Poly Ensae
Poly Ensae
Poly Ensae
DERIVATIVES PRICING
January 2017
Contents
1 Introduction 4
1.1 European and American options . . . . . . . . . . . . . . . . . 5
1.2 No dominance principle and first properties . . . . . . . . . . . 7
1.3 Put-Call Parity . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Bounds on call prices and early exercise of American calls . . . 10
1.5 Risk effect on options prices . . . . . . . . . . . . . . . . . . . 11
1.6 Some popular examples of contingent claims . . . . . . . . . . 13
1
3.5 On discrete-time martingales . . . . . . . . . . . . . . . . . . . 37
3.6 American contingent claims in complete financial markets . . . 42
3.6.1 Problem formulation . . . . . . . . . . . . . . . . . . . 42
3.6.2 Decomposition of supermartingales . . . . . . . . . . . 43
3.6.3 The Snell envelope . . . . . . . . . . . . . . . . . . . . 44
3.6.4 Valuation of American contingent claims in a complete
market . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2
6.2 Itô’s formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.2.1 Extension to Itô processes . . . . . . . . . . . . . . . . 95
6.3 The Feynman-Kac representation formula . . . . . . . . . . . 98
6.4 The Cameron-Martin change of measure . . . . . . . . . . . . 100
6.5 Complement: density of simple processes in H2 . . . . . . . . . 102
3
Chapter 1
Introduction
4
Cameron-Martin Theorem, connection with the heat equation. We then con-
sider the Black-Scholes continuous-time financial market and derive various
versions of the famous Black-Scholes formula. A discussion of the practical
use of these formulae is provided.
i.e. if the time T price of the asset S i is larger than the strike K, then
the buyer receives the payoff STi − K which corresponds to the benefit from
buying the asset from the seller of the contract rather than on the financial
market. If the time T price of the asset S i is smaller than the strike K, the
contract is worthless for the buyer.
A European put option on the asset S i is a contract where the seller
promises to purchase the risky asset S i at the maturity T for some given
exercise price, or strike, K > 0. At time T , the buyer has the possibility,
and not the obligation, to exercise the option, i.e. to sell the risky asset to
the seller at strike K. Of course, the buyer would exercise the option only if
the price which prevails at time T is smaller than K. Therefore, the gain of
5
the buyer out of this contract is
i.e. if the time T price of the asset S i is smaller than the strike K, then
the buyer receives the payoff K − STi which corresponds to the benefit from
selling the asset to the seller of the contract rather than on the financial
market. If the time T price of the asset S i is larger than the strike K, the
contract is worthless for the buyer, as he can sell the risky asset for a larger
price on the financial market.
An American call (resp. put) option with maturity T and strike K > 0
differs from the corresponding European contract in that it offers the pos-
sibility to be exercised at any time before maturity (and not only at the
maturity).
The seller of a derivative security requires a compensation for the risk
that he is bearing. In other words, the buyer must pay the price or the
premium for the benefit of the contrcat. The main interest of this course is
to determine this price. In the subsequent sections of this introduction, we
introduce the no dominance principle which already allows to obtain some
model-free properties of options which hold both in discrete and continuous-
time models.
In the subsequent sections, we shall consider call and put options with
exercise price (or strike) K, maturity T , and written on a single risky asset
with price S. At every time t ≤ T , the American and the European call
option price are respectively denoted by
Similarly, the prices of the American and the Eurpoean put options are re-
spectively denoted by
6
The intrinsic value of the call and the put options are respectively:
i.e. the value received upon immediate exercing the option. An option is said
to be in-the-money (resp. out-of-the-money) if its intrinsic value is positive.
If K = St , the option is said to be at-the-money. Thus a call option is
in-the-money if St > K, while a put option is in-the-money if St < K.
Finally, a zero-coupon bond is the discount bond defined by the fixed
income 1 at the maturity T . We shall denote by Bt (T ) its price at time
t. Given the prices of zero-coupon bonds with all maturity, the price at
time t of any stream of deterministic payments F1 , . . . , Fn at the maturities
t < T1 < . . . < Tn is given by
F1 Bt (T1 ) + . . . + Fn Bt (Tn ) .
7
2 By a similar easy argument, we now show that American and European
call (resp. put) options prices are decreasing (resp. increasing) in the exercise
price, i.e. for K1 ≥ K2 :
Let us justify this for the case of American call options. If the holder of
the low exrecise price call adopts the optimal exercise strategy of the high
exercise price call, the payoff of the low exercise price call will be higher in
all states of the world. Hence, the value of the low exercise price call must
be no less than the price of the high exercise price call.
4 American/European Call and put prices are convex in K. Let us justify
this property for the case of American call options. For an arbitrary time
instant u ∈ [t, T ] and λ ∈ [0, 1], it follows from the convexity of the intrinsic
value that
c (t, St , T, K2 ) − c (t, St , T, K1 )
−Bt (T ) ≤ ≤ 0, K2 > K1 .
K2 − K1
The right hand-side inequality follows from the decreasing nature of the Eu-
ropean call option c in K. To see that the left hand-side inequality holds,
consider the portfolio X consisting of a short position of the European call
8
with exercise price K1 , a long position of the European call with exercise
price K2 , and a long position of K2 − K1 zero-coupon bonds. The value of
this portfolio at the maturity T is
This is a direct consequence of the fact that all stopping strategies of the
shorter maturity option are allowed for the longer maturity one. Notice that
this argument is specific to the American case.
XT = (K − ST )+ + ST − (ST − K)+ − K = 0 .
9
Finally, if the underlying asset pays out some dividends then, the above
argument breaks down because one should account for the dividends received
by holding the underlying asset S. If we assume that the dividends are known
in advance, i.e. non-random, then it is an easy exercise to adapt the put-call
parity to this context. However, id the dividends are subject to uncertainty
as in real life, there is no direct way to adapt the put-call parity.
c(St , τ, K) ≤ C(St , τ, K) ≤ St
10
option is always worth more than its exercise value, so early exercise is never
optimal.
3 Assume that the security price takes values as close as possible to zero.
Then, early exercise of American put options may be optimal before maturity.
Suppose the security price at some time u falls so deeply that Su <
K − KBu (T ).
- Observe that the maximum value that the American put can deliver when
if exercised at maturity is K.
- The immediate exercise value at time u is K − Su > K − [K − KBu (T )] =
KBu (T ) ≡ the discounted value of the maximum amount that the put could
pay if held to maturity,
Hence, in this case waiting until maturity to exercise is never optimal.
11
Definition Let Rti (T ), i = 1, 2 be the return of two securities. We say that
security 2 is more risky than security 1 if
Let C i (t, St , T, K) be the price of the American call option with payoff
+
(St Ri − K) , and C n (t, St , T, K) be the price of the basket option defined
+ +
by the payoff n1 ni=1 St Ri − K = ST + n1 ni=1 εi − K .
P P
We have previously seen that the portfolio of options with common ma-
turity and strike is worth more than the corresponding basket option:
n
1 1X i
C (t, St , T, K) = C (t, St , T, K) ≥ C n (t, St , T, K).
n i=1
Observe that the final payoff of the basket option C n (T, ST , T, K) −→ (ST − K)+
a.s. as n → ∞ by the law of large numbers. Then assuming that the pricing
functional is continuous, it follows that C n (t, St , T, K) −→ C(t, St , T, K),
and therefore: that
C 1 (t, St , T, K) ≥ C(t, St , T, K) .
Notice that the result holds due to the convexity of European/American call/put
options payoffs.
12
1.6 Some popular examples of contingent claims
Example 1.1 (Basket call and put options) Given a subset I of indices in
{1, . . . , n} and a family of positive coefficients (ai )i∈I , the payoff of a Basket
call (resp. put) option is defined by
!+ !+
X X
B = ai STi − K resp. K − ai STi .
i∈I i∈I
Example 1.3 (Asian option)An Asian call option on the asset S i with ma-
turity T > 0 and strike K > 0 is defined by the payoff at maturity:
i +
ST − K ,
i
where S T is the average price process on the time period [0, T ]. With this
definition, there is still choice for the type of Asian option in hand. One can
i
define S T to be the arithmetic mean over of given finite set of dates (outlined
in the contract), or the continuous arithmetic mean...
Example 1.4 (Barrier call options) Let B, K > 0 be two given parameters,
and T > 0 given maturity. There are four types of barrier call options on the
asset S i with stike K, barrier B and maturity T :
• When B > S0 :
13
– an Up and Out Call option is defined by the payoff at the maturity
T:
UOCT + UICT = CT
• When B < S0 :
– an Down and Out Call option is defined by the payoff at the ma-
turity T :
14
The payoff is that of a European call option if the price process of
the underlying asset crosses the barrier B before maturity. Oth-
erwise it is zero. Clearly,
DOCT + DICT = CT
Example 1.5 (Barrier put options) Replace calls by puts in the previous
example
15
Chapter 2
16
Sometimes we will also use the notation Xtθ to emphasize that the portfolio
is constructed using the strategy θ.
The gain of a portfolio strategy is defined by
d
X
G = X1 − X0 = θi ∆S i , with ∆S i = S1i − S0i .
i=0
It is convenient to express the prices of all assets in the units of the risk-free
asset, or, in other words, choose the bank account as numéraire. We therefore
introduce discounted prices
Si
S̃ti := 0t
St
and the discounted portfolio value
d
Xt X
X̃t = 0
= θ0 + θi S̃ti
St i=1
The discoutned gain of the portfolio strategy does not depend on the number
of units of the risk-free asset in the portfolio:
d
X
G̃ = X̃1 − X̃0 = θi ∆S̃ i .
i=1
Exercise 2.1 Prove that the two definitions of arbitrage opportunity are
equivalent.
Under the risk-neutral probability, the price of an asset depends only on its
expected return, but not on its risk.
18
Theorem 2.1 A financial market admits no arbitrage opportunity if and
only if there exists at least one risk-neutral probability.
Further, let C be the d-dimensional set of all expectations of (∆S̃ i )di=1 under
probabilities in Q:
C = {E Q [∆S̃]|Q ∈ Q}.
C is clearly a convex set, and to prove the existence of a risk-neutral proba-
bility it is sufficient to show that 0 ∈ C. Suppose on the contrary that 0 ∈/ C.
d
Then by the separating hyperplane theorem [23] there exists H ∈ R such
that H T x ≥ 0 for all x ∈ C and Hx0 > 0 for some x0 ∈ C. Let θ = (H0 , H)
with H0 arbitrary. For this strategy therefore E Q [G̃θ ] ≥ 0 for all Q ∈ Q and
E Q [G̃θ ] > 0 for some Q ∈ Q.
Suppose that G̃θ (ωi ) < 0 for some ωi ∈ Ω. Taking Qε = Kε + (1 − ε)1ωi we
see that E Qε [G̃θ ] < 0 for ε sufficiently small, which is a contradiction with
the above. Therefore, G̃θ ≥ 0 in all states of nature, and G̃θ > 0 in at least
one state: θ is an arbitrage opportunity. Since we have supposed that the
market is arbitrage free, this is a contradiction, and therefore we conclude
that 0 ∈ C and there exists at least one risk-neutral probability. ♦
Example 2.7 Let us look for a risk-neutral probability in the market of ex-
19
ample 2.6. We need to solve the system of equations
3 1 1 3
− q1 − q2 + q3 + q4 = 0
5 5 5 5
3 3 1 1
− q1 − q2 + q3 + q4 = 0
10 10 10 2
q1 + q2 + q3 + q4 = 1.
means that the enlarged market is not arbitrage-free. To find this opportunity
we need to construct a portfolio which only pays off in states ω2 and ω4 . For
example, the strategy consisting in the purchase of a stock funded from a bank
account and simultaneous sale of a put option is an arbitrage opportunity.
20
A contingent claim B is said to be attainable if there exists a strategy θ
such that B = X1θ . By the absence of arbitrage, the price of the contingent
claim at time 0 must then be given by p(B) = X0θ . Indeed, if the contingent
claim is quoted at a price p′ < p(B), then a strategy which consists in (i)
buying the contingent claim at t = 0; (ii) selling the portfolio X θ , has a zero
terminal pay-off and generates an immediate profit of p(B) − p′ . Similarly, if
the quoted price is greter than p(B), an arbitrage strategy can be constructed.
