition h(T ) = 1. This
induces the unique candidate:
1 (µ − r)2
a(T −t)
h(t) := e with a := γ r + .
2 (1 − γ)σ 2
Hence, the function (t, x) 7−→ xγ h(t) is a classical solution of the HJB equation
(4.6). It is easily checked that the conditions of Theorem 4.1 are all satisfied in
this context. Then V (t, x) = xγ h(t), and the optimal portfolio allocation policy
is given by the linear control process:
µ−r
π̂t = X π̂ .
(1 − γ)σ 2 t
and
2
yDyz v(t, y, z) = 8µ2 λ3 (T − t) v(t, y, z) y 2 ≥ 0 . (4.11)
β
b b
arbitrary real parameter β, we denote by A the generator the process
2. For an
Y ,Z :
1 2 2 1
Aβ := β Dyy + y 2 Dzz
2 2
+ βyDyz .
2 2
In this step, we intend to prove that for all t ∈ [0, T ] and y, z ∈ R :
The second equality follows from the fact that {v(t, Yt1 , Zt1 ), t ≤ T } is a mar-
tingale . As for the first equality, we see from (4.10) and (4.11) that 1 is a
maximizer of both functions β 7−→ β 2 Dyy
2
v(t, y, z) and β 7−→ βyDyz 2
v(t, y, z) on
[−1, 1].
3. Let b be some given predictable process valued in [−1, 1], and define the
sequence of stopping times
We are now able to prove the law of the iterated logarithm for double stochas-
tic integrals by a direct adaptation of the case of the Brownian motion. Set
1
h(t) := 2t log log for t > 0 .
t
58 CHAPTER 4. THE VERIFICATION ARGUMENT
where b̃ := δ −1 (β̄ − b) is valued in [−1, 1]. It then follows from the second
inequality that :
2Xtb
lim sup ≥ β̄ − δ = β0 .
t&0 h(t)
We now prove the second inequality. Clearly, we can assume with no loss of
generality that kbk∞ ≤ 1. Let T > 0 and λ > 0 be such that 2λT < 1. It
follows from Doob’s maximal inequality for submartingales that for all α ≥ 0,
P max 2Xtb ≥ α = P max exp(2λXtb ) ≥ exp(λα)
0≤t≤T 0≤t≤T
h b
i
≤ e−λα E e2λXT .
Since k≥0 k −(1+η) < ∞, it follows from the Borel-Cantelli lemma that, for
P
almost all ω ∈ Ω, there exists a natural number K θ,η (ω) such that for all
k ≥ K θ,η (ω),
max 2Xtb (ω) < (1 + η)2 h(θk ) .
0≤t≤θ k
4.2. Examples 59
h(t)
2Xtb (ω) < (1 + η)2 h(θk ) ≤ (1 + η)2 .
θ
Hence,
2Xtb (1 + η)2
lim sup < a.s.
t&0 h(t) θ
and the required result follows by letting θ tend to 1 and η to 0 along the
rationals. ♦
where b and σ satisfy the usual Lipschitz and linear growth conditions. Given
the functions k, f : [0, T ] × Rd −→ R and g : Rd −→ R, we consider the optimal
stopping problem
Z τ
V (t, x) := sup E β(t, s)f (s, Xst,x )ds + β(t, τ )g(Xτt,x ) , (4.15)
t
τ ∈T[t,T t
]
Before stating the main result of this section, we observe that for many inter-
esting examples, it is known that the value function V does not satisfy the C 1,2
regularity which we have been using so far for the application of Itô’s formula.
Therefore, in order to state a result which can be applied to a wider class of
problems, we shall enlarge in the following remark the set of function for which
Itô’s formula still holds true.
Remark 4.6. Let v be a function in the Sobolev space W 1,2 (S). By definition,
for such a function v, there is a sequence of functions (v n )n≥1 ⊂ C 1,2 (S) such
that v n −→ v uniformly on compact subsets of S, and
Then, Itô’s formula holds true for v n for all n ≥ 1, and is inherited by v by
sending n → ∞.
Theorem 4.7. Let T < ∞ and v ∈ W 1,2 ([0, T ), Rd ). Assume further that v
and f have quadratic growth. Then:
(i) If v is a supersolution of (4.16), then v ≥ V .
(ii) If v is a solution of (4.16), then v = V and
Notice that τn −→ τ a.s. Then, since f and v have quadratic growth, we may
pass to the limit n → ∞ invoking the dominated convergence theorem, and we
get:
h Z T i
v(t, x) ≥ E βT v(T, XTt,x ) + βs f (s, Xst,x )ds .
t
4.2. Examples 61
Since v(T, .) ≥ g by the supersolution property, this concludes the proof of (i).
(ii) Let τt∗ be the stopping time introduced in the theorem. Then, since v(T, .) =
g, it follows that τt∗ ∈ T[t,T
t
] . Set
Observe that v > g on [t, τtn ) ⊂ [t, τt∗ ) and therefore −∂t v − Lv − f = 0 on
[t, τtn ). Then, proceeding as in the previous step, it follows from Itô’s formula
that:
h Z τtn i
v(t, x) = E βτtn v(τtn , Xτt,x
n ) +
t
β s f (s, Xs
t,x
)ds .
t
Since τtn −→ τt∗ a.s. and f, v have quadratic growth, we may pass to the limit
n → ∞ invoking the dominated convergence theorem. This leads to:
h Z T i
t,x
v(t, x) = E βT v(T, XT ) + βs f (s, Xst,x )ds ,
t
and the required result follows from the fact that v(T, .) = g. ♦
In view this time independence, it follows that the dynamic programming cor-
responding to this problem is:
1
min{v − (K − s)+ , rv − rsDv − σ 2 D2 v} = 0. (4.18)
2
62 CHAPTER 4. THE VERIFICATION ARGUMENT
(4.20)
Indeed, for an arbitrary stopping time τ ∈ T[0,∞) , it follows from Itô’s formula
that:
Z τ Z τ
1
p(s) = e−rτ p(Sτ0,s ) − e−rt (−rp + rsp0 + σ 2 s2 p00 )(St )dt − p0 (St )σSt dWt
0 2 0
Z τ
−rτ
≥ e t,s +
(K − Sτ ) − p0 (St )σSt dWt
0
sup E g(Sτt,s ) .
P (t, s) := (4.21)
t
τ ∈T[t,T ]
Introduction to Viscosity
Solutions
Throughout this chapter, we provide the main tools from the theory of viscosity
solutions for the purpose of our applications to stochastic control problems. For
a deeper presentation, we refer to the excellent overview paper by Crandall,
Ischii and Lions [3].
The full importance of this condition will be made clear in Proposition 5.2 below.
The first step towards the definition of a notion of weak solution to (E) is
the introduction of sub and supersolutions.
Definition 5.1. A function u : O −→ R is a classical supersolution (resp.
subsolution) of (E) if u ∈ C 2 (O) and
65
66 CHAPTER 5. VISCOSITY SOLUTIONS
Proposition 5.2. Let u be a C 2 (O) function. Then the following claims are
equivalent.
(i) u is a classical supersolution (resp. subsolution) of (E)
(ii) for all pairs (x0 , ϕ) ∈ O × C 2 (O) such that x0 is a minimizer (resp. maxi-
mizer) of the difference u − ϕ) on O, we have
u∗ (x) = lim
0
inf u(x0 ) , u∗ (x) = lim sup u(x0 ) .
x →x x0 →x
We are now ready for the definition of viscosity solutions. Observe that Claim
(ii) in the above proposition does not involve the regularity of u. It therefore
suggests the following weak notion of solution to (E).
Definition 5.3. Let u : O −→ R be a locally bounded function.
(i) We say that u is a (discontinuous) viscosity supersolution of (E) if
for all pairs (x0 , ϕ) ∈ O × C 2 (O) such that x0 is a minimizer of the difference
(u∗ − ϕ) on O.
(ii) We say that u is a (discontinuous) viscosity subsolution of (E) if
for all pairs (x0 , ϕ) ∈ O × C 2 (O) such that x0 is a maximizer of the difference
(u∗ − ϕ) on O.
