Driver PDF
Driver PDF
Driver PDF
Ovidiu Calin
Department of Mathematics
Eastern Michigan University
Ypsilanti, MI 48197 USA
[email protected]
Preface
i
ii O. Calin
Contents
I Stochastic Calculus 3
1 Basic Notions 5
1.1 Probability Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Sample Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Events and Probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Random Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Distribution Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Basic Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Independent Random Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.8 Expectation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.9 Radon-Nikodym’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.10 Conditional Expectation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.11 Inequalities of Random Variables . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.12 Limits of Sequences of Random Variables . . . . . . . . . . . . . . . . . . . . . 20
1.13 Properties of Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.14 Stochastic Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
iii
iv O. Calin
4 Stochastic Integration 59
4.0.3 Nonanticipating Processes . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.0.4 Increments of Brownian Motions . . . . . . . . . . . . . . . . . . . . . . 59
4.1 The Ito Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2 Examples of Ito integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2.1 The case Ft = c, constant . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2.2 The case Ft = Wt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.3 The Fundamental Relation dWt2 = dt . . . . . . . . . . . . . . . . . . . . . . . . 63
4.4 Properties of the Ito Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.5 The Wiener Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.6 Poisson Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.6.1 An Workout Example: the case Ft = Mt . . . . . . . . . . . . . . . . . . 70
5 Stochastic Differentiation 73
5.1 Differentiation Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.2 Basic Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Ito’s Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.3.1 Ito’s formula for diffusions . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.3.2 Ito’s formula for Poisson processes . . . . . . . . . . . . . . . . . . . . . 78
5.3.3 Ito’s multidimensional formula . . . . . . . . . . . . . . . . . . . . . . . 79
8 Martingales 117
8.1 Examples of Martingales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
8.2 Girsanov’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Stochastic Calculus and Applications to Finance v
Stochastic Calculus
3
Chapter 1
Basic Notions
Example 1.2.1 Flip a coin and measure the occurrence of outcomes by 0 and 1: associate a
0 if the outcome does not occur and a 1 if the outcome occurs. We obtain the following four
possible assignments:
so the set of subsets of {H, T } can be represented as 4 sequences of length 2 formed with 0 and
5
6
1: {0, 0}, {0, 1}, {1, 0}, {1, 1}. These corresponds in order to Ø, {T }, {H}, {H, T }, which is
2{H,T } .
Example 1.2.2 Pick a natural number at random. Any subset of the sample space corresponds
to a sequence formed with 0 and 1. For instance, the subset {1, 3, 5, 6} corresponds to the
sequence 10101100000 . . . having 1 on the 1st, 3rd, 5th and 6th places and 0 in rest. It is known
that the number of these sequences is infinite and can be put into bijective correspondence with
the real numbers set R. This can be also written as |2N | = |R|.
Any subset F of 2Ω that satisfies the previous three properties is called a σ-field. The sets
belonging to F are called events. This way, the complement of an event, or the union of events
is also an event. We say that an event occurs if the outcome of the experiment is an element
of that subset.
1. P (Ω) = 1;
The triplet (Ω, F, P ) is called a probability space. This is the main setup in which the
probability theory works.
Example 1.3.1 In the case of flipping a coin, the probability space has the following elements:
Ω = {H, T }, F = {Ø, {H}, {T }, {H, T }} and P defined by P (Ø) = 0, P ({H}) = 21 , P ({T }) =
1
2 , P ({H, T }) = 1.
Example 1.3.2 Consider a finite sample space Ω = {s1 , . . . , sn }, with the σ-field F = 2Ω , and
probability given by P (A) = |A|/n, ∀A ∈ F. Then (Ω, 2Ω , P ) is called the classical probability
space.
7
¡ ¢
Figure 1.1: If any pullback X −1 (a, b) is known, then the random variable X : Ω → R is
2Ω -measurable.
Example 1.4.1 Consider the experiment of flipping three coins. In this case Ω is the set of
all possible triplets. Consider the random variable X which gives the number of tails obtained.
For instance X(HHH) = 0, X(HHT ) = 1, etc. The sets
Example 1.4.2 A graph is a set of elements, called nodes, and a set of unordered pairs of
nodes, called edges. Consider the set of nodes N = {n1 , n2 , . . . , nk } and the set of edges
E = {(n1 , n2 ), . . . , (ni , nj ), . . . , (nk−1 , nk )}. Define the probability space (Ω, F, P ), where
◦ the sample space is Ω = N ∪ E (the complete graph);
◦ the σ-field F is the set of all subgraphs of Ω;
8
◦ the probability is given by P (G) = n(G)/k, where n(G) is the number of nodes of the
graph G.
As an example of a random variable we consider Y : F → R, Y (G) = the total number of edges
of the graph G. Since given F, one can count the total number of edges of each subgraph, it
follows that Y is F-measurable, and hence it is a random variable.
If we have
d
F (x) = p(x),
dx X
then we say that p(x) is the probability density function of X. A useful property which follows
from the Fundamental Theorem of Calculus is
Z b
P (a < X < b) = P (ω; a < X(ω) < b) = p(x) dx.
a
In the case of discrete random variables the aforementioned integral is replaced by the following
sum X
P (a < X < b) = P (X = x).
a<x<b
E[X] = µ, V ar[X] = σ 2 .
Log-normal distribution Let X be normally distributed with mean µ and variance σ 2 . Then
the random variable Y = eX is said log-normal distributed. The mean and variance of Y are
given by
2
E[X] = eµ+σ /2
2 2
V ar[X] = e2µ+σ (eσ − 1).
9
0.4
0.5
0.3 0.4
0.3
0.2
0.2
0.1
0.1
-4 -2 0 2 4 0 2 4 6 8
a b
3.5
Α = 3, Β = 9 Α = 8, Β = 3
3.0
0.20
Α = 3, Β = 2 2.5
0.15
2.0
0.10 1.5
Α = 4, Β = 3
1.0
0.05
0.5
c d
Gamma distribution A random variable X is said to have a gamma distribution with pa-
rameters α > 0, β > 0 if its density function is given by
xα−1 e−x/β
p(x) = , x ≥ 0,
β α Γ(α)
where Γ(α, β) denotes the gamma function, see Fig.1.2c. The mean and variance are
The case α = 1 is known as the exponential distribution, see Fig.1.3a. In this case
1 −x/β
p(x) = e , x > 0.
β
10
0.35 0.10
0.25
0.20
Β=3 0.06
0.15
0.04
0.10
0.02
0.05
0 2 4 6 8 10 5 10 15 20 25 30
a b
The particular case when α = n/2 and β = 2 becomes the χ2 −distribution with n degrees of
freedom. This characterizes a sum of n independent standard normal distributions.
Beta distribution A random variable X is said to have a beta distribution with parameters
α > 0, β > 0 if its probability density function is of the form
xα−1 (1 − x)β−1
p(x) = , 0 ≤ x ≤ 1,
B(α, β)
where B(α, β) denotes the beta function. See see Fig.1.2d for two particular density functions.
In this case
α αβ
E[X] = , V ar[X] = 2
.
α+β (α + β) (α + β + 1)
Proposition 1.7.1 Let X and Y be independent random variables with probability density
functions pX (x) and pY (y). Then the product random variable XY has the probability density
function pX (x) pY (y).