Since, under a risk-neutral probability Q, X0 = E Q [Ṽ1 ], we have the
following result:
Proposition 2.1 The fair price p(B) of an attainable contingent claim sat-
isfies
Q Q B
p(B) = E [B̃] = E ,
S10
where Q is any risk-neutral probability in the market.
Example 2.8 In the market of example 2.6, a put option with strike K = 1
is attainable (using the put-call parity) and one can check that its expected
1
discounted pay-off is equal to 10 under all the risk-neutral probabilities in this
market. The put option with strike K = 23 is not attainable.
Let A be the pay-off matrix: Aik = S1i (ωk ). The contingent claim B is
attainable if the equation θT A = B has a solution θ ∈ Rd+1 . Every contingent
claim is attainable if θT A = B has a solution for every B ∈ RK , which is
the case if A has K linearly independent lines. This means that for a market
to be complete, the number of assets (including the risk-free asset) must be
greater or equal to the number of states of nature.
Example 2.9 If the market of example 2.6 is enlarged with a put option with
strike K = 32 , it becomes complete (four linearly independent assets). If it
is enlarged with a put option with strike K = 1, it remains incomplete (four
assets but not linearly independent).
21
Theorem 2.2 An arbitrage-free market is complete if and only if the risk-
neutral probability is unique.
Proof. Suppose that the market is complete. Then the prices of the con-
tingent claims of the form 1ωk for k = 1, . . . , K are uniquely defined and for
every risk-neutral probability Q we have
For example, in the case of p+ , we can always restrict the set of attainable
claims over which we minimize to those satisfying C(ωi )Q(ωi ) ≤ maxi B(ωi )
(where Q is any risk-neutral measure). Since the latter set is compact, the
inf is attained.
If the claim B is attainable, p+ (B) = p− (B) but for genuinely inattainable
claims p− (B) < p(B) < p+ (B) because equality would create arbitrage.
22
Theorem 2.3 In an arbitrage-free market, the upper and lower superhedging
prices are given by
Proof. Assume first that p+ (B) < supQ E Q [B̃]. Then there exists Q ∈ QRN
with p+ (B) < E Q [B̃] and an attainable claim C with p(C) < E Q [B̃] and C ≥
B. Then E Q [C̃] < E Q [B̃] which is in contradiction with C ≥ B. Therefore
p+ (B) ≥ supQ E Q [B̃]. Similarly, we can show that p− (B) ≤ inf Q E Q [B̃].
Assume now that p− (B) < inf Q E Q [B̃] and consider an extended market
where B is quoted at price p(B) with p− (B) < p(B) < inf Q E Q [B̃]. By
theorem 2.1, in this market there exists an arbitrage opportunity, which will
be denoted by (θ, θ′ ), where θ′ 6= 0 is the coefficient in front of B. If θ′ > 0
then B dominates the attainable portfolio − θ1′ θS, whose price is equal to
p(B) (by definition of the arbitrage opportunity) and hence greater than
p− (B), which gives a contradiction.
If θ′ < 0 then B is dominated by the attainable portfolio − θ1′ θS whose
price is equal to p(B). Therefore, p+ (B) < inf Q E Q [B̃], which gives a contra-
diction with the first part of the proof.
Therefore, we conclude that p− (B) ≥ inf Q E Q [B̃] and in the same way we
can show that p+ (B) ≤ supQ E Q [B̃]. ♦
Example 2.10 Let us compute the arbitrage bounds for a put option with
strike K = 32 in example 2.6. The set {E Q [B̃]|Q ∈ QRN } coincides with
{ 15 + 52 q|0 < q < 38 }, and the bounds are therefore 51 < p(B) < 20
7
.
23
2.5 Portfolio optimization and equilibrium as-
set pricing
In this section, we consider the problem of an economic agent who disposes of
w euros at date t = 0 and wants to invest them optimally using the available
assets. Whereas the price of a contingent claim in a complete arbitrage-free
market is uniquely determined, the optimal investment policy depends on
the preferences of the agent: some agens will choose riskier portfolios than
others.
To measure the preferences of an agent, we introduce the utility function:
a function u : R → R or u : [0, ∞) → R, which is strictly increasing and
strictly concave. The agent prefers holding portfolio X at date t = 1 to
holding the portfolio Y if and only if E[u(X)] ≥ E[u(Y )]. The fact that
u is increasing means that X is always preferred to Y if X ≥ Y , and the
concavity implies (via Jensen’s inequality) that E[X] is always preferred to
X. Since increasing linear transformations of the utility function do not
change the structure of preferences, the utility function is only defined up to
such transformations.
For a random financial flow X, the certainty equivalent of X is the
constant cX ∈ R such that u(cX ) = E[u(X)]. The risk premium is the
difference between the expected value of X and its certainty equivalent:
ρ(X) = E[X]−cX . Let m = E[X]. Then u(cX ) = u(m−ρ) ≈ u(m)−u′ (m)ρ.
On the other hand, u(cX ) = E[u(X)] ≈ u(m) + 21 u′′ (m)Var(X). Therefore,
Var(X)
ρ ≈ α(m) ,
2
′′
where α(x) = − uu′ (x)
(x)
is called the absolute risk aversion coefficient at the
level x.
24
Setting α = 1−γ x
for γ ∈ [0, 1) leads to hyperbolic absolute risk aversion
(HARA) utility function, defined for x ∈ [0, ∞), which models the situation
where the agent’s risk aversion decays with increasing wealth. We get:
1−γ
(− log u′ )′ = ⇒ u′ = Cxγ−1 .
x
If γ = 0, this leads to the logarithmic utility u(x) = log x, and the case γ > 0
γ
corresponds to power utility u(x) = xγ .
subject to the constraint X0θ = w, called the budget constraint. First, let us
observe that if the market is not arbitrage-free, this problem does not have
a (finite) solution. Indeed, let θ̂ be an optimal strategy with E[u(X1θ̂ )] < ∞
and θA be an arbitrage opportunity. Then the strategy θ̂ + θA satisfies the
budget constraint and
which means that θ̂ cannot be optimal. Therefore, from now on, we suppose
that the market is arbitrage-free. This means that there exits a risk-neutral
probability Q such that
X1θ
Q
E =w (2.2)
(1 + R)
The budget constraint can therefore be replaced by the constraint (2.2), and
the utility maximization problem reduces to two independent problems:
• Maximize the utility over the set of all hedgeable claims subject to the
bugdet constraint:
25
• Find a trading strategy θ which replicates X̄. This is a problem that
we already know how to solve from previous sections.
At this point, two remarks are in order. First, if the market is complete, all
claims are hedgeable, and the first problem reduces to
Secondly, exactly the same logic can be used to solve the portfolio optimiza-
tion problem in a multiperiod discrete-time model.
In view of the above discussion, from now on we concentrate on the prob-
lem (2.3), but without necessarily supposing that the market is complete. If
it is, this problem gives us an immediate solution to the portfolio optimiza-
tion problem. In the incomplete market case, it will allow us to understand
the optimal contingent claim profiles and explain the demand for derivative
products in financial markets. The problem of finding an optimal portfolio in
an incomplete market is more difficult and out of scope of this introductory
course.
26
risk-neutral probability. The problem (2.3) becomes
K
X K
X
max pi u(xi ) subject to qi xi = w(1 + R).
x∈RK
i=1 i=1
pi u′ (xi ) = λqi .
The function
K
X qi
f (λ) = qi I λ
i=1
pi
is strictly decreasing and satisfies
Therefore, under (A1), for every initial wealth w ∈ R there exists a solution
λ∗ > 0 of equation (2.4), and under (A2), such a solution exists for every
positive initial wealth w > 0. The optimal claim is then given by
∗ ∗ qi dQ
xi = I λ or X ∗ = I(λ∗ Z) with Z := .
pi dP
27
To show that X ∗ is indeed the unique optimizer of the constrained problem
(2.3), assume that X ′ = (x′1 , . . . , x′k ) is another solution, different from X ∗ .
We then have:
K
X K
X
pi u(x′i ) = L(x′1 , . . . , x′K , λ∗ ) < L(x∗1 , . . . , x∗K , λ∗ ) = pi u(x∗i ),
i=1 i=1
where the first equality holds because X ′ satisfies the budget equation, the
inequality is due to the strict concavity of L and the last equality follows
from the budget equation for X ∗ .
which means that the agents are rewarded for the risk they take.
28
We suppose that the market consists of a finite set A of economic agents with
utility functions (ua )a∈A and initial endowments with pay-offs (Wa )a∈A . The
P
sum of all endowments M := a∈A Wa is the market portfolio, that is, the
aggregate portfolio of all assets available for trading. A feasible allocation
(Xa )a∈A is a set of contingent claims satisfying the market clearing condition:
X
Xa = M.
a∈A
The prices in the market will be determined via a pricing density φ: a random
variable φ > 0 satisfying E[φ] = 1. The price of a contingent claim X is
φX
p(X) = E 1+R .
A microeconomic or Arrow-Debreu equilibrium is a pricing density φ∗ and
a feasible allocation (Xa∗ )a∈A , such that for every a ∈ A, Xa solves the utility
maximization problem for agent a subject to the budget constraint defined
by the pricing rule φ∗ : E[φ∗ Xa ] = E[φ∗ Wa ].
In these introductory lecture notes we consider the simplest case where
all agents have an exponential utility function: ua (x) = 1 − e−αa x . Let φ be
a pricing density. Then the optimal allocation for agent a has the form
1
Xa = E[φWa ] + {E[φ log φ] − log φ} ,
αa
and the market clearing condition becomes
1
M = E[φM] + {E[φ log φ] − log φ}
α
where we set α1 := a∈A 1
P
αa
. This implies that equilibrium pricing rule is
necessarily of the form
e−αM
φ∗ = .
E[e−αM ]
with this pricing rule, the optimal allocation for each agent
α
Xa∗ = E[φ∗ Wa ] + (M − E[φ∗ M])
αa
evidently satisfies the market clearing condition. The couple (φ∗ , X ∗ ) is then
the unique equilibrium in this market. In equilibrium, the agents hold linear
29
shares of the market portfolio, which means that once again we recover the
mutual fund theorem.
30
Chapter 3
31
Trading strategies The trading strategy is a d + 1-dimensional stochastic
i=0,...,d
process (θti )t=1,...,T , where θti represents the amount of i-th asset held in the
portfolio between dates ti−1 and ti . The trading strategies are also assumed
to be predictable: the portfolio composition for the period is known in the
beginning of the period.
For example, a constant strategy θti = θ0i corresponds to buying the assets
at t = 0 and holding them until t = T . In the fixed-mix strategy, one fixes the
proportions ω0 , . . . , ωd of the total wealth to invest into different assets, and
at each date the portfolio is readjusted to maintain the same proportions.
Since the asset prices do not vary proportionnally, the amounts of assets will
be different at each date for this strategy.
The gain process of a strategy is the stochastic process describing the
change in the portfolio value due to the variation of the asset prices:
t X
X d
G0 = 0, Gt = θsi ∆Ssi , ∆Sti = Sti − St−1
i
.
s=1 i=0
32
For a self-financing strategy these two values are equal and the value of
portfolio at date t is defined uniquely by
d
X
Xt = θti Sti = X0 + Gt .
i=0
33
Proof. Let Q be a martingale probability. Then the discounted value (X̃t )
of every self-financing portfolio is a Q-martingale, so that for every strategy
θ with X0θ = 0 we have E Q [X̃Tθ ] = 0, which is in contradiction with VTθ ≥ 0
in all states and VTθ > 0 in at least one state.