(iii) We say that u is a (discontinuous) viscosity solution of (E) if it is both a
viscosity supersolution and subsolution of (E).
Notation We will say that F x, u∗ (x), Du∗ (x), D2 u∗ (x) ≥ 0 in the viscosity
sense whenever u∗ is a viscosity supersolution of (E). A similar notation will be
used for subsolution.
Remark 5.4. An immediate consequence of Proposition 5.2 is that any classical
solution of (E) is also a viscosity solution of (E).
Remark 5.5. Clearly, the above definition is not changed if the minimum or
maximum are local and/or strict. Also, by a density argument, the test function
can be chosen in C ∞ (O).
5.3. First properties 67
In Section 6.1, we will show that the value function V is a viscosity solution
of the DPE (2.18) under the conditions of Theorem 2.6 (except the smoothness
assumption on V ). We also want to emphasize that proving that the value
function is a viscosity solution is almost as easy as proving that it is a classical
solution when V is known to be smooth.
The main difficulty in the theory of viscosity solutions is the interpretation
of the equation in the viscosity sense. First, by weakening the notion of solution
to the second order nonlinear PDE (E), we are enlarging the set of solutions,
and one has to guarantee that uniqueness still holds (in some convenient class
of functions). This issue will be discussed in the subsequent Section 5.4.
where
We leave the easy proof of this proposition to the reader. The next result
shows how limit operations with viscosity solutions can be performed very easily.
A particular change of function in the previous proposition is the multiplica-
tion by a scalar. Another useful tool is the addition, or convex combination, of
viscosity sub or supersolutions. Of course, this is possible only if the equation is
suitable for such operations. For this reason, the following result specializes the
discussion to the case where the nonlinearity F (x, r, p, A) is convex in (r, p, A).
By standard convex analysis, this means that F can be expressed as:
n 1 o
F (x, r, p, A) = sup − fγ (x) + kγ (x)r − bγ (x) · p − Tr σγ (x)2 A (5.1)
,
γ∈G 2
Since u2 is a classical subsolution of the equation (E), and F is convex in (r, p, A),
we now compute directly that
Notice that, in the last simplified proof, the technical condition on the coef-
ficients (fγ , kγ , bγ , σγ )γ of the representation (5.1) of F has not been used.
Consider some r > 0 together with the closed ball B̄ with radius r, centered at
x̄. Of course, we may choose |xn − x̄| < r for all n ≥ 0. Let x̄n be a minimizer
of uεn − ϕ on B̄. We claim that
Before verifying this, let us complete the proof. We first deduce that x̄n is an
interior point of B̄ for large n, so that x̄n is a local minimizer of the difference
uεn − ϕ. Then :
and the required result follows by taking limits and using the definition of F .
It remains to prove Claim (5.2). Recall that (xn )n is valued in the compact
set B̄. Then, there is a subsequence, still named (xn )n , which converges to some
x̃ ∈ B̄. We now prove that x̃ = x̄ and obtain the second claim in (5.2) as a
by-product. Using the fact that x̄n is a minimizer of uεn − ϕ on B̄, together
with the definition of u, we see that
We now obtain (5.2) from the fact that x̄ is a strict minimizer of the difference
(u − ϕ). ♦
Observe that the passage to the limit in partial differential equations written
in the classical or the generalized sense usually requires much more technicalities,
as one has to ensure convergence of all the partial derivatives involved in the
equation. The above stability result provides a general method to pass to the
limit when the equation is written in the viscosity sense, and its proof turns out
to be remarkably simple.
A possible application of the stability result is to establish the convergence
of numerical schemes. In view of the simplicity of the above statement, the
notion of viscosity solutions provides a nice framework for such questions.
(u − ϕn )(xn , yn ) = min(u − ϕn ) ,
I×J
Before proving this, let us complete the proof. Since (x0 , y0 ) is an interior point
of A × B, it follows from the viscosity property of u that
Taking the limits, this provides:it follows from the lower semicontinuity of u
that
Since u is lower semicontinu, this implies that u(x̄, ȳ)−f (ȳ)+lim inf n→∞ n|xn −
x0 |2 ≤ u(x0 , y0 ) − f (y0 ). Then, we must have x̄ = x0 , and
M := sup (u − v) ≤ 0. (5.6)
cl(O)
Assume to the contrary that M > 0. Then since cl(O) is a compact subset of
Rd , and u − v ≤ 0 on ∂O, we have:
where the last inequality follows crucially from the ellipticity of F . This provides
the desired contradiction, under our condition that F is strictly increasing in r.
♦
The objective of this section is to mimic the previous proof in the sense of
viscosity solutions. We first start by the case of first order equations where the
beautiful trick of doubling variables allows for an immediate adaptation of the
previous argument. However, more work is needed for the second order case.
where the concept of doubling variables separates the dependence of the func-
tions u and v on two different variables, and the penalization term involves the
parameter n which is intended to mimick the maximization problem (5.6) for
large n.
5.4. Comparison results 73
Proposition 5.12. Let O be a bounded open subset of Rd . Assume that the non-
lineary F (x, r, p) is independent of the A−variable, Lipshitz in the x−variable
uniformly in (r, p), and that there is a constant c > 0 such that
F (x, r0 , p) − F (x, r, p) ≥ c(r0 − r) for all r0 ≥ r, x ∈ cl(O), p ∈ Rd . (5.9)
Let u, v ∈ C 0 cl(O) be viscosity subsolution and supersolution of (E), respec-
tively, with u ≤ v on ∂O. Then u ≤ v on cl(O).
Proof. Suppose to the contrary that η := (u − v)(x0 ) > 0 for some x0 ∈ O, so
that the maximum value in (5.8) Mn ≥ η. Since u and v are continuous, we
may find for each n ≥ 1 a maximizer (xn , yn ) ∈ cl(O)2 of the problem (5.8):
n
Mn = u(xn ) − v(yn ) − |xn − yn |2 .
2
We shall prove later that
2εn := n|xn − yn |2 −→ 0, as n → ∞, and xn , yn ∈ O for large n. (5.10)
Observe that (5.8) implies that xn is a maximizer of the difference (u − ϕ), and
yn is a minimizer of the difference (v − ψ), where
n n
ϕ(x) := v(yn ) + |x − yn |2 , x ∈ cl(O), ψ(y) := u(xn ) − |xn − y|2 , x ∈ cl(O).
2 2
Then, it follows from the viscosity properties of u and v that
F xn , u(xn ), Dϕ(xn ) ≤ 0 ≤ F yn , v(yn ), Dψ(yn ) , for large n.
Using the Lipshitz property of F in x and the increaseness property (5.9), this
implies that
0 ≥ F xn , v(yn ) + Mn + εn , n(xn − yn ) − F yn , v(yn ), n(xn − yn )
≥ −|Fx |L∞ |xn − yn | + c(Mn + εn )
In view of (5.10) and the fact that Mn ≥ η > 0, this leads to the required
contradiction.
It remains to justify (5.10). Let x∗ ∈ cl(O) ba a maximizer of (u − v)
on cl(O), and denote by mn (x, y) the objective function in the maximization
problem (5.8). Let (x̄, ȳ) be any accumulation point of the bounded sequence
(xn , yn )n , i.e. (xnk , ynk ) −→ (x̄, ȳ) for some subsequence (nk )k . Then, it follows
from the obvious inequality Mn = mn (xn , yn ) ≥ mn (x∗ , x∗ ) that
nk
lim sup |xnk − ynk |2 ≤ u(xnk ) − v(ynk ) − (u − v)(x∗ )
k→∞ 2
≤ u(x̄) − v(ȳ) − (u − v)(x∗ ),
by the upper-semicontinuity of u and the slower-semicontinuity of v. This shows
that x̄ = ȳ, and it follows from the definition of x∗ as a maximizer of (u − v)
that
nk
lim sup |xnk − ynk |2 ≤ (u − v)(x̄) − (u − v)(x∗ ) ≤ 0,
k→∞ 2
74 CHAPTER 5. VISCOSITY SOLUTIONS
In the previous proof, one can easily see that the condition that F is uni-
formly Lipshitz in x can be weakened by exploiting the stronger convergence
of |xn − yn | to 0 in (5.10). For instance, the previous comparison result holds
true under the monotonicity condition (5.9) together with the locally Lipschitz
condition:
−
Motivated by this observation, we introduce the subjet JO v(x0 ) by
n o
−
JO v(x0 ) := (p, A) ∈ Rd × Sd : v(x) ≥ v(x0 ) + q p,A (x − x0 ) + ◦ |x − x0 |2 .