11
Proof: Let pXY (x, y) be the probability density of the product XY . Using the independence
of sets we have
1.8 Expectation
A random variable X : Ω → R is called integrable if
Z Z
|X(ω)| dP (ω) = |x|p(x) dx < ∞.
Ω R
where p(x) denotes the probability density function of X. Customary the expectation of X is
denoted by µ and it is also called mean. In general, for any continuous1 function h : R → R,
we have Z Z
¡ ¢
E[h(X)] = h X(ω) dP (ω) = h(x)p(x) dx.
Ω R
Proposition 1.8.1 The expectation operator E is linear, i.e. for any integrable random vari-
ables X and Y
1. E[cX] = cE[X], ∀c ∈ R;
2. E[X + Y ] = E[X] + E[Y ].
Proof: It follows from the fact that the integral is a linear operator.
Proposition 1.8.2 Let X and Y be two independent integrable random variables. Then
E[XY ] = E[X]E[Y ].
Proof: This is a variant of Fubini’s theorem. Let pX , pY , pXY denote the probability densities
of X, Y and XY , respectively. Since X and Y are independent, by Proposition 1.7.1 we have
pXY = pX pY . Then
ZZ Z Z
E[XY ] = xypXY (x, y) dxdy = xpX (x) dx ypY (y) dy = E[X]E[Y ].
1 in general, measurable
12
Proposition 1.9.1 Consider the probability space (Ω, F, P ), and let G be a σ-field included in
F. It X is a G-predictable random variable such that
Z
X dP = 0 ∀A ∈ G,
A
then X = 0 a.s.
¡ ¢
Proof: In order to show that X = 0 almost surely, it suffices to prove that P ω; X(ω) = 0 = 1.
We shall show first that X takes values as small as possible with probability one, i.e. ∀² > 0
we have P (|X| < ²) = 1. To do this, let A = {ω; X(ω) ≥ ²}. Then
Z Z Z
1 1
0 ≤ P (X ≥ ²) = dP = ² dP ≤ X dP = 0,
A ² A ² A
and hence P (X ≥ ²) = 0. Similarly P (X ≤ −²) = 0. Therefore
then X = Y a.s.
R
Proof: Since A (X − Y ) dP = 0, ∀A ∈ G, by Proposition 1.9.1 we have X − Y = 0 a.s.
X dP = E[X]dP.
Ø Ø
Exercise 1.10.2 Show that E[E[X|G]] = E[X], i.e. all conditional expectations have the same
mean, which is the mean of X.
Proof: Using the definition of expectation and taking A = Ω in the second relation of the
aforementioned definition, yields
Z Z
E[E[X|G]] = E[X|G] dP = XdP = E[X],
Ω Ω
14
Exercise 1.10.3 The conditional expectation of X given the total information F is the random
variable X itself, i.e.
E[X|F] = X.
Proof: The random variables X and E[X|F] are both F-predictable (from the definition of
the random variable). From the definition of the conditional expectation we have
Z Z
E[X|F] dP = X dP, ∀A ∈ F.
A A
Figure 1.4: Jensen’s inequality ϕ(E[X]) < E[ϕ(X)] for a convex function ϕ.
E[X]2 ≤ E[X 2 ].
Since the right side is finite, it follows that E[X] < ∞, so X is integrable.
16
Application 1.11.3 If mX (t) denotes the moment generating function of the random variable
X with mean µ, then
mX (t) ≥ etµ .
Proof: Applying Jensen inequality with the convex function ϕ(x) = ex yields
eE[X] ≤ E[eX ].
Exercise 1.11.4 Prove that a non-constant random variable has a non-zero standard devia-
tion.
1
P (ω; |X(ω)| ≥ λ) ≤ E[|X|p ],
λp
for any p > 0.
Theorem 1.11.8 (Chernoff bounds) Let X be a random variable. Then for any λ > 0 we
have
E[etX ]
1. P (X ≥ λ) ≤ , ∀t > 0;
eλt
E[etX ]
2. P (X ≤ λ) ≤ , ∀t < 0.
eλt
Proof: 1. Let t > 0 and denote Yt = etX . By Markov’s inequality
E[Y ]
P (Y ≥ eλt ) ≤ .
eλt
Then we have
It is easy to see that the quadratic function f (t) = (µ − λ)t + 21 t2 σ 2 has the minimum value
λ−µ
reached for t =
σ2 ³λ − µ´ (λ − µ)2
min f (t) = f = − .
t>0 σ2 2σ 2
Substituting in the previous formula, we obtain the following result:
18
³X
n ´³ X
n ´ Xn
λi a i λi b i ≤ λi ai bi .
i=1 i=1 i=1
Proof: Since the sequences (ai ) and (bi ) are either both increasing or both decreasing
(ai − aj )(bi − bj ) ≥ 0.
Multiplying by the positive quantity λi λj and summing over i and j we get
X
λi λj (ai − aj )(bi − bj ) ≥ 0.
i,j
Expanding yields
³X ´³ X ´ ³X ´³ X ´ ³X ´³ X ´
λj λi ai bi − λi ai λj bj − λj a j λi b i
j i i j j i
³X ´³ X ´
+ λi λj aj bj ≥ 0.
i j
X
Using λj = 1 the expression becomes
j
X ³X ´³ X ´
λi ai bi ≥ λi ai λj bj ,
i i j
Proposition 1.11.11 Let X be a random variable and f and g be two functions, both increas-
ing or both decreasing. Then
Let x0 < x1 < · · · < xn be a partition of the interval I, with ∆x = xk+1 − xk . Using Lemma
1.11.10 we obtain the following inequality between Riemann sums
X ³X ´³ X ´
f (xj )g(xj )p(xj )∆x ≥ f (xj )p(xj )∆x g(xj )p(xj )∆x ,
j j j
where aj = f (xj ), bj = g(xj ), and λj = p(xj )∆x. Taking the limit k∆xk → 0 we obtain
(1.11.4), which leads to the desired result.
Exercise 1.11.12 Let f and g be two differentiable functions such that f 0 (x)g 0 (x) > 0, x ∈ I.
Then
E[f (X)g(X)] ≥ E[f (X)]E[g(X)],
for any random variable X with values in I.
and we shall write ac-lim Xn = X. An important example where this type of limit occurs is
n→∞
the Strong Law of Large Numbers:
If Xn is a sequence of independent and identically distributed random variables with the
X1 + · · · + Xn
same mean µ, then ac-lim = µ.
n→∞ n
It worth noting that this type of convergence is known also under the name of strong
convergence. This is the reason why the the aforementioned theorem bares its name.
This limit will be abbreviated by ms-lim Xn = X. The mean square convergence is useful
n→∞
when defining the Ito integral.
Example 1.12.1 Consider a sequence Xn of random variables such that there is a constant k
with E[Xn ] → k and V ar(Xn ) → 0 as n → ∞. Show that ms-lim Xn = k.
n→∞
Proof: Let ms-lim Yn = Y . Let ² > 0 arbitrary fixed. Applying Markov’s inequality with
n→∞
X = Yn − Y , p = 2 and λ = ², yields
1
0 ≤ P (|Yn − Y | ≥ ²) ≤ E[|Yn − Y |2 ].
²2
The right side tends to 0 as n → ∞. Applying the Squeeze Theorem we obtain
lim P (|Yn − Y | ≥ ²) = 0,
n→∞
Remark 1.12.3 The conclusion still holds true even in the case when there is a p > 0 such
that E[|Xn |p ] → 0 as n → ∞.