Conversely, assume that the market is aribtrage-free. We consider a one-
period market with the same states of nature Ω as the original one, with
one risk-free asset evolving with zero interest rate, and with the risky assets
having pay-offs of the form
and quoted all at zero price at the initial date. Since Ω is finite, there is only
a finite number of such risky assets. This number will be denoted by D. For
i(j) i(j)
j = 1, . . . , D, we denote the pay-off of the j-th asset by 1A(j) (S̃t(j) − S̃t(j)−1 )
We claim that the new market is also arbitrage-free. Indeed, assume that
w ∈ RD+1 is an arbitrage strategy in this market. The discounted gain G
associated to this strategy can be written as
D T X
d X
D
i(j) i(j)
X X
G= wj 1A(j) (S̃t(j) − S̃t(j)−1 ) = i
wj 1i(j)=i 1t(j)=s 1A(j) (S̃si − S̃s−1 )
j=1 s=1 i=1 j=1
T X
X d
= i
θsi (S̃si − S̃s−1 ),
s=1 i=1
where
D
X
θsi = wj 1i(j)=i 1t(j)=s 1A(j) ∈ Fs−1 .
j=1
34
for all At ∈ Ft−1 , which implies that
35
3.4 Complete markets
As in the one-period model, a market is said to be complete if every contin-
gent claim is attainable. Moreover, the same characterization in terms of the
uniqueness of the martingale probability holds.
Proof. The only if part. Assume that the market is complete. This means
that every contingent claim is attainable and has a unique price, in particular
the claims ST0 1ωi for every i, which shows the uniqueness of the martingale
probability.
The if part. By way of contradiction, assume that there exists a non-
attainable contingent claim. We consider a one-period market with the same
states of nature Ω as the original one, with the one risk-free asset evolving
with zero interest rate, and with the risky assets having pay-offs of the form
and quoted all at zero price at the initial date. Since Ω is finite, there is only
a finite number of such risky assets. In this one-period market there is also
a non-attainable claim, and by theorem 2.2, there exists a probability Q∗ on
Ω which is different from Q and satisfies
∗
h i
E Q 1At (S̃ti − S̃t−1
i
) =0
36
3.5 On discrete-time martingales
In this section, we recall some results from the theory of discrete-time mar-
tingales.
is an (F, P)−martingale.
{ν = t} ∈ Ft for every t = 0, . . . , T .
37
Remark 3.1 An equivalent characterization of an F−stopping time is that
{ν ≤ t} ∈ Ft for every t = 0, . . . , T ,
Fν := {A ∈ F : A ∩ {ν = t} ∈ Ft for all t = 0, . . . , T } .
Proof. We only prove this result for martingales ; the case of supermartin-
gales is treated similarly.
(i) We first show that E[MT |Fν ] = Mν for every ν ∈ T . Let A be an arbitrary
event set in Fν . Then, the event set A ∩ {ν = t} ∈ Ft and therefore :
E (MT − Mν )1A∩{ν=t} = E (MT − Mt )1A∩{ν=t} = 0
38
since M is a martingale. Summing up these inequalities, it follows that :
T
X
0 = E (MT − Mν )1A∩{ν=t} = E [(MT − Mν )1A ] .
t=0
since ν ≤ ν̄. ♦
τn := inf {t : |φt | ≥ n} ,
39
The following result provides a sufficient condition in finite discrete time
for a local martingale to be a martingale.
E[MT− ] < ∞ .
Since τn is an (F, P)−stopping time, we deduce that the event set {τn > t−1}
is Ft−1 −measurable. Then :
where the last inequality follows from the fact that M τn is P−integrable. By
sending n to infinity, we deduce that E[|Mt ||Ft−1] < ∞ P -a.s.
Similarly, using the martingale property of the stopped process M τn , we
see that :
τn
E[Mt |Ft−1 ] = E[Mtτn |Ft−1 ] = Mt−1 = Mt−1 on {τn > t − 1} ,
which implies that E[Mt |Ft−1 ] = Mt−1 P -a.s. by sending n to infinity (Notice
that the condition E[MT− ] < ∞ has not been used in this step).
Mt− ≤ E[Mt+1
−
|Ft ] for every t = 0, . . . , T − 1 .
40
Then, E[Mt− ] ≤ E[Mt+1
−
] ≤ . . . ≤ E[MT− ] < ∞ by the condition of the lemma.
We next use Fatou’s lemma to obtain :
h i
+ +
E[Mt ] = E lim inf Mt∧τn
n→∞
−
+
≤ lim inf E Mt∧τ n
= lim inf E Mt∧τn + Mt∧τ n
.
n→∞ n→∞
41
Remark 3.2 In finite discrete-time, Lemma 3.1 provides an easy neces-
sary and sufficient condition for a local martingale to be a martingale (ne-
cessity is trivial). This condition does not extend to the contionuous-time
framework, as we will provide an example of a positive continuous-time (in-
tegrable) local martingale which fails to be a martingale. ♦
42
decides to exercise the option at time t.
This leads to the following concept of American contingent claims.
The seller of an American contingent claim has to face the promised payoff
Bt at any time t before the maturity T , as the buyer can decide to exercise
his right. Then, the relevant super-hedging problem in this context is :
n o
V a (B) := inf x ∈ R : Xtx,θ ≥ Bt , 1 ≤ t ≤ T a.s. for a θ ∈ A . (3.2)
Yt = Mt − At , t = 0, . . . , T ,
43
Proof. For t = 0, the only possible choice is M0 = Y0 since A0 = 0. Next,
we must have :
44
By definition, ȲT ≥ BT = YT . Suppose that Ȳt ≥ Yt . Since Ȳ is a Q-
supermartingale, we have :
Definition 3.11 The process Y is called the Snell envelope of B for the
probability measure Q and the filtration F.
Let T denote the collection of all stopping times with values in {0, . . . , T }.
Our next result provides a representation of the Snell envelope in terms of
the underlying process B and the stopping times of T .
ν ∗ := inf{t = 0, . . . , T : Yt = Bt }
∗
is a stopping time such that the stopped process Y ν is a Q−martingale and
Proof. (i) We first show that ν ∗ defines a stopping time. Since YT = BT , the
random variable ν ∗ takes values in {0, . . . , T }. Clearly {ν ∗ = 0} = {Y0 = B0 }
∈ F0 and for t ≥ 1 :
{ν ∗ = t} = ∩t−1
k=0 {Yk > Bk } ∩ {Yt = Bt } ∈ Ft
45
By definition of the process Y , we have Yt > Bt and Yt = EQ [Yt+1 |Ft ] on
the event set {ν ∗ ≥ t + 1} (Ft -measurable as the complement of {ν ∗ ≤ t}).
Hence :
ν∗ ∗
Yt+1 − Ytν = (Yt+1 − E Q [Yt+1 |Ft ])1{ν ∗ ≥t+1} ,
and the required result is obtained by taking the expected values condition-
ally on Ft and using the fact that {ν ∗ ≥ t + 1} ∈ Ft .
∗
(iii) Since the stopped process Y ν is a Q−martingale, we have :
∗
Y0 = EQ YTν = EQ [YT ∧ν ∗ ] = EQ [Yν ∗ ] = EQ [Bν ∗ ] .
M(S) = {P0 } .
The next result expresses the super-replication problem for American con-
tingent claim (3.2) as an optimal stopping problem.
46
Theorem 3.8 Let B be an American contingent claim satisfying (3.3). Then,
0 0
V a (B) = sup EP [B̃ν ] = EP [B̃ν ∗ ] ,
ν∈T
∗
where ν = inf{t = 0, . . . , T : Yt = Bt } and Y is the Snell envelope of the
process B for the probability measure P0 and the filtration F.
Proof. (i) Let Ỹ be the Snell envelope of the process B̃. This defines the
process Y by the usual relation Y := S 0 Ỹ . As a P0 -supermartingale, the
Snell envelope Ỹ has the Doob-Meyer decomposition :
Ỹt = Mt − At , t = 0, . . . , T ,
where M is a P0 −martingale and A a non-decreasing process with A0 =
0. Since the financial market is complete, the (European) contingent claim
defined by the random variable ST0 MT is attainable, i.e. there exists a self-
financing portfolio strategy θ ∈ A such that :
T −1
P0
X
MT = E [MT ] + θt · S̃t+1 − S̃t
t=0
T −1
X
= Y0 + θt · S̃t+1 − S̃t
t=0
47
for t = 1, . . . , T . By definition of V a , this shows that Y0 ≥ V a (B), and by
Proposition 3.4, we obtain the first inequality
0
h i 0
V a (B) ≤ sup E P B̃ν = E P [B̃ν ∗ ] .
ν∈T
for every ν ∈ T . We next check that V (Bν ) = V̄ (Bν , ν). For x > V̄ (Bν , ν),
there exists a self-financing portfolio strategy θ ∈ A such that Xνx,θ ≥ Bν a.s.
Consider the portfolio strategy
48
0
Corollary
h i3.1 Suppose that the process B̃ is a P -submartingale, i.e. B̃k ≤
0
EP B̃t |Fk for 0 ≤ k ≤ t ≤ T . Then :
0
V a (B) = V (BT ) = EP [B̃T ] .
Remark 3.3 The last corollary applies to the case of American call options.
Let r be a constant interest rate, to simplify, and let us check that the process
{e−rt (St − K)+ , t = 0, . . . , T } is a P0 −submartingale. For 0 ≤ k < t ≤ T ,
we have :
0 0
EP [e−rt (St − K)+ |Fk ] ≥ EP [e−rt (St − K)|Fk ]
= e−rk Sk − e−rt K ≥ e−rk (Sk − K)
0
since S̃ is a P0 −martingale and r ≥ 0. Since EP [(St − K)+ |Fk ] ≥ 0, this
provides :
0
EP [e−rt (St − K)+ |Fk ] ≥ e−rk (Sk − K)+ .
49
Chapter 4
S0 = s , S1 (ωu ) = su , S1 (ωd ) = sd ,
where s, r, u and d are given strictly positive parameters with u > d. Such
50
a financial market can be represented by the binomial tree :
time 0 time 1
Su
Risky asset S0 = s
Sd
X̃1x,θ = x + θ(S̃1 − s) .
0 0
In this context , it turns out that there exists a pair (x0 , θ0 ) such that XTx ,θ =
0 0
B. Indeed, the equality X1x ,θ = B is a system of two (linear) equations with
51
two unknowns which can be solved straightforwardly :
Bu Bd Bu − Bd
x0 = qu + (1 − qu ) = EQ [B̃] and θ0 = .
1+R 1+R su − sd
We also observe in our simple context that such a pair (x0 , θ0 ) is unique.
Then x0 = EQ [Be−r ] and (x0 , θ0 ) is a perfect replication strategy for B, i.e.
0
X1x0 ,θ = B.
(iii) No arbitrage valuation : Let us denote by p(B) the market price of the
contingent claim contract. Under the no-arbitrage condition we have that
p(B) = EQ Be−r .
(4.2)
Hence in the context of the simple binomial model, the no-arbitrage condi-
tion implies a unique valuation rule. This is the no-arbitrage price of the
contingent claim B.
In order to better understand this result, let us show directly this equality.
Suppose that the contingent claim B is available for trading at time 0. We
will now show that, under the no arbitrage condition, the price p(B) of the
contingent claim contract is necessarily given by p(B) = x0 .
(iii-a) Indeed, suppose that p(B) < x0 , and consider the following portfolio
strategy :
- at time 0, pay p(B) to buy the contingent claim, so as to receive the
payoff B at time 1,
- perform the self-financing strategy (−x0 , −θ), this leads to paying −x0
at time 0, and receiving −B at time 1
The initial capital needed to perform this portfolio strategy is p(B) − x0 < 0.
At time 1, the terminal wealth induced by the self-financing strategy exactly
compensates the payoff of the contingent claim. We have then built an
arbitrage opportunity in the financial market augmented with the contingent
claim contract, thus violating the no arbitrage condition on this enlarged
financial market.
(iii-b) If p(B) > x0 , we consider the following portfolio strategy :
52
- at time 0, receive p(B) by selling the contingent claim, so as to pay the
payoff B at time 1,
- perform the self-financing strategy (x0 , θ), this leads to paying x0 at
time 0, and receiving B at time 1
The initial capital needed to perform this portfolio strategy is −p(B) + x0 <
0. At time 1, the terminal wealth induced by the self-financing strategy
exactly compensates the payoff of the contingent claim. This again defines an
arbitrage opportunity in the financial market augmented with the contingent
claim contract, thus violating the no arbitrage condition on this enlarged
financial market.