(5.11)
+
Similarly, we define the superjet JO u(x0 ) of a function u ∈ USC(O) at the point
x0 ∈ O by
n o
+
JO u(x0 ) := (p, A) ∈ Rd × Sd : u(x) ≤ u(x0 ) + q p,A (x − x0 ) + ◦ |x − x0 |2
(5.12)
Then, it can be proved that a function v ∈ LSC(O) is a viscosity supersolution
of the equation (E) if and only if
−
F (x, v(x), p, A) ≥ 0 for all (p, A) ∈ JO v(x).
5.4. Comparison results 75
± ±
where (xn , pn , An ) ∈ Graph(JO w) means that (pn , An ) ∈ JO w(xn ). The follow-
ing result is obvious, and provides an equivalent definition of viscosity solutions.
Proposition 5.13. Consider an elliptic nonlinearity F , and let u ∈ USC(O),
v ∈ LSC(O).
(i) Assume that F is lower-semicontinuous. Then, u is a viscosity subsolution
of (E) if and only if:
and the following inequality holds in the sense of symmetric matrices in S2d :
A 0
− ε−1 + D2 φ(x0 , y0 ) I2d ≤ ≤ D2 φ(x0 , y0 ) + εD2 φ(x0 , y0 )2 .
0 −B
76 CHAPTER 5. VISCOSITY SOLUTIONS
Remark 5.15. If u and v were C 2 functions in Lemma 5.14, the first and second
order condition for the maximization problem (5.13) with the test function (5.14)
is Du(x0 ) = α(x0 − y0 ), Dv(x0 ) = α(x0 − y0 ), and
2
D u(x0 ) 0 Id −Id
≤ α .
0 −D2 v(y0 ) −Id Id
Hence, the right-hand side inequality in (5.16) is worsening the previous second
order condition by replacing the coefficient α by 3α. ♦
Remark 5.16. The right-hand side inequality of (5.16) implies that
A ≤ B. (5.17)
To see this, take an arbitrary ξ ∈ Rd , and denote by ξ T its transpose. From
right-hand side inequality of (5.16), it follows that
A 0 ξ
0 ≥ (ξ T , ξ T ) = (Aξ) · ξ − (Bξ) · ξ.
0 −B ξ
♦
Before turning to the comparison result for second order equations, let us
go back to the
α|xα − yα |2 −→ 0, xα −→ x0 , Mα −→ (u − ϕ)(x0 ),
(5.18)
and u1 (xα ) −→ u1 (x0 ), u2 (yα ) −→ u2 (x0 ).
λ1 D2 ϕ(xα )
A1 0 Id −Id 0
≤ kα + .
0 A2 −Id Id 0 λ2 D2 ϕ(yα )
0 ≥ λ1 Fγ xα , u1 (xα ), λ−1 −1
1 Dx φ(xα , yα ) + Dϕ(xα ), λ1 A1
+λ2 Fγ yα , u2 (yα ), λ−1 −1
2 Dy φ(xα , yα ) + Dϕ(yα ), λ2 A2
= −λ1 fγ (xα ) − λ2 fγ (yα ) + λ1 kγ (xα )u1 (xα ) + λ2 kγ (yα )u2 (yα )
−λ1 bγ (xα ) · Dϕ(xα ) − λ2 bγ (yα ) · Dϕ(yα )
1
− Tr σγ2 (xα )A1 + σγ2 (yα )A2 ,
2
where we used the fact that Dx φ = −Dy φ = α(x − y). Then, fγ , kγ , bγ are
continuous, (5.18) yields
Tr σγ2 (xα )A1 + σγ2 (yα )A2 = Tr σγ2 (xα )D2 ϕ(xα ) + σγ2 (yα )D2 ϕ(yα )
+3Lγ α|xα − yα |2 ,
1
0 ≥ −fγ (x0 ) + kγ (x0 )u(x0 ) − bγ (x0 ) · Dϕ(x0 ) − Tr[σγ2 (x0 )D2 ϕ(x0 )].
2
The required viscosity subsolution property follows from the arbitrariness of
γ ∈ G.
3. It remains to justify (5.18). As (xα , yα )α is bounded, we may find a con-
verging sequence (xαn , yαn )n to some imiting point (x̂, ŷ). By the definition of
(xα , yα ) as the maximizer of φ, we have
Remark 5.18. Assumption 5.17 (ii) implies that the nonlinearity F is elliptic.
To see this, notice that for A ≤ B, ξ, η ∈ Rd , and ε > 0, we have
Aξ · ξ − (B + εId )η · η ≤ Bξ · ξ − (B + εId )η · η
= 2η · B(ξ − η) + B(ξ − η) · (ξ − η) − ε|η|2
≤ ε−1 |B(ξ − η)|2 + B(ξ − η) · (ξ − η)
|B| 1 + ε−1 |B| |ξ − η|2 .
≤
For 3α ≥ (1 + ε−1 |B|)|B|, the latter inequality implies the right-hand side of
(5.16) holds true with (A, B + εId ). For ε sufficiently small, the left-hand side
of (5.16) is also true with (A, B + εId ) if in addition α > |A| ∨ |B|. Then
u ≤ v on ∂O =⇒ u ≤ v on Ō := cl(O).
Step 1: For every α > 0, it follows from the upper-semicontinuity of the differ-
ence (u − v) and the compactness of Ō that
n α o
Mα := sup u(x) − v(y) − |x − y|2
O×O 2
α
= u(xα ) − v(yα ) − |xα − yα |2 (5.20)
2
80 CHAPTER 5. VISCOSITY SOLUTIONS
Step 3: We first use the strict monotonicity Assumption 5.17 (i) to obtain:
γδ ≤ γ u(xn ) − v(yn ) ≤ F (xn , u(xn ), αn (xn − yn ), An )
−F (xn , v(yn ), αn (xn − yn ), An ) .
γδ ≤ $ αn |xn − yn |2 + |xn − yn | .
Sending n to infinity, this leads to the desired contradiction of (5.19) and (5.21).
Step 4: It remains to prove the claims (5.21). By the upper-semicontinuity of
the difference (u − v) and the compactness of Ō, there exists a maximizer x∗ of
the difference (u − v). Then
αn
(u − v)(x∗ ) ≤ Mαn = u(xn ) − v(yn ) − |xn − yn |2 .
2
Sending n → ∞, this provides
1
`¯ := lim sup αn |xn − yn |2 ≤ lim sup u(xαn ) − v(yαn ) − (u − v)(x∗ )
2 n→∞ n→∞
≤ u(x̂) − v(ŷ) − (u − v)(x∗ );
0 ≤ `¯ ≤ (u − v)(x̂) − (u − v)(x∗ ) ≤ 0.
5.4. Comparison results 81
where
1
α|x − y|2 + ε|x|2 + ε|y|2 .
φ(x, y) :=
2
1. Since u(x) = ◦(|x|2 ) and v(y) = ◦(|y|2 ) at infinity, there is a maximizer
(xα , yα ) for the previous problem:
Mα = u(xα ) − v(yα ) − φ(xα , yα ).
Moreover, there is a sequence αn → ∞ such that
(xn , yn ) := (xαn , yαn ) −→ (x̂ε , ŷε ),
and, similar to Step 4 of the proof of Theorem 5.19, we can prove that x̂ε = ŷε ,
αn |xn − yn |2 −→ 0, and Mαn −→ M∞ := sup (u − v)(x) − ε|x|2 . (5.26)
x∈Rd
Notice that Mαn ≥ (u − v)(z) − φ(z, z) ≥ δ − ε|z|2 > 0, by (5.25). then, for
sufficiently small ε > 0, we have
0 < δ − ε|z|2 ≤ lim sup Mαn = lim sup {u(xn ) − v(yn ) − φ(xn , yn )}
n→∞ n→∞
≤ lim sup {u(xn ) − v(yn ))}
n→∞
≤ lim sup u(xn ) − lim inf v(yn )
n→∞ n→∞
≤ (u − v)(x̂ε ).