22
Limit in Distribution
We say the sequence Xn converges in distribution to X if for any continuous bounded function
ϕ(x) we have
lim ϕ(Xn ) = ϕ(X).
n→∞
This type of limit is even weaker than the stochastic convergence, i.e. it is implied by it.
An application of the limit in distribution is obtained if consider ϕ(x) = eitx . In this
case, if Xn converges in distribution to X, then the characteristic function of Xn converges to
the characteristic function of X. In particular, the probability density of Xn approachers the
probability density of X.
It can be shown that the convergence in distribution is equivalent with
1. ms-lim (Xn + Yn ) = 0
n→∞
2. ms-lim (Xn Yn ) = 0.
n→∞
Proof: Since ms-lim Xn = 0, then lim E[Xn2 ] = 0. Applying the Squeeze Theorem to the
n→∞ n→∞
inequality2
0 ≤ E[Xn ]2 ≤ E[Xn2 ]
yields lim E[Xn ] = 0. Then
n→∞
³ ´
lim V ar[Xn ] = lim E[Xn2 ] − lim E[Xn ]2
n→∞ n→∞ n→∞
= lim E[Xn2 ] − lim E[Xn ]2
n→∞ n→∞
= 0.
Similarly, we have lim E[Yn2 ] = 0, lim E[Yn ] = 0 and lim V ar[Yn ] = 0. Then lim σXn =
n→∞ n→∞ n→∞ n→∞
lim σYn = 0. Using the correlation formula
n→∞
Cov(Xn , Yn )
Corr(Xn , Yn ) = ,
σXn σYn
Since lim σXn σXn = 0, from the Squeeze Theorem it follows that
n→∞
lim Cov(Xn , Yn ) = 0.
n→∞
Proposition 1.13.2 If the sequences of random variables Xn and Yn converge in the mean
square, then
Proof: 1. Let ms-lim Xn = L and ms-lim Yn = M . Consider the sequences Xn0 = Xn − L and
n→∞ n→∞
Yn0 = Yn − M . Then ms-lim Xn0 = 0 and ms-lim Yn0 = 0. Applying Lemma 1.13.1 yields
n→∞ n→∞
ms-lim (Xn0 + Yn0 ) = 0.
n→∞
This is equivalent with
ms-lim (Xn − L + Yn − M ) = 0,
n→∞
which becomes
ms-lim (Xn + Yn ) = L + M .
n→∞
The evolution in time of a given state of the world ω ∈ Ω given by the function t 7−→ Xt (ω) is
called a path or realization of Xt . The study of stochastic processes using computer simulations
is based on retrieving information about the process Xt given a large number of it realizations.
24
Consider that all the information accumulated until time t is contained by the σ-field Ft .
This means that Ft contains the information of which events have already occurred until the
time t, and which did not. Since the information is growing in time, we have
Fs ⊂ F t ⊂ F
Xt = E[X|Ft ].
From the definition of the conditional expectation, the random variable Xt is Ft -predictable, and
can be regarded as the measurement of X at time t using the information Ft . If the accumulated
knowledge Ft increases and eventually equals the σ-field F, then X = E[X|F], i.e. we obtain
the entire random variable. The process Xt is adapted to Ft .
Example 1.14.3 Don Joe is asking the doctor how long he still has to live. The age at which
he will pass away is a random variable, denoted by X. Given his medical condition today, which
is contained in Ft , the doctor infers that Mr. Joe will die at the age of Xt = E[X|Ft ]. The
stochastic process Xt is adapted to the medical knowledge Ft .
Remark 1.14.5 The first condition states that the unconditional forecast is finite E[|Xt ]] =
Z
|Xt | dP < ∞. Condition 2 says that the value Xt is known, given the information set Ft .
Ω
The third relation asserts that the best forecast of unobserved future values is the last observation
on Xt .
Example 1.14.1 Let Xt denote Mr. Li Zhu’s salary after t years of work in the same company.
Since Xt is known at time t and it is bounded above, as all salaries are, then the first two
conditions hold. Being honest, Mr. Zhu expects today that his future salary will be the same as
today’s, i.e. Xs = E[Xt |Fs ], for s < t. This means that Xt is a martingale.
Example 1.14.2 If in the previous example Mr. Zhu is optimistic and believes as today that
his future salary will increase, then Xt is a submartingale.
Example 1.14.4 Let Xt and Yt be martingales with respect to the filtration Ft . Show that for
any a, b, c ∈ R the process Zt = aXt + bYt + c is a Ft -martingale.
Example 1.14.5 Let Xt and Yt be martingales with respect to the filtration Ft . Is the process
Xt Yt a martingale with respect to Ft ?
Bt − Bs ∼ N (0, |t − s|).
The process Xt = x + Bt has all the properties of a Brownian motion that starts at x. Since
Bt − Bs is stationary, its distribution function depends only on the time interval t − s, i.e.
From condition 4 we get that Bt is normally distributed with mean E[Bt ] = 0 and V ar[Bt ] = t
Bt ∼ N (0, t).
Let 0 < s < t. Since the increments are independent, we can write
Proposition 2.1.2 A Brownian motion process Bt is a martingale with respect to the infor-
mation set Ft = σ(Bs ; s ≤ t).
27
28
where we used that Bs is Fs -measurable (from where E[Bs |Fs ] = Bs ) and that the increment
Bt − Bs is independent of previous values of Bt contained in the information set Ft = σ(Bs ; s ≤
t).
A process with similar properties as the Brownian motion was introduced by Wiener.
E[(Wt − Ws )2 ] = t − s, s ≤ t;
The only property Bt has and Wt seems not to have is that the increments are normally
distributed. However, there is no distinction between these two processes, as the following
result states.
In stochastic calculus we often need to use infinitesimal notations and its properties. If dWt
denotes the infinitesimal increment of a Wiener process in the time interval dt, the aforemen-
tioned properties become dWt ∼ N (0, dt), E[dWt ] = 0, and E[(dWt )2 ] = dt.
Proposition 2.1.5 If Wt is a Wiener process with respect to the information set Ft , then
Yt = Wt2 − t is a martingale.
Proof: Let s < t. Using that the increments Wt − Ws and (Wt − Ws )2 are independent of the
information set Fs and applying Proposition 1.10.4 yields
Proposition 2.1.6 The conditional distribution of Wt+s , given the present Wt and the past
Wu , 0 ≤ u < t, depends only on the present.
1 These type of processes are called Marcov processes
29
Since Wt is normally distributed with mean 0 and variance t, its density function is
1 − x2
φt (x) = √ e 2t .
2πt
Then its distribution function is
Z x
1 u2
Ft (x) = P (Wt ≤ x) = √ e− 2t du
2πt −∞
Cov(Ws , Wt ) = Cov(Ws , Ws + Wt − Ws )
= Cov(Ws , Ws ) + Cov(Ws , Wt − Ws )
= V ar(Ws ) + E[Ws (Wt − Ws )] − E[Ws ]E[Wt − Ws ]
= s + E[Ws ]E[Wt − Ws ]
= s,
since E[Ws ] = 0.
We can arrive to the same result starting from the formula
so Cov(Ws , Wt ) = s.
30
r
Cov(Ws , Wt ) s s
Corr(Ws , Wt ) = =√ √ = .