Remark 4.4 All the results of this section hold true with a sequence (bn , σn )
satisfying :
√ √
nbn −→ bT and nσn −→ σ T whenever n → ∞ .
53
For n ≥ 1, we consider the price process of a single risky asset S n = {Skn ,
k = 0, . . . , n} defined by :
k
!
X
S0n = s and Skn = s exp kbn + σn Zi , k = 1, . . . , n .
i=1
Rn := er(T /n) − 1 .
For each n ≥ 1, we have then defined a financial market with time step T /n.
In order to ensure that these financial markets satisfy the no-arbitrage
condition, we assume that :
Exercise 4.4 Show that (4.3) is a necessary and sufficient conditon for the
financial markets of this section to satisfy the no-arbitrage condition.
Under this condition, there exists a unique risk neutral measure P0n defined
by :
1 + Rn − dn
P0n [Zi = 1] = qn := .
u n − dn
By Theorem 2.2, the above financial markets are complete.
From the results of the previous chapters, the no-arbitrage price of this Eu-
ropean call option is :
0
pn (B n ) = e−rT EPn (Snn − K)+ .
54
Under the probability measure P0n , the random variables (1 + Zi )/2 are inde-
pendent and identically distributed as a Bernoulli with parameter qn . Then :
" n #
X 1 + Zi
P0n = j = Cnj qnj (1 − qn )n−j for j = 0, . . . , n .
i=1
2
This provides
n
X
n n −rT
g sujn dnn−j Cnj qnj (1 − qn )n−j .
p (B ) = e
j=0
55
pnk (B n ) is the no-arbitrage price of the contingent claim pnk+1 (B n ) :
un Skn
Risky asset Skn
dn Skn
pnk+1 (B n )un
Contingent claim pnk (B n )
pnk+1 (B n )dn
n
pnk+1 (B n )un − pnk+1 (B n )dn
θk+1 = , k = 0, . . . , n − 1 .
un Skn − dn Skn
Remark 4.6 The reference measure P is not involved neither in the val-
uation formula, nor in the hedging formula. This is due to the fact that
the no-arbitrage price in our complete market framework coincides with the
super-hedging cost, which in turn depends on the reference measure only
through the corresponding zero-measure sets.
56
At the terminal date, the price Pnn (B n ) of the American put option coin-
cides with the price pnn (B n ) of the European put option with the same strike,
since both coincide with the pay-off. At a date k < n, the buyer has the
choice between exercising the option and receiving (K − Skn )+ or not exercis-
ing. In the latter case, the option becomes “locally European” and its price
1 n
is 1+R n
E Q [Pk+1 (B n )|Fk ] (called continuation value). The optimal exercise
policy is to choose the largest of these two values:
1
Pkn = max{(K − Skn )+ , n n
(B n )un + (1 − qn )Pk+1 (B n )dn }.
qn Pk+1
1 + Rn
Therefore, the rational buyer will exercise her option as soon as the Pkn
becomes equal to (K − Skn ). For the seller of the option, the same hedging
strategy
n
P n (B n )un − Pk+1 (B n )dn
n
θk+1 = k+1
un Skn − dn Skn
allows to hedge the put option until the exercise date. If the buyer does not
exercise optimally, after the optimal exercise date, the hedging strategy is no
longer self-financing and generates profit for the seller: the seller can with-
draw some money from the account and still be able to face her obligations.
57
Introduce the sequence :
and let
Lemma 4.3 Let (Xk,n )1≤k≤n be a triangular sequence of iid Bernoulli ram-
dom variables with parameter πn :
P[Xk,n = 1] = 1 − P[Xk,n = 0] = πn .
58
Then:
Pn
k=1 Xk,n − nπn
p −→ N (0, 1) in distribution .
nπn (1 − πn )
where
√
−rT ln(s/k) v
K̃ := Ke , d± (s, k, v) := √ ± ,
v 2
Rx 2 √
and N(x) = −∞ e−v /2 dv/ 2π is the cumulative distribution function of
standard N (0, 1) distribution.
59
1. By definition of ηn , we have
sunηn −1 dn−η
n
n +1
≤ K ≤ suηnn dnn−ηn .
Then,
r r !
T T T K
+ O n−1/2 ,
2ηn σ +n b −σ = ln
n n n s
n √ ln(K/s) − bT √
ηn = + n √ + ◦( n) . (4.6)
2 2σ T
We also compute that
σ2
1 r − b − 2
T √ √
nqn = + √ n + ◦( n) . (4.7)
2 2σ T
By (4.6) and (4.7), it follows that
ηn − nqn
lim p = −d− (s, K̃, σ 2 T ) .
n→∞ nqn (1 − qn )
where LQn (Z) denotes the distribution under Qn of the random variable Z.
Then :
"P #
n
Z j − nq n ηn − nqn
lim B(n, qn , ηn ) = lim Qn pk=1 ≥p
n→∞ n→∞ nqn (1 − qn ) nqn (1 − qn )
2 2
= 1 − N −d− (s, K̃, σ T ) = N d− (s, K̃, σ T ) .
60
Proof of Lemma 4.3. (i) We start by recalling a well-known result on
characteristic functions. Let X be a random variable with E[X n ] < ∞.
Then :
n
itX X (it)n
φX (t) := E e = E[X k ] + ◦(tn ) . (4.8)
k=0
n!
To prove this result, we denote F (t, x) := eitx and f (t) = E[F (t, X)]. The
function t 7−→ F (t, x) is differentiable with respect to the t variable. Since
|Ft (t, X)| = |iXF (t, X)| ≤ |X| ∈ L1 , it follows from the dominated conver-
gence theorem that the function f is differentiable with f ′ (t) = E[iXeitX ]. In
particular, f ′ (0) = iE[X]. Iterating this argument, we see that the function
f is n times differentiable with n−th order derivative at zero given by :
61
Chapter 5
The purpose of this and the following chapter is to introduce the concepts
from stochastic calculus which will be needed for the future development
of the course, namely the Black and Scholes model for pricing and hedging
European Vanilla options. We provide a treatment of all of these applications
without appealing to the general theory of stochastic integration.
V : R+ × Ω −→ E
(t, ω) 7−→ Vt (ω)
62
is measurable. For a fixed ω ∈ Ω, the function t ∈ R+ 7−→ Vt (ω) is the
sample path (or trajectory) of V corresponding to ω.
A filtration F = {Ft , t ≥ 0} is an increasing family of sub-σ−algebras
of F . As in the discrete-time context, Ft is intuitively understood as the
information available up to time t. The increasing feature of the filtration,
Fs ⊂ Ft for 0 ≤ s ≤ t, means that information can only increase as time
goes on.
63
Since the map V n is obviously B([0, t]) ⊗ F −measurable, we deduce that the
measurability of the limit map V defined on [0, t] × Ω. ♦
A random time is a random variable τ with values in [0, ∞]. It is called
- a stopping time if the event set {τ ≤ t} is in Ft for every t ∈ R+ ,
- an optional time if the event set {τ < t} is in Ft for every t ∈ R+ .
Obviously, any stopping time is an optional time. It is an easy exercise
to show that these two notions are in fact identical whenever the filtration F
is right-continuous, i.e.
This will be the case in all of the financial applications of this course. An
important example of a stopping time is:
Exercise 5.6 (first exit time) Let X be a stochastic process with continu-
ous paths adapted to F, and consider a closed subset Γ ∈ B(E). Show that
the random time
TΓ := inf {t ≥ 0 : Xt 6∈ Γ}
Fτ := {A ∈ F : A ∩ {τ ≤ t} ∈ Ft for every t ∈ R+ }
Proof. For stopping times τ1 and τ2 taking values in a finite set, the proof
is identical to that of Theorem 3.6. In order to extend the result to gen-
eral stopping times, we approximate the stopping times τi by the decreasing
sequence τin := ⌊nτin⌋+1 , and we use a limiting argument. ♦
65
with supt≥0 E Vt+ < ∞. This is the so-called submartingale convergence
theorem, see e.g. Karatzas and Shreve [16] Theorem 1.3.15. In the con-
text of these lectures, we shall simply apply the following consequence of the
optional sampling theorem:
66
many surprising properties. Kyioshi Itô (1915-) developed the stochastic
differential calculus. The theory benefitted from the considerable activity on
martingales theory, in particular in France around P.A. Meyer.
(i) W is F−adapted,
(ii) W0 = 0 and the sample paths W. (ω) are continuous for a.e. ω ∈ Ω,
(i) W0 = 0 and the sample paths W. (ω) are continuous for a.e. ω ∈ Ω,
(ii) W has independent increments: for all 0 ≤ t1 < t2 ≤ t3 < t4 , Wt4 −Wt3
and Wt2 − Wt1 are independent.
where ⌊t⌋ denotes the largest integer less than or equal to t. The following
figure shows a typical sample path of the process M.
68
√
In the above definition, the normalization by n is suggested by the Central
Limit Theorem. We next set
k
tk := for k ∈ N
n
and we list some obvious properties of the process W n :
• for 0 ≤ i ≤ j ≤ k ≤ ℓ ≤ n, the increments Wtnℓ − Wtnk and Wtnj − Wtni
are independent,
• for 0 ≤ i ≤ k, the two first moments of the increment Mtk − Mti are
given by
Hence, except the Gaussian feature of the increments, the discrete-time pro-
cess Wtnk , k ∈ N is approximately a Brownian motion. One could even
obtain Gaussian increments with the required mean and variance by replac-
ing the distribution (5.1) by a convenient normal distribution. However, since
our objective is to imitate the Brownian motion in the asymptotics n → ∞,
the Gaussian distribution of the increments is expected to hold in the limit
by a central limit type of argument.
The following figures represent a typical sample path of the process W n .
Another interesting property of the rescaled random walk, which will be
inherited by the Brownian motion, is the following quadratic variation result:
k
X 2
n n
[W , W ]tk = Wtnj − Wtnj−1 = tk for k ∈ N .
j=1
69
Figure 5.2: The rescaled random walk
70
Figure 5.3: Approximation of a sample path of a Brownian motion
71
Figure 5.4: A sample path of the two-dimensional Brownian motion
motion W 2 is a submartingale:
72
consequence of the independence of the increments of the Brownian motion.
We shall see later in Corollary 5.2 that the Markov property holds in a
stronger sense by replacing the deterministic time t by an arbitrary stopping
time τ .
• The Brownian motion is a centered Gaussian process as it follows from
its definition that the vector random variable (Wt1 , . . . , Wtn ) is Gaussian for
every 0 ≤ t1 < . . . < tn . Centered Gaussian processes can be character-
ized in terms of their covariance function. A direct calculation provides the
covariance function of the Brownian motion
Exercise 5.9 For 0 ≤ t1 < t̂ < t2 , show that the conditional distribution
of Wt̂ given (Wt1 , Wt2 ) = (x1 , x2 ) is Gaussian, and provided its mean and
variance in closed form. ♦
1 (y−x)2
p(t, x, T, y)dy := P [WT ∈ [y, y + dy]|Wt = x] = p e− 2(T −t) dy
2π(T − t)
An important observation is that this density function satisfies the heat equa-
tion for every fixed (t, x):
∂p 1 ∂2p
= ,
∂T 2 ∂y 2
73
as it can be checked by direct calculation. One can also fix (T, y) and express
the heat equation in terms of the variables (t, x):
∂p 1 ∂ 2 p
+ = 0. (5.2)
∂t 2 ∂x2
We next consider a function g with polynomial growth, say, and we define
the conditional expectation:
Z
V (t, x) = E [g (WT ) |Wt = x] = g(y)p(t, x, T, y)dy . (5.3)
74
Proof. Properties (i), (ii) and (iii) of Definition 5.15 are immediately
checked. ♦
Bt := Wt+τ − Wτ , t ≥ 0,
This would imply that B has independent increments with the required Gaus-
sian distribution.
Observe that we may restrict our attention to the case where τ has a finite
support {s1 , . . . , sn }. Indeed, given that (5.5) holds for such stopping times,
75
one may approximate a general stopping time τ by the decreasing sequence
of stopping times τ n := (⌊nτ ⌋ + 1) /n, apply (5.5) for each n ≥ 1, and pass
to the limit by the monotone convergence theorem thus proving that (5.5)
holds for τ .