5.4. Comparison results 83
ε|x̂ε | −→ 0 as ε & 0,
which we now prove. Indeed, we have that (u−v)(x̂ε )−ε|x̂ε |2 ≥ (u−v)z −ε|z|2 ,
and therefore
(u − v)(x̂ε ) δ
ε|x̂ε | ≤ − for small ε > 0.
|x̂ε | 2|x̂ε |
Using our condition |u(x)| = ◦(|x|) and |v(x)| = ◦(|x|) at infinity, we see that
whenever |x̂ε | is unbounded in ε, the last inequality implies that ε|x̂ε | → 0. ♦
Proof. For each ε > 0, define W (x) := U (x) + εx, x ∈ O. Then W satisfies in
the viscosity sense DW ≥ ε in O, i.e. for all (x0 , ϕ) ∈ O × C 1 (O) such that
together with the boundary conditions v(x1 ) = v(x2 ) = W (x2 ). Observe that
W is a lower semicontinuous viscosity supersolution of the above equation. From
the comparison result of Example 5.20, this implies that
sup (v − W ) = max {(v − W )(x1 ), (v − W )(x2 )} ≤ 0.
[x1 ,x2 ]
Hence W (x) ≥ v(x) = W (x2 ) for all x ∈ [x1 , x2 ]. Applying this inequality at
x0 ∈ (x1 , x2 ), and recalling that the test function ϕ is strictly increasing on
[x1 , x2 ], we get :
(W − ϕ)(x0 ) > (W − ϕ)(x2 ),
contradicting (5.31). ♦
Lemma 5.25. Let O be an open interval of R, and U : O −→ R be a lower
semicontinuous viscosity supersolution of the equation −D2 U ≥ 0 on O. Then
U is concave on O.
Proof. Let a < b be two arbitrary elements in O, and consider some ε > 0
together with the function
√ √ √ √
U (a) e ε(b−s) − e− ε(b−s) + U (b) e ε(s−a) − e− ε(s−a)
v(s) := √ √ for a ≤ s ≤ b.
e ε(b−a) − e− ε(b−a)
Clearly, v solves the equation
εv − D2 v = 0 on (a, b), v=U on {a, b}.
Since U is lower semicontinuous it is bounded from below on the interval [a, b].
Therefore, by possibly adding a constant to U , we can assume that U ≥ 0, so
that U is a lower semicontinuous viscosity supersolution of the above equation.
It then follows from the comparison theorem 6.6 that :
sup (v − U ) = max {(v − U )(a), (v − U )(b)} = 0.
[a,b]
Hence,
√ √ √ √
U (a) e ε(b−s) − e− ε(b−s) + U (b) e ε(s−a) − e− ε(s−a)
U (s) ≥ v(s) = √ √
e ε(b−a) − e− ε(b−a)
Observe that
λ 2 λ
v̂ λ (x) + v(y) − |y|2 + λx · y
|x| = sup
2 y∈RN 2
(w − ϕ)(x0 ) = max
2
(w − ϕ).
O
Proof. Step 1: We first observe that we may reduce the problem to the case
1
O = Rd , x0 = 0, u1 (0) = u2 (0) = 0, and ϕ(x) = Ax · x for some A ∈ Sd .
2
The reduction to x0 = 0 follows from an immediate change of coordinates.
Choose any compact subset of K ⊂ O containing the origin and set ūi = ui on
K and −∞ otherwise, i = 1, 2. Then, the problem can be stated equivalently in
terms of the functions ūi which are now defined on Rd and take values on the
extended real line. Also by defining
¯1 (x1 )+ū
ū ¯2 (x2 )−ϕ̄¯(x1 , x2 ) < ū
¯1 (x1 )+ū
¯2 (x2 )−ϕ̄(x1 , x2 ) ≤ ū
¯1 (0)+ū
¯2 (0)−ϕ̄(0) = 0.
2w(x) ≤ Ax · x
= A(x − y) · (x − y)Ay · y − 2Ay · (y − x)
1
≤ A(x − y) · (x − y)Ay · y + εA2 y · y + |x − y|2
ε
1
= A(x − y) · (x − y) + |x − y| + (A + εA2 )y · y
2
ε
≤ (ε−1 + |A|)|x − y|2 + (A + εA2 )y · y.
Set λ := ε−1 + |A| and B := A + εA2 . The latter inequality implies the following
property of the sup-convolution:
1
ŵλ (y) − By · y ≤ ŵ(0) = 0.
2
Step 3: Recall from (5.33) that ŵλ is λ−semiconvex. Then, it follows from
2,+ 2,−
Lemma 5.26 that there exist X ∈ S2d such that (0, X) ∈ J ŵλ (0) ∩ J ŵλ (0)
and −λI2d ≤ X ≤ B. Moreover, it is immediately checked that ŵλ (x1 , x2 ) =
ûλ1 (x1 ) + ûλ2 (x2 ), implying that X is bloc-diagonal with blocs X1 , X2 ∈ Sd .
Hence:
X1 0
−(ε−1 + |A|)I2d ≤ ≤ A + εA2
0 X2
2,+ λ
and (0, Xi ) ∈ J ûi (0) for i = 1, 2 which, by Lemma 5.27 implies that (0, Xi ) ∈
2,+
J uλi (0). ♦
88 CHAPTER 5. VISCOSITY SOLUTIONS
We coninue by turning to the proofs of Lemmas 5.26 and 5.27. The main
tools which will be used are the following properties of any semiconvex function
ϕ : RN −→ R whose proofs are reported in [3]:
• Aleksandrov lemma: ϕ is twice differentiable a.e.
• Jensen’s lemma: if x0 is a strict maximizer of ϕ, then for every r, δ > 0,
the set
x̄ ∈ B(x0 , r) : x 7−→ ϕ(x)+p·x has a local maximum at x̄ for some p ∈ Bδ
Proof of Lemma 5.26 Notice that ϕ(x) := f (x) − 21 Bx · x − |x|4 has a strict
maximum at x = 0. Localizing around the origin, we see that ϕ is a semiconvex
function. Then, for every δ > 0, by the above Aleksandrov and Jensen lemmas,
there exists qδ and xδ such that
We may then write the first and second order optimality conditions to see that:
and, in view of (5.34), we may send δ to zero along some subsequence and obtain
a limit point (0, X) ∈ J¯2,+ f (0) ∩ J¯2,− f (0). ♦
Dynamic Programming
Equation in Viscosity
Sense
b : (t, x, u) ∈ S × U −→ b(t, x, u) ∈ Rn
91
92 CHAPTER 6. DPE IN VISCOSITY SENSE
and
for some constant K independent of (t, x, y, u). For each admissible control
process ν ∈ U0 , the controlled stochastic differential equation :
has a unique solution X, for all given initial data ξ ∈ L2 (F0 , P) with
ν 2
E sup |Xs | < C(1 + E[|ξ|2 ])eCt for all t ∈ [0, T ] (6.4)
0≤s≤t
for some constant C. Finally, the gain functional is defined via the functions:
f, k : [0, T ) × Rd × U −→ R and g : Rd −→ R
for some constant K independent of (t, x, u). The cost function J on [0, T ] ×
Rd × U is:
"Z #
T
J(t, x, ν) := E β ν (t, s)f (s, Xst,x,ν , νs )ds + β ν (t, T )g(XTt,x,ν ) , (6.5)
t
and {Xst,x,ν , s ≥ t} is the solution of (6.3) with control process ν and initial
condition Xtt,x,ν = x. The stochastic control problem is defined by the value
function:
on [0, T ) × Rd .
ηn := V (tn , xn ) − ϕ(tn , xn ) −→ 0.
where α > 0 is some given constant, B denotes the unit ball of Rn , and
p
hn := |ηn |1{ηn 6=0} + n−1 1{ηn =0} .