σ(Wt )σ(Ws ) s t t
Remark 2.1.8 Removing the order relation between s and t, the previous relations can be also
stated as
Cov(Ws , Wt ) = min{s, t};
s
min{s, t}
Corr(Ws , Wt ) = .
max{s, t}
The following exercises state the translation and the scaling invariance of the Brownian
motion.
Exercise 2.1.9 For any t0 ≥ 0, show that the process Xt = Wt+t0 −Wt0 is a Brownian motion.
Exercise 2.1.10 For any λ > 0, show that the process Xt = √1 Wλt is a Brownian motion.
λ
Exercise 2.1.11 Let 0 < s < t < u. Show the following multiplicative property
Corr(Ws , Wt )Corr(Wt , Wu ) = Corr(Ws , Wu ).
Research topic: Find all stochastic processes with the aforementioned property.
Exercise 2.1.12 (a) Use the martingale property of Wt2 − t to find
E[(Wt2 − t)(Ws2 − s)];
(b) Evaluate E[Wt2 Ws2 ];
(c) Compute Cov(Wt2 , Ws2 );
(d) Find Corr(Wt2 , Ws2 ).
Exercise 2.1.13 Consider the process Yt = tW 1t , t > 0.
(a) Find the distribution of Yt ;
(b) Find Cov(Ys , Yt );
(c) Is Yt a Brownian motion process?
Exercise 2.1.14 The process Xt = |Wt | is called Brownian motion reflected at the origin.
Show that p
(a) E[|Wt |] = 2t/π;
(b) V ar(|Wt |) = (1 − π2 )t.
Research topic: Find all functions g(t) and h(t) such that the process Xt = g(t)Wh(t) is a
Brownian motion.
Exercise 2.1.15 Let 0 < s < t. Find E[Wt2 |Fs ].
Exercise 2.1.16 Let 0 < s < t. Show that
(a) E[Wt3 |Fs ] = 3(t − s)Ws + Ws3 ;
(b) E[Wt4 |Fs ] = 3(t − s)2 + 6(t − s)Ws2 + Ws4 .
31
3
-0.5
-1.0
a b
Figure 2.1: a Three simulations of the Brownian motion process Wt ; b Two simulations of the
geometric Brownian motion process eWt .
Proposition 2.2.2 The geometric Brownian motion Xt = eWt is log-normally distributed with
mean et/2 and variance e2t − et .
Proof: Since Wt is normally distributed, then Xt = eWt will have a log-normal distribution.
Using Lemma 2.2.1 we have
1 2
√
e−(ln x) /(2t) , if x > 0,
d x 2πt
p(x) = F (x) =
dx Xt
0, elsewhere.
Exercise 2.2.3 If Xt = eWt , find the covariance Cov(Xs , Xt ).
Exercise 2.2.4 Let Xt = eWt .
(a) Show that Xt is not a martingale.
t
(b) Show that e− 2 Xt is a martingale.
Theorem 2.3.1 If Xj are independent random variables normally distributed with mean µj
and variance σj2 , then the sum X1 +· · ·+Xn is also normally distributed with mean µ1 +· · ·+µn
and variance σ12 + · · · + σn2 .
Then
³ ´ ³ n(n + 1)(2n + 1) ´
X1 + · · · + Xn ∼ N 0, (1 + 22 + 32 + · · · + n2 )∆s = N 0, ∆s ,
6
t
with ∆s = . Using (2.3.1) yields
n
W s + · · · + W sn ³ (n + 1)(2n + 1) ´
t 1 ∼ N 0, t3 .
n 6n2
Taking the limit we get
³ t3 ´
Zt ∼ N 0, .
3
Proposition 2.3.2 The integrated Brownian motion Zt has a Gaussian distribution with mean
0 and variance t3 /3.
The mean and the variance can be also computed in a direct way as follows. By Fubini’s
theorem we have
Z t Z Z t
E[Zt ] = E[ Ws ds] = Ws ds dP
0 R 0
Z tZ Z t
= Ws dP ds = E[Ws ] ds = 0,
0 R 0
where
D1 = {(u, v); u < v, 0 ≤ u ≤ t}, D2 = {(u, v); u > v, 0 ≤ u ≤ t}
The first integral can be evaluated using Fubini’s theorem
ZZ ZZ
min{u, v} dudv = v dudv
D1 D1
Z t ³Z u ´ Z t
u2 t3
= v dv du = du = .
0 0 0 2 6
Similarly, the latter integral is equal to
ZZ
t3
min{u, v} dudv = .
D2 6
34
(b) Use the first part to find the mean and variance of Zt .
Exercise 2.3.4 Let s < t. Show that the covariance of the integrated Brownian motion is given
by ³t s´
Cov[Zs , Zt ] = s2 − .
2 6
Exercise 2.3.5 Show that
(a) Cov[Zt , Zt − Zt−h ] = 21 t2 h + o(h).
t2
(b) Cov[Zt , Wt ] = .
2
Z t
Exercise 2.3.6 Consider the process Xt = eWs ds.
0
(a) Find the mean of Xt ;
(b) Find the variance of Xt ;
(c) What is the distribution of Xt ?
Vt = eZt
2t3 t3
V ar[Vt ] = E[Vt2 ] − E[Vt ]2 = e 3 −e3 .
using that the increments Wt − W0 and W1 − Wt are independent and normally distributed,
with
Wt − W0 ∼ N (0, t), W1 − Wt ∼ N (0, 1 − t),
it follows that Xt is normally distributed with
This can be also stated by saying that the Brownian bridge tied at 0 and 1 is a Gaussian process
with mean 0 and variance t(1 − t).
E[Yt ] = µt + E[Wt ] = µt
and variance
V ar[Yt ] = V ar[µt + Wt ] = V ar[Wt ] = t.
Proof: Since the Brownian motions W1 (t), . . . , Wn (t) are independent, their joint density
function is
In the next computation we shall use the following formula of integration that follows from
the use of polar coordinates
Z Z ρ
n−1
f (x) dx = σ(S ) rn−1 g(r) dr, (2.7.3)
{|x|≤ρ} 0
2π n/2
σ(Sn−1 ) =
Γ(n/2)
Differentiating yields
d σ(Sn−1 ) n−1 − ρ2
pt (ρ) = FR (ρ) = ρ e 2t
dρ (2πt)n/2
2 2
n−1 − ρ2t
= ρ e , ρ > 0, t > 0.
(2t)n/2 Γ(n/2)
It worth noting that in the 2-dimensional case the aforementioned density becomes a partic-
ular case of a Weibull distribution with parameters m = 2 and α = 2t, called Wald’s distribution
1 − x2
pt (x) = xe 2t , x > 0, t > 0.
t
37
where o(h) denotes a quantity such that limh→0 o(h)/h = 0. This means the probability that a
jump of size 1 occurs in the infinitesimal interval dt is equal to λdt, and the probability that at
least 2 events occur in the same small interval is zero. This implies that the random variable
dNt may take only two values, 0 and 1, and hence satisfies
P (dNt = 1) = λ dt (2.8.6)
P (dNt = 0) = 1 − λ dt. (2.8.7)
In particular, the random variable Nt is Poisson distributed with E[Nt ] = λt and V ar[Nt ] = λt.
The parameter λ is called the rate of the process. This means that the events occur at the
constant rate λ.