For a stopping time τ with finite support {s1 , . . . , sn }, we have:
E φ (Bt4 − Bt3 ) ψ (Bt2 − Bt1 ) f (Ws ) 1{s≤τ }
n
X
= E φ (Bt4 − Bt3 ) ψ (Bt2 − Bt1 ) f (Ws ) 1{s≤τ } 1{τ =si }
i=1
n
X h i
= E φ Wti4 − Wti3 ψ Wti2 − Wti1 f (Ws ) 1{τ =si ≥s}
i=1
i=1
n
X
= E E [φ (Wt4 − Wt3 )] E [ψ (Wt2 − Wt1 )] f (Ws ) 1{τ =si ≥s}
i=1
where the last equality follows from the independence of the increments of
the Brownian motion and the symmetry if the Gaussian distribution. Hence
E φ (Bt4 − Bt3 ) ψ (Bt2 − Bt1 ) f (Ws ) 1{s≤τ }
Xn
= E [φ (Wt4 − Wt3 )] E [ψ (Wt2 − Wt1 )] E f (Ws ) 1{τ =si ≥s}
i=1
76
Corollary 5.2 The Brownian motion satisfies the strong Markov property:
Wt∗ := sup Ws , t ≥ 0.
0≤s≤t
The key-idea for this result is to make use of the Brownian motion started
at the first hitting time of some level y:
Observe that
{Wt∗ ≥ y} = {Ty ≤ t} ,
P [Wt∗ ≥ y] = P [|Wt | ≥ y] .
77
Furthermore, the joint distribution of the Brownian motion and the corre-
sponding running maximum is characterized by
Proof. From Exercise 5.6 and Proposition 5.7, the first hitting time Ty of
the level y is a stopping time, and the process
Bt := Wt+Ty − WTy , t ≥ 0,
is a Brownian motion independent of FTy . Since Bt and −Bt have the same
distribution and WTy = y, we compute that
where the last equality follows from the fact that {Wt∗ ≥ y} ⊂ {Wt ≥ 2y −x}
as x ≥ y. As for the marginal distribution of the running maximum, we
decompose:
where the two last equalities follow from the first part of this proof together
with the symmetry of the Gaussian distribution. ♦
78
5.6 On the filtration of the Brownian motion
Because the Brownian motion has a.s. continuous sample paths, the corre-
sponding canonical filtration is left-continuous, i.e. ∪s<t FsW = FtW . How-
ever, FW is not right-continuous. To see this, observe that the event set
W
{W has a local maximum at t} is in Ft+ := ∩s>t FsW , but is not in FtW .
This difficulty can be overcome by slightly enlarging the canonical filtra-
tion by the collection of zero-measure sets. We define N := {A ∈ Ω : P(A) =
0} and the filtration by F̄tW := σ(FtW ∪ N ).
The resulting filtration F̄W is called the completed canonical filtration,
and W remains a Brownian motion with respect to F̄W , because the four
properties of the Definition 5.14 do not depend on the null sets of P. Impor-
tantly, F̄W can be shown to be continuous. We send the interested reader to
Karatzas and Shreve [16] for a rigorous treatment of this question. In the
rest of these notes, we always work with the completed filtration, and since
there is no ambiguity, we shall from now on omit the bar in F̄W , that is, FW
will denote the completed filtration.
The following result appears simple, but has far-reaching consequences
for the study of local behavior of Brownian paths.
79
Theorem 5.11 For a Brownian motion W , we have
Wt
−→ 0 P − a.s. as t → ∞ .
t
Proof. We first decompose
Wt Wt − W⌊t⌋ ⌊t⌋ W⌊t⌋
= +
t t t ⌊t⌋
By the law of large numbers, we have
⌊t⌋
W⌊t⌋ 1 X
= (Wi − Wi−1 ) −→ 0 P − a.s.
⌊t⌋ ⌊t⌋ i=1
is a Brownian motion.
80
Proof. All of the properties (i)-(ii)-(iii) of Definition 5.15 are obvious
except the sample path continuity at zero. But this is equivalent to the law
of large numbers stated in Proposition 5.11. ♦
The following two results show the complexity of the sample paths of the
Brownian motion.
almost surely.
Observe that the event “W changes sign infinitely many times in any time
interval [0, t]” is in F0B , and therefore its probability is either zero or one.
Assume that it is zero. Then, with probability one, either lim inf tց0 Wt t = 0
or lim suptց0 Wt t = 0, which is a contradiction with Proposition 5.11.
The following result shows that the sample paths of the Brownian motion
are unbounded a.s., and the Brownian motion is recurrent, meaning that this
process returns to the point 0 an infinite number of times.
81
Proposition 5.13 For a standard Brownian motion W , we have
almost surely.
Proof.
From Proposition 5.10, Bt := tW1/t defines a Brownian motion. Since
Wt /t = B1/t , it follows that the behavior of Wt /t for t ց 0 corresponds to
the behavior of Bu for u ր ∞. The required limit result is then a restatement
of Proposition 5.11 for t0 = 0. ♦
We conclude this section by stating, without proof, the law of the iterated
logarithm for the Brownian motion. The interested reader may consult [16]
for the proof.
As we shall see shortly, the Brownian motion has infinite total variation.
In particular, this implies that classical integration theories are not suitable
for the case of the Brownian motion. The key-idea in order to define an
integration theory with respect to the Brownian motion is that the quadratic
variation is finite. In the context of these notes, we shall not present the
82
complete stochastic integration theory. We will restrict our attention to the
quadratic variation along dyadic sequences defined by
X 2
Vtn := W − W for t > 0 .
n n
ti+1 ti
tn
i ≤t
Before stating the main result of this section, we observe that the quadratic
variation (along any subdivision) of a continuously differentiable function f
converges to zero. Indeed, ti ≤t |f (ti+1 ) − f (ti )|2 ≤ kf ′ kL∞ ([0,t]) ti ≤t |ti+1 −
P P
Then:
Z t
n
P lim Vt (ϕ) = ϕs ds , for every t ≥ 0 = 1 .
n→∞ 0
hR i
Proof. (i) We first assume that E R+ |ϕt |2 dt < ∞, a condition which
will be relaxed in Step (iii) below. We first fix t > 0 and show that Vtn (ϕ) −→
Rt
ϕ ds P−a.s. as n → ∞, or equivalently:
0 s
2
n→∞
X
−n
Zi −→ 0 P − a.s. where Zi := ϕtni Wtni+1 − Wtni − 2 .
tn
i ≤t
Observe that E[Zi Zj ] = 0 for i 6= j, and E[Zj2 ] = C2−2n E |ϕtni |2 for some
83
Rt
E |ϕtni |2 2−n −→ 0 E [|ϕt |2 ] dt, we conclude that
P
Since tn
i ≤t
2
X X
lim inf E Zi < ∞ .
N →∞
n≤N tn
i ≤t
P P 2
In particular, this shows that the series n≥1 n
ti ≤t Z i is a.s. finite, and
P
therefore tn ≤t Zi −→ 0 P−a.s. as n → ∞.
i
(ii) From the first step of this proof, we can find a zero measure set Ns
for each rational number s ∈ Q. For an arbitrary t ≥ 0, let (sp ) and (s′p )
be two monotonic sequences of rational numbers with sp ր t and s′p ց t.
Then, except on the zero-measure set N := ∪s∈Q Ns , it follows from the
monotonicity of the quadratic variation that
Z sp
ϕu du = lim Vsnp (ϕ) ≤ lim inf Vtn (ϕ)
0 n→∞ n→∞
Z s′p
n n
≤ lim sup Vt (ϕ) ≤ lim Vs′p (ϕ) = ϕu du .
n→∞ n→∞ 0
Rt
Sending p → ∞ shows that Vtn (ϕ) −→ 0 ϕu du as n → ∞ for every ω outside
the zero-measure set N.
(iii) We now consider the general case where ϕ is only known to satisfy
|ϕt |2 dt < ∞ a.s. We introduce the sequence of stopping times
R
R+
Z t
2
Tk := inf t > 0 : |ϕu | du ≥ k , k ≥ 1,
0
and we observe that Tk −→ ∞ P−a.s. as k → ∞. From the previous steps
of this proof, we have P−a.s.
X 2
ϕtni Wtni+1 − Wtni
tn
i ≤t∧Tk
Z t Z t∧Tk
Vtn
= ϕ1[0,Tk ] −→ ϕu 1[0,Tk ] (u)du = ϕu du .
0 0
84
The required limit result follows immediately by sending k → ∞. ♦
Remark 5.8 Proposition 5.14 has a natural direct extension to the multi-
dimensional setting. Let W be a standard Brownian motion in Rd , and let
ϕ be an adapted process with (d × d)−matrix values. Then, P−a.s.
X T Z t
Tr ϕtni Wtni+1 − Wtni Wtni+1 − Wtni −→ Tr [ϕt ] dt , t ≥ 0.
tn 0
i ≤t
Remark 5.9 Inspecting the proof of Proposition 5.14, we see that the
quadratic variation along any subdivision 0 = sn0 < . . . < snn = t satisfies:
n
X 2 X
Wsi+1 − Wsi −→ t P − a.s. whenever sup |sni+1 − sni | −→ 0 .
n n
1≤i≤n
i=1 n≥1
We also observe that the above convergence result holds in probability for
an arbitrary partition (sni , 1 ≤ i ≤ n)n of [0, t] with mesh sup1≤i≤n |sni+1 − sni |
shrinking to zero. See e.g. Karatzas and Shreve [16], Theorem 1.5.8 for the
case ϕ ≡ 1.
85
Chapter 6
Because of this property of the Brownian, one can not hope to define the
stochastic integral with respect to the Brownian motion pathwise. To un-
derstand this, let us forget for a moment about stochastic process. Let
ϕ, f : [0, 1] −→ R be continuous functions, and consider the Riemann sum:
X
ϕ tni−1 f (tni ) − f tni−1
Sn := .
tn
i ≤1
The, if the total variation of f is infinite, one can not guarantee that the
above sum converges for every continuous function ϕ.
In order to circumvent this limitation, we shall make use of the finiteness
of the quadratic variation of the Brownian motion, which allows to obtain
an L2 −definition of stochastic integration.
86
6.1.1 Stochastic integrals of simple processes
Throughout this section, we fix a final time T > 0. A process φ is called
simple if there exists a strictly increasing sequence (tn )n≥0 in R and a sequence
of random variables (ϕn )n≥0 such that
∞
X
φt = ϕ0 1{0} (t) + ϕn 1(tn ,tn+1 ] (t), t ≥ 0,
n=0
and
87
Proposition 6.15 The set of simple processes S is dense in H2 , i.e. for
every φ ∈ H2 , there is a sequence φ(n) n≥0 of processes in S such that
kφ − φ(n) kH2 −→ 0 as n → ∞.
b. We next show that the limit It (φ) does not depend on the choice of the
approximating sequence φ(n) n . Indeed, for another approximating sequence
ψ (n) n of φ, we have
0 (n)
2 ≤
I 0 ψ (n) − I 0 φ(n)
2 +
I 0 φ(n) − It (φ)
2
I ψ
t − I t (φ) L t t t
(n) (n)
0 L (n)
L
≤ ψ −φ
H 2 +
It φ − It (φ) 2
L
−→ 0 as n → ∞.
c. We finally show that the uniquely defined limit It (φ) satisfies the ana-
logue properties of (6.2)-(6.3):
88
To see that (6.4) holds, we directly compute with φ(n) n
an approximating
sequence of φ in H2 that
+
I 0 (φ(n) ) − Is (φ)
2
s L
by (6.2). By the Jensen inequality and the law of iterated expectations, this
provides
Z t Z t h 2 i
2
|φ(n) 2
= lim E I 0 φ(n) = E I (φ)2 .
E |φs | ds = lim E s | ds
0 n→∞ 0 n→∞
is the unique limit in L2 of the sequence It0 (φ(n) ) n for every choice of an
approximating sequence φ(n) n of φ in H2 . Moreover, {It (φ), t ≥ 0} is a
square integrable martingale satisfying the Itô isometry
"Z 2 #
t Z t
E φs dWs =E |φs |2 ds , t ≥ 0.