Notice that θn −→ t as n −→ ∞.
1. From the first inequality in the dynamic programming principle of Theorem
2.3, it follows that:
" Z θn #
n n
0 ≤ E V (tn , xn ) − β(tn , θn )V∗ (θn , Xθn ) − β(tn , r)f (r, Xr , νr )dr .
tn
Now, in contrast with the proof of Proposition 2.4, the value function is not
known to be smooth, and therefore we can not apply Itô’s formula to V . The
94 CHAPTER 6. DPE IN VISCOSITY SENSE
We now send n to infinity. The a.s. convergence of the random value in-
side the expectation is easily obtained by the mean value Theorem; recall
that for n ≥ N (ω) sufficiently large, θn (ω) = hn . Since the random vari-
R θn
able h−1
n t
β(tn , r)(L· ϕ − f )(r, Xrn , u)dr is essentially bounded, uniformly in
n, on the stochastic interval [tn , θn ], it follows from the dominated convergence
theorem that :
We next wish to show that V satisfies the nonlinear partial differential equa-
tion (6.9) with equality, in the viscosity sense. This is also obtained by a slight
modification of the proof of Proposition 2.5.
Proposition 6.3. Assume that the value function V is locally bounded on S.
Let the function H be finite and upper semicontinuous on [0, T ) × Rd × Rd × Sd ,
and kk + k∞ < ∞. Then, V is a viscosity subsolution of the equation
on [0, T ) × Rn .
h := ∂t ϕ + H ., ϕ, Dϕ, D2 ϕ < 0 on Nr .
(6.13)
and we observe that θnν , Xθtnν ,xn ,ν ∈ ∂Nr by the pathwise continuity of the
n
controlled process. Then, with βsν := β ν (tn , s), it follows from (6.14) that:
V (tn , xn ) ≥ −η + ϕ(tn , xn )
ν
h Z θn i
−η + E βθνnν ϕ θnν , Xθtnν ,xn ,ν − βsν (∂t + Lνs )ϕ s, Xstn ,xn ,ν ds
=
n
tn
ν
h Z θn i
−η + E βθνnν ϕ θnν , Xθtnν ,xn ,ν + βsν f (., νs ) − h s, Xstn ,xn ,ν ds
≥
n
tn
ν
h Z θn i
ν ν tn ,xn ,ν
≥ −η + E βθnν ϕ θn , Xθν + βsν f s, Xstn ,xn ,ν , νs ds
n
tn
−∂t V − H ·, V, DV, D2 V = 0 on S.
(6.17)
The partial differential equation (6.17) has a very simple and specific depen-
dence in the time-derivative term. Because of this, it is usually referred to as a
parabolic equation.
In order to a obtain a characterization of the value function by means of
the dynamic programming equation, the latter viscosity property needs to be
complemented by a uniqueness result. This is usually obtained as a consequence
of a comparison result.
In the present situation, one may verify the conditions of Theorem 5.23. For
completeness, we report a comparison result which is adapted for the class of
equations corresponding to stochastic control problems.
Consider the parabolic equation:
where BS (t, x; γ) is the collection of elements (s, y) in S such that |t−s|2 +|x−y|2
≤ γ 2 . We report, without proof, the following result from [6] (Theorem V.8.1
and Remark V.8.1).
Assumption 6.5. The above operators satisfy:
lim supε&0 {G+γε (tε , xε , pε , Aε ) − G−γε (sε , yε , pε , Bε )}
(6.19)
≤ Const (|t0 − s0 | + |x0 − y0 |) [1 + |p0 | + α (|t0 − s0 | + |x0 − y0 |)]
Theorem 6.6. Let Assumption 6.5 hold true, and let u ∈ USC(S̄), v ∈ LSC(S̄)
be viscosity subsolution and supersolution of (6.18), respectively. Then
sup(u − v) = sup(u − v)(T, ·).
S Rn
A sufficient condition for (6.19) to hold is that f (·, ·, u), k(·, ·, u), b(·, ·, u),
and σ(·, ·, u) ∈ C 1 (S) with
kbt k∞ + kbx k∞ + kσt k∞ + kσx k∞ < ∞
|b(t, x, u)| + |σ(t, x, u)| ≤ Const(1 + |x| + |u|) ;
see [6], Lemma V.8.1.
Here, X t,x denotes the unique strong solution of (3.1) with initial condition
Xtt,x = x. Then, the optimal stopping problem is defined by:
V (t, x) := sup J(t, x, τ ) for all (t, x) ∈ S. (6.22)
τ ∈T[t,T ]
The next result derives the dynamic programming equation for the previous
optimal stopping problem in the sense of viscosity solution, thus relaxing the
C 1,2 regularity condition in the statement of Theorem 3.4. As usual, the same
methodology allows to handle seemingly more general optimal stopping prob-
lems:
V̄ (t, x) := ¯ x, τ ),
sup J(t, (6.23)
τ ∈T[t,T ]
where
"Z #
T
¯ x, τ )
J(t, := E β(t, s)f (s, Xst,x )ds + β(t, τ )g(Xτt,x )1{τ <∞} ,
t
Z s
β(t, s) := exp − k(u, Xut,x )du .
t
98 CHAPTER 6. DPE IN VISCOSITY SENSE
Then θhnn ∈ T[ttn ,T ] for sufficiently small hn , and it follows from (3.10) that:
h i
V (tn , xn ) ≥ E V∗ θhnn , Xθhnn .
We continue by fixing
p
hn := |ηn |1{ηn 6=0} + n−1 1{ηn =0} ,
as in the proof of Proposition 6.2. Then, the rest of the proof follows exactly the
line of argument of the proof of Theorem 3.4 combined with that of Proposition
6.2.
(ii) We next prove that V is a viscosity subsolution of the equation (6.24). Let
(t0 , x0 ) ∈ S and ϕ ∈ C 2 (S) be such that
V (tn , xn ) = ηn + ϕ(tn , xn )
" #
Z τ ∧θn
= ηn + E ϕ (τ ∧ θn , Xτ ∧θn ) − (∂t + A)ϕ(t, Xt )dt ,
tn
where the diffusion term has zero expectation because the process (t, Xttn ,xn ) is
confined to the compact subset Nh on the stochastic interval [tn , τ ∧ θn ]. Since
−(∂t + A)ϕ ≥ 0 on Nh by (6.25), this provides:
V (tn , xn ) ≥ ηn + E ϕ (τ, Xτ ) 1{τ <θn } + ϕ (θn , Xθn ) 1{τ ≥θn }
≥ ηn + E (g (Xτ ) + δ) 1{τ <θn } + (V ∗ (θn , Xθn ) + γ) 1{θn ≥τ }
Proof. This is an easy adaptation of the proof of Theorem 5.23. We adapt the
same notations so that, in the present, x stands for the pair (t, x). The only
difference appears at Step 3 which starts from the fact that
Stochastic Target
Problems
7.1.1 Formulation
Let T > 0 be the finite time horizon and W = {Wt , 0 ≤ t ≤ T } be a d-
dimensional Brownian motion defined on a complete probability space (Ω, F, P).
We denote by F = {Ft , 0 ≤ t ≤ T } the P-augmentation of the filtration gener-
ated by W .
We assume that the control set U is a convex compact subset of Rd with
non-empty interior, and we denote by U the set of all progressively measurable
processes ν = {νt , 0 ≤ t ≤ T } with values in U .
The state process is defined as follow: given the initial data z = (x, y) ∈
Rd × R, an initial time t ∈ [0, T ], and a control process ν ∈ U, let the controlled
process Z t,z,ν = (X t,x,ν , Y t,z,ν ) be the solution of the stochastic differential
equation :
101
102 CHAPTER 7. STOCHASTIC TARGET PROBLEMS
(7.2)
Remark 7.1. By introducing the subset of control processes:
we may re-write the value function of the stochastic target problem into:
This follows from the fact that the state process X t,x,ν is independent of y and
YTt,x,y,ν is nondecreasing in y.