38
where we used that Ns is Fs -measurable (and hence E[Ns |Fs ] = Ns ) and that the increment
Nt − Ns is independent of previous values of Ns and the information set Fs . Subtracting λt
yields
E[Nt − λt|Fs ] = Ns − λs,
or E[Mt |Fs ] = Ms . Since it is obvious that Mt is Ft -adapted, it follows that Mt is a martingale.
It worth noting that the Poisson process Nt is not a martingale. The martingale process
Mt = Nt − λt is called the compensated Poisson process.
Exercise 2.8.5 (i) Show that the moment generating function of the random variable Nt is
x
mNt (x) = eλt(e −1)
.
Exercise 2.8.6 Find the mean and variance of the process Xt = eNt .
Exercise 2.8.7 (i) Show that the moment generating function of the random variable Mt is
x
mMt (x) = eλt(e −x−1)
.
E[Mt − Ms ] = 0,
E[(Mt − Ms )2 ] = λ(t − s),
E[(Mt − Ms )3 ] = λ(t − s),
E[(Mt − Ms )4 ] = λ(t − s) + 3λ2 (t − s)2 .
Proposition 2.8.9 The random variables Tn are independent and exponentially distributed
with mean E[Tn ] = 1/λ.
Proof: We start by noticing that the events {T1 > t} and {Nt = 0} are the same, since both
describe the situation that no events occurred until time t. Then
d
fT1 (t) = FT (t) = λe−λt .
dt 1
40
It follows that T1 is has an exponential distribution, with E[T1 ] = 1/λ. The conditional
distribution of T2 is
d λe−λt (λt)n−1
fSn (t) = FSn (t) = .
dt (n − 1)!
Writing
tn−1 e−λt
fSn (t) = ,
(1/λ)n Γ(n)
it turns out that Sn has a gamma distribution with parameters α = n and β = 1/λ. It follows
that
n n
E[Sn ] = , V ar[Sn ] = 2 .
λ λ
The relation lim E[Sn ] = ∞ states that the expectation of the waiting time gets unbounded
n→∞
large as n → ∞.
Figure 2.2: The Poisson process Nt and the waiting times S1 , S2 , · · · Sn . The shaded rectangle
has area n(Sn+1 − t).
called the integrated Poisson process. The next result provides a relation between the process
Ut and the partial sum of the waiting times Sk .
Let Nt = n. Since Nt is equal to k between the waiting times Sk and Sk+1 , the process Ut ,
which is equal to the area of the subgraph of Nu between 0 and t, can be expressed as
Z t
Ut = Nu du = 1 · (S2 − S1 ) + 2 · (S3 − S2 ) + · · · + n(Sn+1 − Sn ) − n(Sn+1 − t).
0
Since Sn < t < Sn+1 , the difference of the last two terms represents the area of last the
rectangle, which has the length t − Sn and the height n. Using associativity, a computation
yields
where we replaced n by Nt .
42
The following result deals with the quadratic variation of the compensated Poisson process
Mt = Nt − λt.
Proposition 2.8.12 Let a < b and consider the partition a = t0 < t1 < · · · < tn−1 < tn < b.
Then
n−1
X
ms– lim (Mtk+1 − Mtk )2 = Nb − Na , (2.8.9)
k∆n k→∞
k=0
Proof: For the sake of simplicity we shall use the following notations
∆tk = tk+1 − tk , ∆Mk = Mtk+1 − Mtk , ∆Nk = Ntk+1 − Ntk .
The relation we need to prove can be also written as
n−1
X £ ¤
ms-lim (∆Mk )2 − ∆Nk = 0.
n→∞
k=0
Let
Yk = (∆Mk )2 − ∆Nk = (∆Mk )2 + ∆Mk + λ∆tk .
It suffices to show that
h n−1
X i
E Yk = 0, (2.8.10)
k=0
h n−1
X i
lim V ar Yk = 0. (2.8.11)
n→∞
k=0
The first identity follows from the properties of Poisson processes, see Exercise 2.8.7
h n−1
X i n−1
X n−1
X
E Yk = E[Yk ] = E[(∆Mk )2 ] − E[∆Nk ]
k=0 k=0 k=0
n−1
X
= (λ∆tk − λ∆tk ) = 0.
k=0
For the proof of the identity (2.8.11) we need to find first the variance of Yk .
V ar[Yk ] = V ar[(∆Mk )2 − (∆Mk + λ∆tk )] = V ar[(∆Mk )2 − ∆Mk ]
= V ar[(∆Mk )2 ] + V ar[∆Mk ] − 2Cov[∆Mk2 , ∆Mk ]
= λ∆tk + 2λ2 ∆t2k + λ∆tk
h i
−2 E[(∆Mk )3 ] − E[(∆Mk )2 ]E[∆Mk ]
= 2λ2 (∆tk )2 ,
43
where we used Exercise 2.8.7 and the fact that E[∆Mk ] = 0. Since Mt is a process with
independent increments, then Cov[Yk , Yj ] = 0 for i 6= j. Then
h n−1
X i n−1
X X n−1
X
V ar Yk = V ar[Yk ] + 2 Cov[Yk , Yj ] = V ar[Yk ]
k=0 k=0 k6=j k=0
n−1
X n−1
X
= 2λ2 (∆tk )2 ≤ 2λ2 k∆n k ∆tk = 2λ2 (b − a)k∆n k,
k=0 k=0
h n−1
X i
and hence V ar Yn → 0 as k∆n k → 0. By Proposition 3.3.1 we obtain the desired limit in
k=0
mean square sense.
The previous result states that the quadratic variation of the martingale Mt between a and
b is equal to the jump of the Poisson process between a and b.
while the left side can be regarded as a stochastic integral with respect to (dMt )2
Z b n−1
X
(dMt )2 := ms- lim (Mtk+1 − Mtk )2 .
a n→∞
k=0
since this is thought as a vanishing integral of the increment process dMt with respect to dt
Z b
dMt dt = 0, ∀a, b ∈ R.
a
Denote
n−1
X n−1
X
Xn = (tk+1 − tk )(Mtk+1 − Mtk ) = ∆tk ∆Mk .
k=0 k=0
1. E[Xn ] = 0;
2. lim V ar[Xn ] = 0.
n→∞
h n−1
X i n−1
X
E[Xn ] = E ∆tk ∆Mk = ∆tk E[∆Mk ] = 0.
k=0 k=0
Since the Poisson process Nt has independent increments, the same property holds for the
compensated Poisson process Mt . Then ∆tk ∆Mk and ∆tj ∆Mj are independent for k 6= j, and
using the properties of variance we have
h n−1
X i n−1
X n−1
X
V ar[Xn ] = V ar ∆tk ∆Mk = (∆tk )2 V ar[∆Mk ] = λ (∆tk )3 ,
k=0 k=0 k=0
where we used
V ar[∆Mk ] = E[(∆Mk )2 ] − (E[∆Mk ])2 = λ∆tk ,
see Exercise 2.8.7, (ii). If let k∆n k = max ∆tk , then
k
n−1
X n−1
X
V ar[Xn ] = λ (∆tk )3 ≤ λk∆n k2 ∆tk = λ(b − a)k∆n k2 → 0
k=0 k=0
dt dMt = 0. (2.8.15)
Since the Brownian motion Wt and the process Mt have independent increments and ∆Wk is
independent of ∆Mk , we have
n−1
X n−1
X
E[Yn ] = E[∆Wk ∆Mk ] = E[∆Wk ]E[∆Mk ] = 0,
k=0 k=0
where we used E[∆Wk ] = E[∆Mk ] = 0. Using also E[(∆Wk )2 ] = ∆tk , E[(∆Mk )2 ] = λ∆tk ,
invoking the independence of ∆Wk and ∆Mk , we get
as n → ∞. Since Yn is a random variable with mean zero and variance decreasing to zero, it
follows that Yn → 0 in the mean square sense. Hence we proved that
The relations proved in this section will be useful in the sequel, when developing the stochas-
tic model of a stock price that exhibits jumps modeled by a Poisson process.