0 0
Observe that, since the Brownian motion has continuous trajectories, the
stochastic integral of a simple process I 0 (φ) also has continuous sample paths
89
by the definition in (6.1). It is not clear whether this property is inherited
by the extension I of I0 to H2 . The following result shows that it is indeed
the case, and provides the useful Doob’s maximal inequality for martingales
(6.6).
Rt
Proposition 6.16 Let φ be a process in H2 , and set Mt := 0 φs ·dWs , t ≥ 0.
Then M has continuous sample paths a.s. and
Z t
2 2
E sup Ms ≤ 4E |Ms | = E |φs | ds . (6.6)
s≤t 0
Proof. 1. Set Mt∗ := sups≤t Ms . In this step, we prove that the maximal
inequality (6.6) holds if M is known to to be continuous. Consider the
sequence of stopping times:
θn := inf {t ≥ 0 : Mt∗ ≥ n}
Then
2
Z t∧θn
∗ ∗
Mt∧θ n
= 2Mt∧θn Mt∧θ n
−2 Ms∗ dMs
0
and we obtain the maximal inequality by sending n to infinity and using the
monotone convergence theorem.
90
2. We now show that the process M is continuous. By definition of the
stochastic integral, Mt is the L2 −limit of Mtn := It0 (φn ) for some sequence
(φn )n of simple processes converging to φ in H2 . By definition of the stochas-
tic integral of simple integrands in (6.1), notice that the process {Mtn −Mtm =
It0 (φn ) − It0 (φm ), t ≥ 0} is a.s. continuous. We then deduce from the first
step that:
n m n m m n
E sup |Ms − Ms | ≤ E sup(Ms − Ms ) + E sup(Ms − Ms )
s≤t s≤t s≤t
Z t
≤ 8E |φns − φm 2
s | ds .
0
91
To do this, we consider the sequence of stopping times
Z t
2
τn := inf t > 0 : |φu | du ≥ n .
0
τn −→ ∞ P − a.s. when n → ∞.
For fixed n > 0, the stopped process φnt := φt 1t≤τn is in H2 . Then the
stochastic integral It (φn ) is well-defined by Theorem 6.13. Since (τn )n is a
non-decreasing and P [τn ≥ t for some n ≥ 1] = 1, it follows that the limit
Proposition 6.17 Let φ be a process satisfying (6.7). Then, for every T >
0, the process {It (φ), 0 ≤ t ≤ T } is a local martingale.
92
Proof. The above defined sequence (τn )n is easily shown to be a localizing
sequence. The result is then a direct consequence of the martingale property
of the stochastic integral of a process in H2 . ♦
for every T ≥ 0.
Proof. 1 We first fix T > 0, and we show that the above Itô’s formula
holds with probability 1. By possibly adding a constant to f we may assume
that f (0, 0) = 0. Let tni := i2−n for i ≥ 1, and n(T ) := ⌊T 2n ⌋, so that
tnn(T ) ≤ T < tnn(T )+1 .
1.a We first decompose
Xh i
f tnn(T )+1 , Wtnn(T )+1 = f tni+1 , Wtni+1 − f tni , Wtni+1
tn
i ≤T
Xh i
+ f tni , Wtni+1 − f tni , Wtni .
tn
i ≤T
93
By a Taylor expansion, this provides:
X
ITn (fx ) := fx tni , Wtni · Wtni+1 − Wtni
tn
i ≤T
X
= f tnn(T )+1 , Wtnn(T )+1 − ft τin , Wtni+1 2−n
tn
i ≤T
1 X
Tr fxx (tni , ξin ) − fxx tni , Wtni ∆ni W ∆ni W T
−
2 tn ≤T
i
1 X
Tr fxx tni , Wtni ∆ni W ∆ni W T ,
− (6.9)
2 tn ≤T
i
where ∆ni W := Wtni+1 − Wtni , τin is a random variable with values in tni , tni+1 ,
and ξin = λni Wtni + (1 − λni ) Wtni+1 for some random variable λni with values in
[0, 1].
1.b Since a.e. sample path of the Brownian motion is continuous, and
therefore uniformly continuous on the compact interval [0, T + 1], it follows
that
n
f tn(T )+1 , Wtn(T )+1 −→ f (T, WT ) , P − a.s.
n
RT
n −n
P
n
t ≤T ft τi , Wtn
i+1
2 −→ 0
ft (t, Wt )dt P − a.s.
i
1.c In order to complete the proof of Itô’s formula for fixed T > 0, it
remains to prove that
Z T
n
IT (fx ) −→ fx (t, Wt ) · dWt P − a.s. along some subsequence.(6.10)
0
94
Notice that ITn (fx ) = IT0 φ(n) where φ(n) is the simple process defined by
(n)
X
fx tni , Wtni 1[tn ,tn ) (t),
φt = t ≥ 0.
i i+1
tn
i ≤T
Remark 6.11 Since, with probability 1, the Itô formula holds for every
T ≥ 0, it follows that the Itô’s formula holds when the deterministic time T
is replaced by a random time τ .
where µ and σ are adapted processes with values in Rn and MR (n, d), re-
spectively and satisfying
Z T Z T
|µs |ds + |σs |2 ds < ∞ a.s.
0 0
95
Observe that stochastic integration with respect to the Itô process X re-
duces to the stochastic integration with respect to W : for any F−adapted
RT RT
Rn −valued process φ with 0 |σtT φt |2 dt + 0 |φt · µt |dt < ∞,a.s.
Z T Z T Z T
φt · dXt = φt · µt dt + φt · σt dWt
0 0 0
Z T Z T
= φt · µt dt + σtT φt · dWt .
0 0
for every T ≥ 0.
n Rt Rt o
Proof. Let θN := inf t : max |Xt − X0 |, 0 σs2 ds, 0 |µs |ds ≥ N . Obvi-
ously, θN → ∞ a.s. when N → ∞, and it is sufficient to prove Itô’s formula
on [0, θN ], since any t ≥ 0 can be reached by sending N to infinity. In view
Rt Rt
of this, we may assume without loss of generality that X, 0 µs ds, 0 σs2 ds
are bounded and that f has compact support. We next consider an approx-
imation of the integrals defining Xt by step functions which are constant on
intervals of time [tni−1 , tni ) for i = 1, . . . , n, and we denote by Xn the resulting
simple approximating process. Notice that Itô’s formula holds true for X n
on each interval [ti−1 , ti ) as a direct consequence of Theorem 6.14. The proof
of the theorem is then concluded by sending n to infinity, and using as much
as needed the dominated convergence theorem. ♦
Xt := X0 + bt + σWt , t ≥ 0 ,
96
where b is a vector in Rd and σ is an (d×d)−matrix. Let f be a C 1,2 (R+ , Rd )
function. Show that
2
∂f ∂f 1 ∂ f T
df (t, Xt) = (t, Xt )dt + (t, Xt ) · dXt + Tr (t, Xt )σσ .
∂t ∂x 2 ∂x∂xT
We conclude this section by providing an example of local martingale
which fails to be a martingale, although positive and uniformly integrable.
This shows the limitation of the discrete-time result of Lemma 3.1.
97
This shows that supt≥0 E [Xt2 ] < ∞. In particular, X is uniformly integrable.
where g is a given function from Rd to R, k and f are functions from [0, T ]×Rd
to R, b is a constant vector in Rd , and σ is a constant d × d matrix. This
is the so-called Dirichlet problem. The following result provides a stochastic
representation for the solution of this purely deterministic problem.
Theorem 6.16 Assume that the function k is uniformly bounded from below,
and f has polynomial growth in x uniformly in t. Let v be a C 1,2 [0, T ), Rd
solution of (6.11) with polynomial 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
k(u,Xut,x )du
where Xst,x := x + b(s − t) + σ(Ws − Wt ) and βst,x := e− t for
t ≤ s ≤ T.
98
by the PDE satisfied by v in (6.11). Then:
E βτt,x t,x
v τ n , X − v(t, x)
Z nτn τn
t,x ∂v t,x
= E βs −f (s, Xs )ds + s, Xs σdWs .
t ∂x
Now observe that the integrands in the stochastic integral is bounded by def-
inition of the stopping time τn and the smoothness of v. 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 . (6.13)
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
−→ (6.14)
t Z
T
βst,x f s, Xst,x ds + βTt,x g XTt,x
=
t
99
We also observe that the Feynman-Kac representation formula has an
important numerical implication. Indeed it opens the door to the use of
Monte Carlo methods in order to obtain a numerical approximation of the
solution of the partial differential equation (6.11). For sake of simplicity, we
provide the main idea in the case f = k = 0. Let X (1) , . . . , X (k) be an iid
sample drawn in the distribution of XTt,x , and compute the mean:
k
1X
g X (i) .
v̂k (t, x) :=
k i=1
By the Law of Large Numbers, it follows that v̂k (t, x) −→ v(t, x) P−a.s.
Moreover 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,
100
a2
introduce the equivalent measure Q := eaN − 2 · P. Then, the above equality
says that the Q−distribution of N coincides with the P−distribution of N +a,
i.e.
101
Rt
Since the random variable s (h(u) + λ) · dWu is a centered Gaussian with
Rt
variance s |h(u) + λ|2 du, independent of Fs , this provides:
1 t 2 1 t 2
R R
EQ eλ·(Wt −Ws ) Fs = e− 2 s |h(u)| du e 2 s |h(u)+λ| du
1 2 (t−s)+λ
Rt
= e2λ s h(u)du
.
Let us observe that the above result can be extended to the case where
h is an adapted stochastic process satisfying some convenient integrability
conditions. This is the so-called Girsanov theorem. In the context of these
lectures, we shall only need the Gaussian framework of the Cameron-Martin
formula.
102
Lemma 6.5 Let φ be a bounded F−progressively measurable process. Then
φ can be approximated by a sequence of simple processes in H2 .
Proof. In the present setting, the process Y (k) , defined in the proof of the
previous Lemma 6.5, is measurable but is not known to be adapted. For each
ε > 0, there is an integer k ≥ 1 such that Y ε := Y (k) satisfies kY ε −XkH2 ≤ ε.
Then, with Xt = X0 for t ≤ 0:
103
We now introduce
X
ϕn (t) := 1{0} (t) + tni 1(tni−1 ,tni ] ,
i≥1
and
(n,s)
Xt := Xϕn (t−s)+s , t ≥ 0, s ∈ (0, 1].
and the required result follows from the dominated convergence theorem.
♦
Proof. We only have to extend Lemma 6.6 to the case where φ is not
necessarily bounded. This is easily achieved by applying Lemma 6.6 to the
bounded process φ ∧ n, for each n ≥ 1, and passing to the limit as n → ∞.
♦
104
Chapter 7
105
for 1 ≤ i ≤ d. Here b = b1 , . . . , bd and σ = (σ ij ) are determinstic functions
on [0, T ] with values respectively in Rd and MR (d, d), the set of d×d matrices
with real coefficients. In order for the above stochastic integral to be well-
defined, we assume that
Z T
|b(t)| + |σ(t)|2 dt < ∞.
0
or in vector notations:
where, for z ∈ Rn , diag[z] denotes the diagonal matrix with diagonal elements
z i . We assume that the MR (d, d)−matrix σ(t) is invertible for every t ∈
[0, T ], and we introduce the function
called the risk premium. We shall frequently make use of the discounted
process
Z t
St
S̃t := 0 = St exp − r(u)du ,
St 0
where 1 = (1, . . . , 1) ∈ Rd .
106
7.1 Portfolio and wealth process
A portfolio strategy is an F−adapted process π = {πt , 0 ≤ t ≤ T } with
values in Rd . For 1 ≤ i ≤ n and 0 ≤ t ≤ T , πti is the amount (in Euros)
invested in the risky asset S i .
As in the discrete-time financial market, we shall consider self-financing
strategies. Let Xt denote the wealth process at time t induced by the portfolio
strategy π. Then, the amount invested in the non-risky asset is Xt − ni=1 πti
P
= Xt − πt · 1.