Proof. Let z = (x, y) and ν ∈ A(t, z), and denote Zt,z,ν := (X t,x,ν , Y t,z,ν ).
t,z,ν t,x,ν
Then YT ≥ g XT P−a.s. Notice that
τ,Zτt,z,ν ,ν
ZTt,z,ν = ZT .
Then, by the definition of the set A, it follows that ν ∈ A (τ, Zτt,z,ν ), and
therefore V (τ, Xτt,x,ν ) ≤ Yτt,z,ν , P−a.s. ♦
In the next subsection, we will prove that the value function V is a viscosity
supersolution of the corresponding dynamic programming equation which will
be obtained as the infinitesimal counterpart of (7.3). The following remark
comments on the full geometric dynamic programming principle in the context
of stochastic target problems.
Remark 7.3. The statement (7.3) in Theorem 7.2 can be strengthened to the
following geometric dynamic programming principle:
n o
V (t, x) = inf y ∈ R : Y t,x,y,ν ≥ V τ, Xτt,x,ν , P − a.s. for some ν ∈ U0 .
(7.4)
Let us provide an intuitive justification of this. Denote ŷ := V (t, x). In view of
(7.3), it is easily seen that (7.4) is implied by
In words, there is no control process ν which allows to reach the value function
V (τ, Xτt,x,ν ) at time τ , with full probability, starting from an initial data stricly
below the value function V (t, x). To see this, suppose to the contrary that there
exist ν ∈ U0 , η > 0, and τ ∈ T[t,T ] such that:
In view of Remark 7.1, this implies that Yτt,x,ŷ−η,ν ∈ Y (τ, Xτt,x,ν ), and therefore,
there exists a control ν̂ ∈ U0 such that
τ,Z t,x,ŷ−η,ν ,ν̂ τ,X t,x,ν ,ν̂
YT τ ≥ g XT τ , P − a.s.
t,x,y ∗ −η,ν
t,x,ν
Since the process X τ,Xτ ,ν̂
, Y τ,Zτ ,ν̂
depends on ν̂ only through its
realizations in the stochastic interval [t, θ], we may chose ν̂ so as ν̂ = ν on
τ,Z t,x,ŷ−η,ν ,ν̂
[t, τ ] (this is the difficult part of this proof). Then ZT τ = ZTt,x,ŷ−η,ν̂
and therefore ŷ − η ∈ Y(t, x), hence ŷ − η ≤ V (t, x). Recall that by definition
ŷ = V (t, x) and η > 0. ♦
104 CHAPTER 7. STOCHASTIC TARGET PROBLEMS
where ψ was introduced in (7.1). In particular, this implies that ψ(t, x, DV (t, x)) ∈
U . By (7.7), this is equivalent to:
−∂t V (t, x) + H t, x, V (t, x), DV (t, x), D2 V (t, x), ψ(t, x, DV (t, x)) ≥ (7.14)
0.
Plugging this into (7.19) and applying Itô’s formula, we then see that:
Z t∧θn Z t∧θn
n
(ε ∧ γ)1{t=θn } ≤ cn + δsn ds + N ν̂s (s, Xsn , Dϕ(s, Xsn )) · dWs
tn tn
Z t∧θn
≤ Mn := cn + δsn 1An (s)ds (7.20)
tn
Z t∧θn
n
+ N ν̂s (s, Xsn , Dϕ(s, Xsn )) · dWs
tn
7.1. Stochastic target 107
where
δsn := −∂t ϕ(s, Xsn ) + H s, Zsn , Dϕ(s, Xsn ), D2 ϕ(s, Xsn ), ν̂s
and
we see that the process M n Ln is a local martingale which is bounded from below
by the constant ε ∧ γ. Then M n Ln is a supermartingale, and it follows from
(7.20) that
which can not happen because cn −→ 0. Hence, our starting point (7.22) can
not happen, and the proof of the supersolution property is complete.
γn := yn − ϕ(tn , xn ) −→ 0. (7.25)
108 CHAPTER 7. STOCHASTIC TARGET PROBLEMS
where the drift term α(·) ≥ η is defined in (7.23) and the diffusion coefficient
vanishes by the definition of the function ψ in (7.1). Since ε, ζ > 0 and γn → 0,
this implies that
Y n (θn ) ≥ V (θn , X n (θn )) for sufficiently large n.
Recalling that the initial position of the process Y n is yn = V (tn , xn ) − n−1 <
V (tn , xn ), this is clearly in contradiction with the second part of the geometric
dynamic programming principle discussed in Remark 7.3. ♦
Remark 7.5. We observe that the normalization of the first asset to unity
does not entail any loss of generality as we can always reduce to this case by
discounting or, in other words, by taking the price process of this asset as a
numéraire.
Also, the formulation of the above process X as a martingale is not a re-
striction as our subsequent superhedging problem only involves the underlying
probability measure through the corresponding zero-measure sets. Therefore,
under the no-arbitrage condition (or more precisely, no free-lunch with vanish-
ing risk), we can reduce the model to the above martingale case by an equivalent
change of measure. ♦
Under the self-financing condition, the liquidation value of the portfolio is
defined by the controlled state process:
where π is the control process, with πti representing the amount invested in the
i−th risky asset X i at time t.
We introduce portfolio constraints by imposing that the portfolio process π
must be valued in a subset U of Rd . We shalll assume that
We then define the controls set by Uo as in the previous sections, and we defined
the superhedging problem under portfolio constraints by the stochastic target
problem:
the map λ 7−→ h(λ) := λδU (ξ) − V∗ (t, x ? eλξ ) is nondecreasing, (7.30)
110 CHAPTER 7. STOCHASTIC TARGET PROBLEMS
where eλξ is the vector of Rd with entries (eλξ )i = eλξi . Then h(1) ≥ h(0)
provides:
We next observe that V∗ (T, .) ≥ g (just use the definition of V , and send t % T ).
Then, we deuce from the latter inquality that
In other words, in order to superhedge the derivative security with final payoff
g(XT ), the constraints on the portfolio require that one hedges the derivative
security with larger payoff ĝ(XT ). The function ĝ is called the face-lifted payoff,
and is the smallest majorant of g which satisfies the gradient constraint x ?
Dg(x) ∈ U for all x ∈ Rd+ .
Combining (7.31) with (7.28), it follows from the comparison result for the
linear Black-Scholes PDE that
Since ĝ has linear growth, it follows that π̂ ∈ H2 . We also observe that the
random variable ln XTt,x is gaussian, so that the function v can be written in:
2
w−x+ 1 σ 2 (T −t)
Z
1 − 21 √2
w σ T −t
v(t, x) = ĝ(e ) p e dw.
2πσ 2 (T − t)
Under this form, it is clear that v is a smooth function. Then the above hedging
portfolio is given by:
In order to avoid degenerate results, we restrict the analysis to the case where the
Y process takes non-negative values, by simply imposing the following conditions
on the coefficients driving its dynamics:
and
Proposition 7.6. For all t ∈ [0, T ] and x̂ = (x, p) ∈ Rd × [0, 1], we have
n o
V̂ (t, x̂) = inf y ∈ R+ : Ĝ X̂Tt,x̂,ν̂ , YTt,x,y,ν ≥ 0 for some ν̂ = (ν, α) ∈ Û .
We first show that V̂ ≥ y > V̂ (t, x, p), we can find ν ∈ U such that
v. For
p0 := P G XTt,x,ν , YTt,x,y,ν ≥ 0 ≥ p. By the stochastic integral representation
α
Since p0 ≥ p, it follows that 1{Y t,x,y,ν ≥g(X t,x,ν ,)} ≥ Pt,p (T ), and therefore
T T
y ≥ v(t, x, p) from the definition of the problem v.
We next show that v ≥ V̂ . For y > v(t, x, p), we have Ĝ X̂Tt,x̂,ν̂ , YTt,x,y,ν ≥
α
0 for some ν̂ = (ν, α) ∈ Û. Since Pt,p is a martingale, it follows that
h i
P YTt,x,y,ν ≥ g XTt,x,ν = E 1{Y t,x,y,ν ≥g(X t,x,ν )} ≥ E PTt,p,α = p,
T T
2. It is easy to show that one can moreover restrict to controls α such that
the process P t,p,α takes values in [0, 1]. This is rather natural since this process
should be interpreted as a conditional probability, and this corresponds to the
natural domain [0, 1] of the variable p. We shall however avoid to introduce this
state constraint, and use th efact that the value function V̂ (·, p) is constant for
p ≤ 0 and equal ∞ for p > 1, see (7.36).