46
Chapter 3
Properties of Stochastic
Processes
The first result deals with the hitting time for a Brownian motion to reach the barrier a ∈ R,
see Fig.3.1.
Lemma 3.1.1 Let Ta be the first time the Brownian motion Wt hits a. Then
Z ∞
2 2
P (Ta ≤ t) = √ √ e−y /2
dy.
2π |a|/ t
2.5
Wt
2.0
a
1.5
Ta
47
48
P (A) = P (A ∩ B) + P (A ∩ B)
= P (A|B)P (B) + P (A|B)P (B). (3.1.1)
Let a > 0. Using formula (3.1.1) for A = {ω; Wt (ω) ≥ a} and B = {ω; Ta (ω) ≤ t} yields
If Ta > t, the Brownian motion did not reached the barrier a yet, so we must have Wt < a.
Therefore
P (Wt ≥ a|Ta > t) = 0.
If Ta ≤ t, then WTa = a. Since the Brownian motion is a Markov process, it starts fresh at Ta .
Due to symmetry of the density function of a normal variable, Wt has equally chances to go up
or go down after the time interval t − Ta . It follows that
1
P (Wt ≥ a|Ta ≤ t) = .
2
Substituting in (3.1.2) yields
P (Ta ≤ t) = 2P (Wt ≥ a)
Z ∞ Z ∞
2 −x2 /(2t) 2 −y 2 /2
= √ e dx = √ √ e dy.
2πt a 2π a/ t
If a < 0, symmetry reasons imply that the distribution of Ta is the same as that of T−a , so we
get Z ∞
2 −y 2 /2
P (Ta ≤ t) = P (T−a ≤ t) = √ √ e dy.
2π −a/ t
Theorem 3.1.2 Let a ∈ R be fixed. Then the Brownian motion hits a in finite time with
probability 1.
Remark 3.1.3 Even if the hitting time is finite with probability 1, its expectation E[Ta ] is
infinite. This means that the expected time to hit the barrier is infinite.
Corollary 3.1.4 A Brownian motion process returns to the origin in finite time with probability
1.
Exercise 3.1.5 Show that the distribution function of the process Xt = max Ws is given by
s∈[0,t]
Z √
a/ t
2 2
P (Xt ≤ a) = √ e−y /2
dy.
2π −∞
The fact that a Brownian motion returns or hits a barrier almost surely is a property char-
acteristic to the dimension 1 only. The next result states that in larger dimensions this is no
more possible.
¡ ¢
Theorem 3.1.6 Let (a, b) ∈ R2 . The 2-dimensional Brownian motion W (t) = W1 (t), W2 (t)
hits the point (a, b) with probability zero. The same result is valid for any n-dimensional Brow-
nian motion, with n ≥ 2.
Research topic. This deals with the hitting time of a 2-dimensional Brownian motion to reach
a given disk. Let D² (x0 ) = {x ∈ R2 ; |x − x0 | ≤ ²}. Find the probability P (∃t; W (t) ∈ D² (x0 ))
as a function of x0 and ². Let TD² (x0 ) = inf t≥0 {t; W (t) ∈ D² (x0 )}. Find the distribution
and the expectation of the random variable TD² (x0 ) . It is known that P (TD² (x0 ) < ∞) = 1.
However, in the n-dimensional version this probability is zero.
Research topic. This deals with the exit time of a 2-dimensional Brownian motion from the
disc of radius a. Let
Find the distribution function of the exit time Ta and its expectation. It is known that P (Ta <
2
∞) = R2 . You may try to apply the same argument as in the proof of Lemma 3.1.1; however,
because of convexity
1
P (Rt ≥ a|Ta ≤ t) > .
2
What is the exact value of this probability?
Theorem 3.1.7 (The law of Arc-sine) The probability that a Brownian motion Wt does not
have any zeros in the interval (t1 , t2 ) is equal to
r
2 t1
P (Wt 6= 0, t1 ≤ t ≤ t2 ) = arcsin .
π t2
Proof: Let A(a; t1 , t2 ) denote the event that the Brownian motion Wt takes on the value a
between t1 and t2 . In particular, A(0; t1 , t2 ) denotes the event that Wt has (at least) a zero
between t1 and t2 . Substituting A = A(0; t1 , t2 ) and X = Wt1 in the following formula of
conditional probability
Z Z
P (A) = P (A|X = x) PX (x) = P (A|X = x)fX (x) dx
50
0.4
0.2
t1 t2
50 100 150 200 250 300 350
-0.2
a
-0.4
Wt
-0.6
yields
Z
¡ ¢ ¡ ¢
P A(0; t1 , t2 ) = P A(0; t1 , t2 )|Wt1 = x fWt (x) dx (3.1.3)
1
Z ∞
1 ¡ ¢ x2
= √ P A(0; t1 , t2 )|Wt1 = x e− 2t1 dx
2πt1 −∞
Using the properties of Wt with respect to time translation and symmetry we have
¡ ¢ ¡ ¢
P A(0; t1 , t2 )|Wt1 = x = P A(0; 0, t2 − t1 )|W0 = x
¡ ¢
= P A(−x; 0, t2 − t1 )|W0 = 0
¡ ¢
= P A(|x|; 0, t2 − t1 )|W0 = 0
¡ ¢
= P A(|x|; 0, t2 − t1 )
¡ ¢
= P T|x| ≤ t2 − t1 ,
the last identity stating that Wt hits |x| before t2 − t1 . Using Lemma 3.1.1 yields
Z ∞
¡ ¢ 2 − y2
P A(0; t1 , t2 )|Wt1 = x = p e 2(t2 −t1 ) dy.
2π(t2 − t1 ) |x|
Substituting in (3.1.3) we obtain
Z ∞ ³ Z ∞ ´ x2
¡ ¢ 1 2 − y2
P A(0; t1 , t2 ) = √ p e 2(t2 −t1 ) dy e− 2t1 dx
2πt1−∞ 2π(t2 − t1 ) |x|
Z ∞Z ∞
1 − y2
−x
2
= p e 2(t2 −t1 ) 2t1 dydx.
π t1 (t2 − t1 ) 0 |x|
Exercise 3.1.10 Find the probability that a Brownian motion Wt does not take the value a in
the interval (t1 , t2 ).
Exercise 3.1.11 Let a 6= b. Find the probability that a Brownian motion Wt does not take any
of the values {a, b} in the interval (t1 , t2 ). Formulate and prove a generalization.
Research topic. What is the probability that a 2-dimensional Brownian process hits a set D
between time instances t1 and t2 ?
We provide below a similar result without proof.