Under the self-financing condition, the dynamics of the wealth process is
given by
n
X πti Xt − πt · 1
dXt = i
dSti + 0
dSt0
i=1
S t S t
We still have to place further technical conditions on π, at least for the above
wealth process to be well-defined as a stochastic integral. Before this, let us
observe that, assuming that the risk premium function satisfies:
Z T
0 2
λ (t) dt < ∞ ,
0
107
is a Brownian motion under the equivalent probability measure
P0 := ZT0 · P on FT , (7.4)
where
Z T T
1
Z
ZT0 := exp − 0
λ (u) · dWu − 0 2
|λ (u)| du . (7.5)
0 2 0
inf X̃tx,π ≥ −C
0≤t≤T
for some constant C > 0 which may depend on π. The collection of all
admissible portfolio processes will be denoted by A.
The finite credit line condition is needed in order to exclude any arbitrage
opportunity, and will be justified in the subsequent subsection.
108
Definition 7.19 We say that the financial market contains no arbitrage op-
portunities if for any admissible portfolio process π ∈ A,
The purpose of this section is to show that the financial market described
above contains no arbitrage opportunities. Our first observation is that P0
is a risk neutral measure, or an equivalent martingale measure, for the price
process S, i.e.
n o
the process S̃t , 0 ≤ t ≤ T is a P0 − martingale. (7.7)
109
developed in Example 6.14. Observe that the latter example also shows that
the finite credit line condition on admissible portfolios does not guarantee
that the discounted wealth process is a P0 −martingale.
In the continuous-time framework, one has to rule out the so-called class
of doubling strategies which are naturally excluded in finite discrete-time.
Doubling strategies are well-known in Casinos. Suppose that a player starts
playing in the Casino without any initial funds. So he borrows some capital
x0 > 0 (with zero interest rate, say) and invests it on some game whose payoff
is 2x0 with some positive probability p, and 0 with probability 1 − p. At the
end of this first round, the player’s wealth is X1 = x0 with probability p, and
X1 = −x0 with probability 1−p. If X1 = −x0 , then the player borrows again
the amount x1 = 2x0 , i.e. doubles his debt, and plays again. At the end of
this second round, the player’s wealth, conditionally on X1 = −x0 , is again
either X2 = x0 = x0 with probability p, and X2 = −2x0 with probability
1 − p. And so on... If the player is allowed to play an infinite number of
rounds, then he would perform this doubling strategy up to the first moment
that his wealth is positive, which happens at some finite time with probability
1.
In conclusion, this strategy allows to perform some positive gain without
any initial investment ! the drawback of this strategy is, however, that the
wealth process can be arbitrarily large negative. So it is only feasible if there
is no restriction on borrowing, which is not realistic.
Since the continuous-time analogue of the discrete-time random walk is
the Brownian motion, see Chapter 4.2, the above doubling strategies provide
a good intuition for our continuous-time model. The following result from
stochastic integration is due to Dudley [10], and is stated for completeness
in order to show how things can get even worse in continuous-time.
110
satisfying
Z T Z T
ξ = φt dWt and |φt |2 dt < ∞ P − a.s.
0 0
Proof. Let (Tn )n be a localizing sequence of stopping times for the local
martingale M, i.e. Tn −→ ∞ a.s. and {Mt∧Tn , 0 ≤ t ≤ T } is a martingale
for every n. Then :
Exercise 7.14 Show that the conclusion of Lemma 7.8 holds true under the
weaker condition that the local martingale M is bounded from below by a
martingale.
111
Theorem 7.19 The continuous-time financial market described above con-
tains no arbitrage opportunities.
0,π
Proof. For π ∈ A, the corresponding h idiscounted wealth process X̃ is a
P0 −super-martingale. Then EP X̃T0,π ≤ 0. Recall that P0 is equivalent
0
over all admissible consumption/investment policies (those for which the in-
tegrals exist and the wealth remains nonnegative). This problem was first
studied by R. Meron in his seminal paper [18] and is known as the Merton
consumption/investment problem.
112
The first step is to reduce the dimension of the problem by introducing
portfolio fractions
θti Sti
ωti = PN j j
if Xt 6= 0 and 0 otherwise, i = 1, . . . , N.
j=0 θt St
The fractions allow to get rid of the dependence on S in the portfolio dy-
namics which becomes
Note that the positivity constraint is essential to be able to work with port-
folio fractions.
Observe that v(0, x) gives the solution to our optimization problem and
v(T, x) = U(x) produces the terminal condition. We now need to link the
value functions for different values of t via a kind of “infinitesimal” backward
induction. The basic idea is that if we know the value function at date t + h,
we can use this information to compute the value function at date t. The
relationship between the two is called the dynamic programming principle:
Z t+h
−ρ(s−t) −ρh
v(t, x) = sup E e u(cs )ds + e v(t + h, Xt+h )|Xt = x .
ω,c t
113
Suppose that v is a C 1,2 function. Then we can apply the Itô formula (The-
orem 6.15) obtaining
"Z
t+h n ∂v ∂v ∂v
0 = sup E e−ρ(s−t) u(cs ) − cs − ρv(s, Xs ) + + Xs (r + ωsT σλ)
ω,c t ∂x ∂t ∂x
Z t+h #
Xs2 ∂ 2 v T T o ∂v
+ ω σσ ωs ds + e−ρ(s−t) Xs ωsT σdWs .
2 ∂x2 s t ∂x
x2 ∂ 2 v T T
∂v ∂v ∂v T
0 = sup u(c) − c − ρv(t, x) + + x (r + ω σλ) + ω σσ ω ,
ω,c ∂x ∂t ∂x 2 ∂x2
from which the optimal investment policy can be immediately computed (if
v is known):
∂v
ω=− ∂x
∂2 v
σ −1 λ.
x ∂x 2
114
of ρ reduces to multiplying the terminal utility by a constant, which means
that we can set ρ = 0 without loss of generality. With these assumptions,
the HJB equation simplifies to
x2 ∂ 2 v T T
∂v ∂v T
0 = sup + x (r + ω σλ) + ω σσ ω .
ω ∂t ∂x 2 ∂x2
From the form of the terminal condition, we make an “educated guess” that
xγ
v(t, x) = f (t)
γ
for some function t. This substitution reduces the HJB equation to
∂f
+ γf (t) sup r + ω T σλ + (γ − 1)ω T σσ T ω ,
∂t ω
which is solved by
σ −1 λ
ω∗ =
1−γ
γkλk2
f (t) = eε(T −t) , ε= + rγ.
2(1 − γ)
The optimal investment strategy in Merton’s model with power utility there-
fore consists in keeping a constant proportion of one’s wealth in each asset,
including the bank account.
It remains to check the optimality of the strategy ω ∗, that is, to show
∗
that v(0, x) = E[U(XTω )] and that v(0, x) ≥ E[U(XTω )] for every admissible
strategy ω. Let ω be one such strategy. Since v satisfies the HJB equation,
Itô formula yields
Z T
∂v ω T
v(0, x) + Xt ωt σdWt ≥ U(XTω ),
∂x
Z0 t
∂v ω T
v(0, x) + Xs ωs σdWs ≥ v(0, Xtω ) ≥ 0 ∀t ≤ T.
0 ∂x
The second line shows that the stochastic integral is a local martingale
bounded from below and thus a supermartingale. Taking the expectation
115
of the first line, we then get
Z T
∂v ω T
v(0, x) ≥ v(0, x) + E X ω σdWt ≥ E[U(XTω )].
0 ∂x t t
For the optimal strategy, we have equality in the HJB equation, which implies
Z T
(Xtω )γ ωt∗T σdWt = U(XTω )
∗ ∗
v(0, x) +
0
∗
Moreover, using the explicit form of Xtω , it is easy to check that
Z T
ω ∗ 2γ
E (Xt ) dt < ∞,
0
x2 ∂ 2 v T T
∂v ∂v T
0 = sup u(c) − c − ρv(x) + x (r + ω σλ) + ω σσ ω .
ω,c ∂x ∂x 2 ∂x2
116
γ
Once again, we guess the solution: v(x) = Kx γ
, and substituting this into the
equation, we see, as before, that the optimal investment policy is a constant
proportion strategy:
σ −1 λ
ω∗ = .
1−γ
After some more computations, we find the constant K and the optimal
consumption policy:
K = (α∗ )γ−1
and c∗ = α∗ x,
γkλk2
∗ 1
where α = ρ − γr − .
1−γ 2(1 − γ)
The agent must therefore consume a constant fraction of his/her current
wealth. The portfolio value corresponding to the optimal investment and
consumption policy is easily computed:
kλk2 λT Wt
∗ r−ρ
Xt = x exp + t+ .
1 − γ 2(1 − γ) 1−γ
To verify that the strategy we have found is indeed optimal, we first show, by
the same arguments as in the previous example, that for any consumption-
investment pair (c, ω),
Z T Z T
−ρt −ρT ω −ρt
v(x) ≥ E e u(ct )dt + e v(XT ) ≥ E e u(ct )dt ,
0 0
and by making T go to ∞,
Z ∞
−ρt
v(x) ≥ E e u(ct )dt .
0
117
7.3.4 The martingale approach
In this section we study a different approach to the problem considered in
section 7.3.2:
sup E[U(XTω )],
ω∈A
E[e−rT ZT XTω ] ≤ x.
Similarly to how this was done in discrete time, we decompose the portfolio
optimization problem into a problem of finding the optimal contingent claim
under the budget constraint, and the problem of replicating this claim with
an admissible portfolio. The optimal claim problem is
XT∗ = I(ye−rT ZT ),
E[e−rT ZT I(ye−rT ZT )] = x.
118
In the case of power utility,
xγ 1 1 − γ γ−1
γ
U(x) = ⇒ I(y) = y γ−1 V (y) = y
γ γ
and the candidate optimal claim is given by
1
ZTγ−1 λT WT 1 − 2γ kλk2 T
XT∗ rT
= xe γ
rT
= xe exp + ,
E[ZT γ−1
] 1−γ (1 − γ)2 2
On the other hand, for the optimal claim X ∗ it is easy to see that there exists
an y ∗ for which
E[U(XT∗ )] = E[V (y ∗e−rT ZT )] + xy ∗ .
119
Chapter 8
Theorem 8.20 (i) Suppose that the function g is bounded. Then, the no
arbitrage price of the Vanilla option G at time 0 is given by
h RT i
0
p0 (G) = v(0, S0 ) where v(t, s) := EP e− t r(u)du g (ST ) St = s .
(ii) Assume further that the market price p0 is a continuous map from L1 (P0 )
to R. Then (i) holds whenever g(ST ) ∈ L1 (P0 ).
120
(iii) The above function v ∈ C 1,2 [0, T ), Rd . If g is bounded from below,
then
∂v
π̂t := diag[St ] (t, St ) , 0 ≤ t ≤ T ,
∂s
defines an admissible portfolio strategy in π̂ ∈ A which perfectly replicates G,
v(0,S ),π̂
i.e. XT 0 = g (ST ).
and we argue that the market price p0 (G) lies inside the no-arbitrage bounds:
1. For an arbitrary x > V0 (G), we have X̃Tx,π ≥ G̃ a.s. for hsomei π ∈ A. Since
0
X̃ x,π is a P0 −supermartingale, we deduce that x ≥ EP G̃ , and therefore
h i
0
V0 (G) ≥ EP G̃ .
h i
0
2. In order to prove the reverse inequality V0 (G) ≥ EP G̃ , we show in
Steph 3 ibelow that G can be perfectly hedged starting from the initial capital
0
EP G̃ . Since −G satisfies the same conditions as G, the conclusions of
Step 1 and the first part of this step hold also for −G. The required results
then follow from the no-arbitrage bounds (8.1).
3. It remains to show that G can be perfectly hedged h byi means of an ad-
P0
missible portfolio, starting from the initial capital E G̃ . Since ln Sti has a
Gaussian distribution, satisfying the heat equation, the function v satisfies
∂2v
∂v ∂v 1 T
+ rs · + Tr diag[s]σσ diag[s] − rv = 0
∂t ∂s 2 ∂s∂sT
121
for t < T , s ∈ Rd+ . We then deduce from Itô’s formula that
Z T
−rT v(0,S ),π̂
e v(T, ST ) = v(0, S0) + e−ru π̂u · σdWu0 = X̃T 0 .