We assume that the coefficients µ and σ are such that X t,x ∈ (0, ∞)d P−a.s.
for all initial conditions (t, x) ∈ [0, T ] × (0, ∞)d . In order to avoid arbitrage, we
also assume that σ is invertible and that
We shall assume throughout that 0 ≤ g(x) ≤ C(1 + |x|) for all x ∈ Rd+ . By the
usual buy-and-hold hedging strategies, this implies that 0 ≤ V (t, x) ≤ C(1+|x|).
Under the above assumptions, the corresponding super-hedging cost V (t, x) :=
V̂ (t, x, 1) is continuous and is given by
t,x
g(XTt,x ) ,
V (t, x) = EQ
1
= −∂t V − Tr σ 2 Dx2 V .
0 (7.42)
2
For later use, let us denote by
Z ·
t,x
WQ := W + λ(Xst,x )ds, s ∈ [t, T ],
t
Lemma 7.8. For all x ∈ Rd+ and p ∈ [0, 1], we have V̂∗ (T, x, p) ≥ pg(x).
1{Y tn ,xn ,yn ,νn −g(X tn ,xn )≥0} ≥ PTtn ,pn ,αn .
T T
R
T RT
where we denotes LTtn ,xn := exp − tn λ(Xstn ,xn ) · dWs − 21 tn |λ(Xstn ,xn )|2 ds .
Then
yn ≥ E PTtn ,pn ,αn g(x) + E PTtn ,pn ,αn LtTn ,xn g(XTtn ,xn ) − g(x)
where we used the fact that P tn ,pn ,αn is a nonnegative martingale. Now, since
this process is also bounded by 1, we have
E PTtn ,pn ,αn LTtn ,xn g(XTtn ,xn ) − g(x) ≤ E LtTn ,xn g(XTtn ,xn ) − g(x) −→ 0
Using the above supersolution property of V̂∗ , we shall prove below that v is an
upper-semicontinuous viscosity subsolution on [0, T ) × (0, ∞)d × (0, ∞) of
1 1 2
−∂t v − Tr σ 2 Dx2 v − |λ| q 2 Dq2 v − Tr [σλDxq v] ≤ 0 (7.48)
2 2
116 CHAPTER 7. STOCHASTIC TARGET PROBLEMS
on [0, T ] × (0, ∞)d × (0, ∞), where the process Qt,x,q is defined by the dynamics
dQt,x,q
s t,x Qt,x
t,x,q = λ(Xs ) · dWs and Qt,x,q (t) = q ∈ (0, ∞). (7.51)
Qs
Given the explicit representation of v̄, we can now provide a lower bound for
the primal function V̂ by using (7.47).
We next deduce from (7.50) a lower bound for the quantile hedging problem
convp
V̂ . Recall that the convex envelop V̂∗ of V̂∗ with respect to p is given by
the bi-conjugate function:
convp
V̂∗ (t, x, p) = sup pq − v(t; x; q) ,
q
and is the largest convex minorant of V̂∗ . Then, since V̂ ≥ V̂∗ , it follows from
(7.50) that:
V̂ (t, x, p) ≥ V̂∗ (t, x, p) ≥ sup pq − v̄(t, x, q) (7.52)
q
e- The explicit solution We finally show that the above explicit minorant
y(t, x, p) is equal to V̂ (t, x, p). By the martingale representation theorem, there
exists a control process ν ∈ U0 such that
t,x,y(t,x,p),ν
≥ g XTt,x 1{q̄(t,x,p)Qt,x,1 ≥g(X t,x )} .
YT
T T
7.2. Quantile stochastic target problems 117
h i
Since P q̄(t, x, p)Qt,x,1
T ≥ g(XTt,x ) = p, by (7.53), this implies that V̂ (t, x, p) =
y(t, x, p).
on (0, T )×(0, ∞)d ×(0, ∞) ⊂ int(dom(ϕ)), where q 7→ J(·, q) denotes the inverse
of p 7→ Dp ϕ̃(·, p), recall that ϕ̃ is strictly convex in p.
We now deduce from the assumption q0 > 0 and (7.47) that we can find
p0 ∈ [0, 1] such that v(t0 , x0 , q0 ) = p0 q0 − V̂∗ (t0 , x0 , p0 ) which, by using the very
definition of (t0 , x0 , p0 , q0 ) and v, implies that
and
where the last equality follows from (7.54) and the strict convexity of the map
p 7→ pq0 − ϕ̃(t0 , x0 , p) in the domain of ϕ̃.
We conclude the proof by discussing three alternative cases depending on
the value of p0 .
• If p0 ∈ (0, 1), then (7.54) implies that ϕ̃ satisfies (7.43) at (t0 , x0 , p0 ) and
the required result follows by exploiting the link between the derivatives of
ϕ̃ and the derivatives of its p-Fenchel transform ϕ, which can be deduced
from (7.54).
• If p0 = 1, then the first boundary condition in (7.44) and (7.54) imply
that (t0 , x0 ) is a local minimizer of V̂∗ (·, 1) − ϕ̃(·, 1) = V − ϕ̃(·, 1) such
that (V − ϕ̃(·, 1))(t0 , x0 ) = 0. This implies that ϕ̃(·, 1) satisfies (7.42) at
(t0 , x0 ) so that ϕ̃ satisfies (7.43) for α = 0 at (t0 , x0 , p0 ). We can then
conclude as in 1. above.
♦
Chapter 8
119
120 CHAPTER 8. BSDE AND STOCHASTIC CONTROL
motivation was also from optimal control which was an important field of inter-
est for Shige Peng. However, the natural connections with problems in finan-
cial mathematics was very quickly realized, see Elkaroui, Peng and Quenez [?].
Therefore, a large development of the theory was achieved in connection with
financial applications and crucially driven by the intuition from finance.
the set of control processes U0 is defined as in Section 2.1, and the controlled
state process is defined by some initial date X0 and the SDE with random
coefficients:
Observe that we are not emphasizing on the time origin and the position of
the state variable X at the time origin. This is a major difference between the
dynamic programming approach, developed by the American school, and the
Pontryagin maximul principle approach of the Russian school.
For every u ∈ U , we define:
so that
∂Lu (t, x, y, z) ∂Lu (t, x, y, z)
b(t, x, u) = and σ(t, x, u) = .
∂y ∂z
We also introduce the function
dŶt = −∇x Lν̂t (t, X̂t , Ŷt , Ẑt )dt + Zt dWt , and ŶT = ∇g(X̂T ), (8.1)
8.1. Motivation and examples 121
where X̂ := X ν̂ ,
(ii) ν̂ satisfies the maximum principle:
Lν̂t (t, X̂t , Ŷt , Ẑt ) = `(t, X̂t , Ŷt , Ẑt ). (8.2)
(iii) The functions g and `(t, ., y, z) are concave, for fixed t, y, z, and
∇x Lν̂t (t, X̂t , Ŷt , Ẑt ) = ∇x `(t, X̂t , Ŷt , Ẑt ) (8.3)
where the diffusion terms have zero expectations because the processes Ŷ and
Ẑ are in H2 . By Conditions (ii) and (iii), this implies that
hZ T
J0 (ν̂) − J0 (ν) ≥ E `(t, X̂t , Ŷt , Ẑt ) − `(t, Xt , Ŷt , Ẑt )
0
i
−(X̂t − Xtν ) · ∇x `(t, X̂t , Ŷt , Ẑt ) dt
≥ 0
If existence holds for the latter problem, then there would exist a pair (Y, Z) in
H2 such that
Z T
Y0 + b(t, Wt , Yt , Zt )dt + Zt · dWt ≥ g(WT ), P − a.s.