Rt
Theorem 3.1.12 (Arc-sine Law of Lévy) Let L+ +
t = 0 sgn Ws ds be the amount of time a
Brownian motion Wt is positive during the time interval [0, t]. Then
r
2 τ
P (L+
t ≤ τ ) = arcsin .
π t
Research topic. Let Lt (D) be the amount of time spent by a 2-dimensional Brownian motion
W (t) inside the set D. Find P (Lt (D) ≤ τ ). When D is the half-plane {(x, y); y > 0} we retrieve
the previous result.
Theorem 3.1.13 Let Xt = µt + Wt denote a Brownian motion with nonzero drift rate µ, and
consider α, β > 0. Then
e2µβ − 1
P (Xt goes up to α before down to − β) = .
e2µβ− e−2µα
Limit in Distribution
We say that Xt converges in distribution to X if for any continuous bounded function ϕ(x) we
have
lim ϕ(Xt ) = ϕ(X).
t→∞
It worth noting that the stochastic convergence implies the convergence in distribution.
Proposition 3.3.1 Consider a stochastic process Xt such that E[Xt ] → k, constant, and
V ar(Xt ) → 0 as t → ∞. Then ms-lim Xt = k.
t→∞
Next we shall provide a few applications that show how some processes compare with powers
of t for t large.
Wt
Corollary 3.3.3 We have ms-lim = 0.
t→∞ t
53
Rt
Application 3.3.4 Let Zt = 0
Ws ds. If β > 3/2, then
Zt
ms-lim = 0.
t→∞ tβ
Zt E[Zt ] 1 t3 1
Proof: Let Xt = β
. Then E[Xt ] = β
= 0, and V ar[Xt ] = 2β V ar[Zt ] = 2β = 2β−3 ,
t t t 3t 3t
1
for any t > 0. Since 2β−3 → 0 as t → ∞, applying Proposition 3.3.1 leads to the desired
3t
result.
Remark 3.3.7 The strongest result regarding limits of Brownian motion is called the law of
iterated logarithms and was first proved by Lamperti:
Wt
lim sup √ = 1,
t→∞ 2t ln ln t
almost certainly.
Another convergence result can be obtained if consider the continuous analog of Exercise
1.12.2:
Proposition 3.3.10 Let Xt be a stochastic process such that there is a p > 0 such that
E[|Xt |p ] → 0 as t → ∞. Then st-lim Xt = 0.
t→∞
The following result can be regarded as the L’Hospital’s rule for sequences:
Lemma 3.3.12 (Cesaró-Stoltz) Let xn and yn be two sequences of real numbers, n ≥ 1. If
xn+1 − xn
the limit lim exists and it is equal to L, then the following limit exists
n→∞ yn+1 − yn
xn
lim = L.
n→∞ yn
Proof: (Sketch) Assume there are differentiable functions f and g such that f (n) = xn and
g(n) = yn . (How do you construct these functions?) From Cauchy’s theorem1 there is a
cn ∈ (n, n + 1) such that
xn+1 − xn f (n + 1) − f (n) f 0 (cn )
L = lim = lim = lim 0 .
n→∞ yn+1 − yn n→∞ g(n + 1) − g(n) n→∞ g (cn )
(Here one may argue against this, but we recall the freedom of choice for the functions f and
g such that cn can be any number between n and n + 1). By l’Hospital’s rule we get
f (t)
lim = L.
t→∞ g(t)
1 This says that if f and g are differentiable on (a, b) and continuous on [a, b], then there is a c ∈ (a, b) such
f (a) − f (b) f 0 (c)
that = 0 .
g(a) − g(b) g (c)
55
xn
Making t = n yields lim = L.
t→∞ yn
The next application states that if a sequence is convergent, then the arithmetic average of
its terms is also convergent, and the sequences have the same limit.
Example 3.3.1 Let an be a convergent sequence with lim an = L. Let
n→∞
a1 + a2 + · · · + an
An =
n
be the arithmetic average of the first n terms. Show that An is convergent and
lim An = L.
n→∞
Gn = (b1 · b2 · · · · · bn )1/n
be the geometric average of the first n terms. Show that Gn is convergent and
lim Gn = L.
n→∞
The following result extends the Cesaró-Stoltz lemma to sequences of random variables.
Proposition 3.3.14 Let Xn be a sequence of random variables on the probability space (Ω, F, P ),
such that
Xn+1 − Xn
ac-lim = L.
n→∞ Yn+1 − Yn
Then
Xn
ac-lim = L.
n→∞ Yn
Xn (ω)
B = {ω ∈ Ω; lim = L}.
n→∞ Yn (ω)
56
Since for any given state of the world ω, the sequences xn = Xn (ω) and yn = Yn (ω) are
numerical sequences, Lemma 3.3.12 yields the inclusion A ⊂ B. This implies P (A) ≤ P (B).
Since P (A) = 1, it follows that P (B) = 1, which leads to the desired conclusion.
Remark 3.3.16 Let Xn and Yn denote the prices of two stocks in the day n. The previous
result states that if Corr(Xn+1 − Xn , Yn+1 − Yn ) → 1, as n → ∞, then Corr(Xn , Yn ) → 1. So,
if the correlation of the the daily changes of the stock price tends to 1 in the long run, then the
stock prices correlation does the same.
Example 3.3.17 Let Sn denote the price of a stock in the day n, and assume that
ac-lim Sn = L.
n→∞
Then
S1 + · · · + Sn
ac-lim = L and ac-lim (S1 · · · · · Sn )1/n = L.
n→∞ n n→∞
This says, that if almost all future simulations of the stock price approach the steady state
limit L, the arithmetic and geometric averages converge to the same limit. The statement is a
consequence of Proposition 3.3.14 and follows a similar proof as Example 3.3.1. Asian options
have payoffs depending on these type of averages, as we shall se in the sequel.
Research topic: Extend Proposition 3.3.14 to the continuous case of stochastic processes.
Proof: For any state of the world ω ∈ Ω consider the sequences xn = Xn (ω), yn = Yn (ω) and
zn = Zn (ω) and apply the usual Squeeze Theorem to them.
Remark 3.3.20 The previous theorem remains valid if n is replaced by a continuous positive
parameter t.
Wt sin(Wt )
Example 3.3.3 Show that ac-lim = 0.
t→∞ t
Wt sin(Wt ) Wt
Proof: Consider the sequences Xt = 0, Yt = and Zt = . From Application
t t
3.3.11 we have ac-lim Zt = 0. Applying the Squeeze Theorem we obtain the desired result.
t→∞
58
Chapter 4
Stochastic Integration
This chapter deals with one of the most useful stochastic integral, called the Ito integral. This
type of stochastic integral was introduced in 1944 by the Japanese mathematician K. Ito, and
was originally motivated by a construction of diffusion processes.
59
60
Remark 4.0.22 The infinitesimal version of the previous result is obtained by replacing t − s
with dt
1. E[dWt2 ] = dt;
2. V ar[dWt2 ] = 2dt2 .
We shall see in a next section that in fact dWt2 and dt are equal in a mean square sense.
Divide the interval [a, b] into n subintervals using the partition points
We emphasize that the intermediate points are the left endpoints of each interval, and this is
the way they should be always chosen. Since the process Ft is nonanticipative, the random
variables Fti and Wti+1 − Wti are independent; this is an important feature in the definition of
the Ito integral.