0
v(0,S ),π̂
Since v(T, s) = g(s), this shows that XT 0 = g (ST ). Since g (ST ) is
square integrable, as a consequence of the polynomial growth condition on
g, it follows that X̃ v(0,S0 ),π̂ is a P0 −square integrable martingale (exercise !),
and therefore
Rt
v(0,S0 )π̂
X̃t = e− 0
r(u)du
v(t, St )
Remark 8.12 In the above proof, we used the fact that the option price
v(t, s) satisfies the linear partial differential equation
∂2v
∂v ∂v 1 T
+ rs · + Tr diag[s]σσ diag[s] − rv = 0 , (t, s) ∈ [0, T ) × R+
∂t ∂s 2 ∂s∂sT
v(T, s) = g(s) , s ≥ 0 . (8.2)
122
are accepting without proof that p is a deterministic function of time and
the spot price, this has been in fact proved in the previous section.
2. The holder of the contingent claim completes his portfolio by some in-
vestment in the risky assets. At time t, he decides to holds −∆i shares of
the risky asset S i . Therefore, the total value of the portfolio at time t is
Pt := p(t, St ) − ∆ · St , 0≤t<T.
3. Considering delta as a constant vector in the time interval [t, t + dt), and
assuming that the function p is of class C 1,2 , the variation of the portfolio
value is given by :
∂p
dPt = Lp(t, St )dt + (t, St ) · dSt − ∆ · dSt .
∂s
∂p 2
∂ p
where Lp = ∂t
+ 12 Tr[diag[s]σσ T diag[s] ∂s∂s T ]. In particular, by setting
∂p
∆ = ,
∂s
we obtain a portfolio value with finite quadratic variation
4. The portfolio Pt is non-risky since the variation of its value in the time
interval [t, t + dt) is known in advance at time t. Then, by the no-arbitrage
argument, we must have
∂2p
∂p ∂p 1 T
+ rs · + Tr diag[s]σσ diag[s] − rp = 0 ,
∂t ∂s 2 ∂s∂sT
which is exactly the PDE obtained in the previous section.
123
8.3 The Black and Scholes model for European
call options
8.3.1 The Black-Scholes formula
In this section, we consider the one-dimensional Black-Scholes model d = 1
so that the price process S of the single risky asset is given in terms of the
P0 −Brownian motion W 0 :
σ2
0
St = S0 exp r − t + σWt , 0≤t≤T. (8.5)
2
Observe that the random variable St is log-normal for every fixed t. This is
the key-ingredient for the next explicit result.
Proposition 8.18 Let G = (ST − K)+ for some K > 0. Then the no-
arbitrage price of the contingent claim G is given by the so-called Black-
Scholes formula :
2 2
p − 0(G) = S0 N d+ (S0 , K̃, σ T ) − K̃ N d− (S0 , K̃, σ T ) ,(8.6)
where
ln (s/k) 1 √
K̃ := Ke−rT , d± (s, k, v) := √ ± v, (8.7)
v 2
and the optimal hedging strategy is given by
π̂t = St N d+ (St , K̃, σ 2 (T − t) , 0≤t≤T. (8.8)
Proof. This formula can be derived by various methods. One can just
calculate directly the expected value by exploiting the explicit probability
density function of the random variable ST . One can also guess a solution
for the valuation PDE corresponding to the call option. We shall present
another method which relies on the technique of change of measure and
reduces considerably the computational effort. We first decompose
0
h i h i
p0 (G) = EP S̃T 1{S̃T ≥K̃} − K̃ P0 S̃T ≥ K̃ (8.9)
124
where as usual, the tilda notation corresponds to discounting, i.e. multipli-
cation by e−rT in the present context.
1. The second term is directly computed by exploiting the knowledge of the
distribution of S̃T :
" #
2 2
h i ln ( S̃ T /S 0 ) + (σ /2)T ln ( K̃/S 0 ) + (σ /2)T
P0 S̃T ≥ K̃ = P0 √ ≥ √
σ T σ T
!
ln (K̃/S0 ) + (σ 2 /2)T
= 1−N √
σ T
= N d− (S0 , K̃, σ 2 T ) .
2. As for the first expected value, we define the new measure P1 := ZT1 · P0
on FT , where
σ2
1 S̃T
ZT := exp σWT − T = .
2 S0
125
Figure 8.1: The Black-Scholes formula as a function of S and t
126
Black-Sholes formula:
p0 (G) = S0 N d+ (S0 , K̃, v(T )) − K̃ N d− (S0 , K̃, v(T )) (8.10)
where
RT
Z T
− r(t)dt
K̃ := Ke 0 , v(T ) := σ 2 (t)dt . (8.11)
0
127
8.3.3 Option on a dividend paying stock
When the risky asset S pays out some dividend, the previous theory requires
some modifications. We shall first consider the case where the risky asset
pays a lump sum of dividend at some pre-specified dates, assuming that the
process S is defined by the Black-Scholes dynamics between two successive
dates of dividend payment. This implies a downward jump of the price
process upon the payment of the dividend. We next consider the case where
the risky asset pays a continuous dividend defined by some constant rate.
The latter case can be viewed as a model simplification for a risky asset
composed by a basket of a large number of dividend paying assets.
128
and the no-arbitrage European call option price is given by
P0
rT +
2 2
E e (ST − K) = Ŝ0 N d+ (Ŝ0 , K̃, σ T ) − K̃ N d− (Ŝ0 , K̃, σ T ) ,
with K̃ = Ke−rT , i.e. the Black-Scholes formula with modified spot price
from S0 to Ŝ0 .
In other words, we can reduce the problem the non-dividend paying security
case by increasing the
n value of the security.
o By the no-arbitrage theory, the
−rt (δ)
discounted process r St , t ≥ 0 must be a martingale under the risk
neutral measure P0 :
2 2
(δ) (δ) r− σ2 t+σ Wt0 r− σ2 t+σ Wt0
St = S0 e = S0 e , t ≥ 0,
We are now in a position to provide the call option price in closed form:
0 0
EP e−rT (ST − K)+ = e−δT EP e−(r−δ)T (ST − K)+
h
−δT (δ) 2
= e S0 N d+ (S0 , K̃ , σ T )
i
− K̃ (δ) N d− (S0 , K̃ (δ) , σ 2 T ) (8.14)
,
129
where
K̃ (δ) := Ke−(r−δ)T .
130
is the maximum likelihood estimator for the parameter σ 2 . The estimator
σ̂n is called the historical volatility parameter.
The natural way to implement the Black-Scholes model is to plug the
historical volatility into the Black-Scholes formula
2 2
BS (St , σ, K, T ) := St N d+ St , K̃, σ T − K̃N d− St , K̃, σ T(8.15)
to compute an estimate of the option price, and into the optimal hedge ratio
∆ (St , σ, K, T ) := N d+ St , K̃, σ 2 T (8.16)
in order to implement the optimal hedging strategy.
Unfortunately, the options prices estimates provided by this method per-
forms very poorly in terms of fitting the observed data on options prices.
Also, the use of the historical volatility for the hedging purpose leads to a
very poor hedging strategy, as it can be verified by a back-testing procedure
on observed data.
3. On liquid options markets, prices are given to the practitioners and are
determined by the confrontation of demand and supply on the market. There-
fore, their main concern is to implement the corresponding hedging strategy.
131
To do this, they use the so-called calibration technique, which in the present
context reduce to the calculation of the implied volatility parameter.
It is very easily checked the Black-Scholes formula (8.15) is a on-to-one
function of the volatility parameter, see (8.21) below. Then, given the ob-
servation of the call option price Ct∗ (K, T ) on the financial market, there
exists a unique parameter σ imp which equates the observed option price to
the corresponding Black-Scholes formula:
BS St , σtimp(K, T ), K, T = Ct∗ (K, T ) , (8.17)
provided that Ct∗ satisfies the no-arbirage bounds of Subsection 1.4. This
defines, for each time t ≥ 0, a map (K, T ) 7−→ σtimp (K, T ) called the implied
volatility surface. For their hedging purpose, the option trader then computes
the hedge ratio
∆imp
t (T, K) := ∆ S t , σt
imp
(K, T ), K, T .
If the constant volatility condition were satisfied on the financial data, then
the implied volatility surface would be expected to be flat. But this is not the
case on real life financial markets. For instance, for a fixed maturity T , it is
usually observed that that the implied volatility is U-shaped as a function of
the strike price. Because of this empirical observation, this curve is called the
volatility smile. It is also frequently argued that the smile is not symmetric
but skewed in the direction of large strikes.
From the conceptual point of view, this practice of options traders is in
contradiction with the basics of the Black-Scholes model: while the Black-
Scholes formula is established under the condition that the volatility pa-
rameter is constant, the practical use via the implied volatility allows for a
stochastic variation of the volatility. In fact, by doing this, the practioners
are determining a wrong volatility parameter out of a wrong formula !
Despite all the criticism against this practice, it is the standard on the
derivatives markets, and it does perform by far better than the statistical
method. It has been widely extended to more complex derivatives markets as
the fixed income derivatives, defaultable securities and related derivatives...
132
Risk control variables: the Greeks
With the above definition of the implied volatility, all the parameters
needed for the implementation of the Black-Scholes model are available. For
the purpose of controlling the risk of their position, the practitioners of the
options markets various sensitivities, commonly called Greeks, of the Black-
Scholes formula to the different variables and parameters of the model. The
following picture shows a typical software of an option trader, and the ob-
jective of the following discussion is to understand its content.
133
2 /2
where N′(x) = (2π)−1/2 e−x .
2. Gamma: is defined by
∂ 2 BS
Γ (St , σ, K, T ) := (St , σ, K, T )
∂s2
1
= √ N′ d+ St , K̃, σ 2 T . (8.18)
St σ T − t
The interpretation of this risk control coefficient is the following. While the
simple Black-Scholes model assumes that the underlying asset price process
is continuous, practitioners believe that large movemements of the prices,
or jumps, are possible. A stress scenario consists in a sudden jump of the
underlying asset price. Then the Gamma coefficient represent the change in
the hedging strategy induced by such a stress scenario. In other words, if
the underlying asset jumps immediately from St to St + ξ, then the option
hedger must immediately modify his position in the risky asset by buying
Γt ξ shares (or selling if Γt ξ < 0).
Given this interpretation, a position with a large Gamma is very risky, as
it would require a large adjustment in case of a stress scenario.
3. Rho: is defined by
∂BS
ρ (St , σ, K, T ) := (St , σ, K, T )
∂r
2
= K̃(T − t)N d− St , K̃, σ T , (8.19)
4. Theta: is defined by
∂BS
θ (St , σ, K, T ) := (St , σ, K, T )
∂T
1 √
= St σ T − tN′ d− St , K̃, σ 2 T , (8.20)
2
134
is also called the time value of the call option. This coefficient isolates the
depreciation of the option when time goes on due to the maturity shortening.
∂BS
V (St , σ, K, T ) := (St , σ, K, T )
∂σ
√
= St T − tN′ d− St , K̃, σ 2 T . (8.21)
This control variable provides the exposition of the call option price to the
volatility risk. Practitioners are of course aware of the stochastic nature
of the volatility process (recall the smile surface above), and are therefore
seeking a position with the smallest possible Vega in absolute value.
135
Bibliography
[1] Black F. and Scholes M. (1973), The pricing of options and cor-
porate liabilities, Journal of Political Economy, 81, 637-654.
[2] Carr P., Ellis K. and Gupta V. (1998), Static Hedging of Exotic
Options, Journal of Finance, June 1998, pp. 1165-90.
[4] Cox J.C., Ross S.A. and Rubinstein M. (1979), Option pricing : a
simplified approach, Journal of Financial Economics, 7, 229-263.
136
[10] Dudley, R.M.(1977), Wiener functionals as Itô integrals. Annals
of Probability 5, 140-141.
[12] El Karoui N. and Rochet J.C. (1989), A pricing formula for op-
tions on zero-coupon bonds. Preprint.
137
[20] Napp C. (2000), Thèse de Doctorat, Université Paris I Panthéon-
Sorbonne.
138