0
Then, under the self-financing condition, the liquidation value of the portfolio
is defined by
We are then reduced to a problem of solving the BSDE (8.5)-(8.6). This problem
Rt
can be solved very easily if the process λ is so that the process {Wt + 0 λs ds, t ≥
0} is a Brownian motion under the so-called equivalent probability measure
Q. Under this condition, it suffices to get rid of the linear term in (8.5) by
discounting, then π̂ is obtained by the martingale representation theorem in the
present Brownian filtration under the equivalent measure Q.
We finally provide an example where the dependence of Y in the control
variable Z is nonlinear. The easiest example is to consider a financial market
with different lending and borrowing rates rt ≤ rt . Then the dynamics of
liquidation value of the portfolio (8.6) is replaced by the following SDE:
f : [0, T ] × Ω × Rn × Rn×d −→ R,
Hence, the process Y is in the space of continuous processes with finite S 2 −norm.
For later use, we report the following necessary and sufficient condition for
a martingale to be uniformly integrabile.
Lemma 8.2. Let M = {Mt , t ∈ [0, T )} be a scalar local martingale. Then, M
is uniformly integrable if and only if
lim λP sup |Mt | > λ = 0.
λ→∞ t≤T
Since E|M∞ | ≤ lim inf n E|Mn | = M0 < ∞, the second term on the right hand-
side convergence to E|M∞ | as λ → ∞. Since the left hand-side term is non-
decreasing in λ, we deduce that
lim E|MTλ | = p + E|M∞ | where p := lim λP sup |Mt | > λ . (8.12)
λ→∞ λ→∞ t<T
Then p = 0 iff Mθn −→ M∞ in L1 for all sequence (θn )n satisfying (8.11), which
is now equivalent to the uniform integrability of M . ♦
ft (y, z) := at + bt y + ct · z, (8.13)
we have reduced to the case where the generator does not depend on z. We
next get rid of the linear term in y by introducing:
Rt Rt Rt
bs ds bs ds bs ds
Y t := Yt e 0 so that dY t = −at e 0 dt + Zt e 0 dBt .
126 CHAPTER 8. BSDE AND STOCHASTIC CONTROL
Finally, defining
Z t Ru
bs ds
Y t := Y t + au e 0 du,
0
we arrive at a BSDE with zero generator for Y t which can be solved by the
martingale representation theorem under the equivalent probability measure Q.
Of course, one can also express the solution under P:
" Z T #
t t
Yt = E ΓT ξ + Γs as dsFt , t ≤ T,
t
where
Z s Z s Z s
1
Γts := exp bu du − |cu |2 du + cu · dWu , 0 ≤ t ≤ s ≤ T. (8.14)
t 2 t t
Theorem 8.3. Assume that {ft (0, 0), t ∈ [0, T ]} ∈ H2 and, for some constant
C > 0,
for all t ∈ [0, T ] and (y, z), (y 0 , z 0 ) ∈ Rn × Rn×d . Then, for every ξ ∈ L2 , there
is a unique solution (Y, Z) ∈ S 2 × H2 to the BSDE(ξ, f ).
Proof. Denote S = (Y, Z), and introduce the equivalent norm in the correspond-
ing H2 space:
"Z #
T
kSkα := E eαt (|Yt |2 + |Zt |2 )dt .
0
φ : s = (y, z) ∈ H2 7−→ S s = (Y s , Z s )
defined by:
Z T Z T
Yts = ξ+ fu (yu , zu )du − Zus · dWu , t ≤ T.
t t
1. First, since |fu (yu , zu )| ≤ |fu (0, 0)| + C(|yu | + |zu |), we see that the pro-
cess {fu (yu , zu ), u ≤ T } is in H2 . Then S s is well-defined by the martingale
representation theorem and S s = φ(s) ∈ H2 .
8.2. Wellposedness of BSDEs 127
0 0
2. For s, s0 ∈ H2 , denote δs := s−s0 , δS := S s −S s and δf := ft (S s )−ft (S s ).
Since δYT = 0, it follows from Itô’s formula that:
Z T Z T
αt 2 αu 2
eαu 2δYu · δfu − α|δYu |2 du
e |δYt | + e |δZu | du =
t t
Z T
−2 eαu (δZu )T δYu · dWu .
t
which converges to zero, provided that V ∈ L1 . To check that the latter condi-
tion hold true, we estimate by the Burkholder-Davis-Gundy inequality that:
!1/2
Z T
E[V ] ≤ CE e2αu |δYu |2 |δZu |2 du
0
!1/2
Z T
≤ C 0 E sup |δYu | |δZu |2 du
u≤T 0
"Z #!
T
C0
2 2
≤ E sup |δYu | +E |δZu | du < ∞.
2 u≤T 0
and then using the Lipschitz property of the generator and the Burkholder-
Davis-Gundy inequality. ♦
where δZ 0,j denotes the j−th component of δZ, and for every z 0 , z 1 ∈ Rd ,
z 1 ⊕j z 0 := z 1,1 , . . . , z 1,j , z 0,j+1 , . . . , z 0,d for 0 < j < n, z 1 ⊕0 z 0 := z 0 ,
z 1 ⊕n z 0 := z 1 .
Since f 1 is Lipschitz-continuous, the processes α and β are bounded. Solving
the linear BSDE (8.18) as in subsection 8.2.2, we get:
" Z T #
t t
δYt = E ΓT δYT + Γu δ0 fu duFt , t ≤ T,
t
where the process Γt is defined as in (8.14) with (δ0 f, α, β) substituted to (a, b, c).
Then Condition (8.17) implies that δY ≥ 0, P−a.s. ♦
Our next result compares the difference in absolute value between the solu-
tions of the two BSDEs, and provides a bound which depends on the difference
between the corresponding final datum and the genrators. In particular, this
bound provide a transparent information about the nature of conditions needed
to pass to limits with BSDEs.
Theorem 8.6. Let (Y i , Z i ) be the solution of BSDE(f i , ξ i ) for some pair
(ξ i , f i ) satisfying the conditions of Theorem 8.3, i = 0, 1. Then:
kY 1 − Y 0 k2S 2 + kZ 1 − Z 0 k2H2 ≤ C kξ 1 − ξ 0 k2L2 + k(f 1 − f 0 )(Y 0 , Z 0 )k2H2 ,
By the same argument as in the proof of Theorem 8.3, we see that the stochastic
integral term has zero expectation. Then
" Z T #
βt 2 βT 2 βu 2 2
e |δYt | = Et e |δξ| + e 2δYu · δfu − |δZu | − β|δYu | du , (8.19)
t
We then choose ε2 := C 2 /6 and β := ε−2 + 1/2, and plug the latter estimate in
(8.19). This provides:
"Z # " #
T
C 2 T βu
Z
βt 2 2 βT 2 1 0 2
e |δYt | + Et |δZu | du ≤ Et e |δξ| + e |δfu (Y , Zu )| du ,
t 2 0
which implies the required inequality by taking the supremum over t ∈ [0, T ] and
using the Doob’s maximal inequality for the martingale {Et [eβT |δξ|2 ], t ≤ T }.
♦
where
dPν
RT RT RT
1 2
:= e 0 λt (νt )·dWt − 2 0 |λt (νt )| dt ν
and βt,T := e− t
ku (νu )du
.
dP FT
We assume that all coefficients involved in the above expression satisfy the
required conditions for the problem to be well-defined.
We first notice that for every ν ∈ U, defining
" Z T #
ν Pν ν ν ν
Yt := E βt,T ξ + βu,T `u (νu )duFt
t
satisfy the conditions of Theorem 8.3, and let (Y, Z) be the corresponding solu-
tion. Then U0 = Y0 .
132 CHAPTER 8. BSDE AND STOCHASTIC CONTROL
Proof. The first claim is obvious, and the second one follows from the fact that
s,X t,x
Xrt,x = Xr s . ♦
Proposition 8.10. Let u be the function defined in Proposition 8.9, and assume
that u ∈ C 1,2 ([0, T ), Rd ). Alors
−∂t u − Au − f (., u, σ T Du) = 0 on [0, T ) × Rd .
Proof. This an easy application of Itô’s lemma together with the usual localiza-
tion technique. ♦
135