The Ito integral is the limit of the partial sum Sn
61
Z T
ms-lim Sn = Ft dWt ,
n→∞ 0
provided the limit exists. It can be shown that the choice of partition does not influence the
value of the Ito integral. This is the reason why, for practical purposes, it suffices to assume
the intervals equidistant, i.e.
(b − a)
ti+1 − ti = a + , i = 0, 1, · · · , n − 1.
n
The previous convergence is in the mean square sense, i.e.
h³ Z b ´2 i
lim E Sn − Ft dWt = 0.
n→∞ a
In particular, taking c = 1, since the Brownian motion starts at 0, we have the following formula
Z T
dWt = WT .
0
62
Since
1
xy = [(x + y)2 − x2 − y 2 ],
2
letting x = Wti and y = Wti+1 − Wti yields
1 2 1 1
Wti (Wti+1 − Wti ) = Wti+1 − Wt2i − (Wti+1 − Wti )2 .
2 2 2
Then after pair cancelations the sum becomes
n−1 n−1 n−1
1X 2 1X 2 1X
Sn = Wti+1 − W ti − (Wti+1 − Wti )2
2 i=0 2 i=0 2 i=0
n−1
1 2 1X
= Wtn − (Wti+1 − Wti )2
2 2 i=0
Using tn = T , we get
n−1
1 2 1X
Sn = W T − (Wti+1 − Wti )2 .
2 2 i=0
Since the first term is independent of n, we have
n−1
1 2 1X
ms- lim Sn = WT − ms- lim (Wti+1 − Wti )2 . (4.2.1)
n→∞ 2 n→∞ 2
i=0
n−1
X n−1
X
V ar[Xn ] = V ar[(Wti+1 − Wti )2 ] = 2(ti+1 − ti )2
i=0 i=0
2T
= ,
n2
63
where we used that the partition is equidistant. Since Xn satisfies the conditions
E[Xn ] = T, ∀n ≥ 1;
V ar[Xn ] → 0, n → ∞,
n−1
X
ms- lim (Wti+1 − Wti )2 = T. (4.2.2)
n→∞
i=0
This states that the quadratic variation of the Brownian motion is T . Hence (4.2.1) becomes
1 2 1
ms- lim Sn = W − T.
n→∞ 2 T 2
We have obtained the following explicit formula of a stochastic integral
Z T
1 2 1
Wt dWt = W − T.
0 2 T 2
It worth noting that the right side contains random variables depending on the limits of inte-
gration a and b.
while the left side can be regarded as a stochastic integral with respect to dWt2
Z T n−1
X
(dWt )2 := ms- lim (Wti+1 − Wti )2 .
0 n→∞
i=0
dWt2 = dt.
Roughly speaking, the process dWt2 , which is the square of infinitesimal increments of a Brow-
nian motion, is totally predictable. This relation is plays a central role in Stochastic Calculus
and it will be useful when dealing with Ito’s Lemma.
Proposition 4.4.1 Let f (Wt , t), g(Wt , t) be nonanticipating processes and c ∈ R. Then we
have
1. Additivity:
Z T Z T Z T
[f (Wt , t) + g(Wt , t)] dWt = f (Wt , t) dWt + g(Wt , t) dWt .
0 0 0
2. Homogeneity:
Z T Z T
cf (Wt , t) dWt = c f (Wt , t) dWt .
0 0
3. Partition property:
Z T Z u Z T
f (Wt , t) dWt = f (Wt , t) dWt + f (Wt , t) dWt , ∀0 < u < T.
0 0 u
n−1
X
Xn = f (Wti , ti )(Wti+1 − Wti )
i=0
n−1
X
Yn = g(Wti , ti )(Wti+1 − Wti ).
i=0
65
Z T Z T
Since ms-lim Xn = f (Wt , t) dWt and ms-lim Yn = g(Wt , t) dWt , using Proposition
n→∞ 0 n→∞ 0
1.13.2 yields
Z T ³ ´
f (Wt , t) + g(Wt , t) dWt
0
X³
n−1 ´
= ms-lim f (Wti , ti ) + g(Wti , ti ) (Wti+1 − Wti )
n→∞
i=0
h n−1
X³ n−1
X i
= ms-lim f (Wti , ti )(Wti+1 − Wti ) + g(Wti , ti )(Wti+1 − Wti )
n→∞
i=0 i=0
= ms-lim (Xn + Yn ) = ms-lim Xn + ms-lim Yn
n→∞ n→∞ n→∞
Z T Z T
= f (Wt , t) dWt + g(Wt , t) dWt .
0 0
The proof of the parts 2 and 3 are left as an exercise to the reader.
Some other properties, such as monotonicity, do not hold in general. It is possible to have a
RT
nonnegative random variable Ft for which the random variable 0 Ft dWt has negative values.
Some of the random variable properties of the Ito integral are given by the following result.
Proposition 4.4.2 We have
1. Zero mean:
hZ b i
E f (Wt , t) dWt = 0.
a
2. Isometry:
h³ Z b ´2 i hZ b i
E f (Wt , t) dWt =E f (Wt , t)2 dt .
a a
3. Covariance:
h³ Z b ´³ Z b ´i hZ b i
E f (Wt , t) dWt g(Wt , t) dWt =E f (Wt , t)g(Wt , t) dt .
a a a
We shall discuss the previous properties giving rough reasons why they hold true. The
detailed proofs are beyond the goal of this books.
Pn−1
1. The Ito integral is the mean square limit of the partial sums Sn = i=0 fti (Wti+1 −
Wti ), where we denoted fti = f (Wti , ti ). Since f (Wt , t) is nonanticipative process, then fti is
independent of the increments Wti+1 − Wti , and then we have
h n−1
X i n−1
X
E[Sn ] = E fti (Wti+1 − Wti ) = E[fti (Wti+1 − Wti )]
i=0 i=0
n−1
X
= E[fti ]E[(Wti+1 − Wti )] = 0,
i=0
because the increments have mean zero. Since each partial sum has zero mean, their limit,
which is the Ito Integral, will also have zero mean.
66
Their product is
³ n−1
X ´³ n−1
X ´
Sn Vn = fti (Wti+1 − Wti ) gtj (Wtj+1 − Wtj )
i=0 j=0
n−1
X n−1
X
= fti gti (Wti+1 − Wti )2 + fti gtj (Wti+1 − Wti )(Wtj+1 − Wtj )
i=0 i6=j
it follows that
n−1
X
E[Sn Vn ] = E[fti gti ]E[(Wti+1 − Wti )2 ]
i=0
n−1
X
= E[fti gti ](ti+1 − ti ),
i=0
67
Rb
which is the Riemann sum for the integral E[ft gt ] dt. a
Rb
From 1 and 2 it follows that the random variable a f (Wt , t) dWt has mean zero and variance
hZ b i hZ b i
V ar f (Wt , t) dWt = E f (Wt , t)2 dt .
a a
Corollary 4.4.3 (Cauchy’s integral inequality) Let f (t) = f (Wt , t) and g(t) = g(Wt , t).
Then
³Z b ´2 ³ Z b ´³ Z b ´
2
E[ft gt ] dt ≤ E[ft ] dt E[gt2 ] dt .
a a a
Proof: It follows from the previous theorem and from the correlation formula |Corr(X, Y )| =
|Corr(X, Y )|
≤ 1.
[V ar(X)V ar(Y )]1/2
Let Ft be the information set at time t. This implies that fti and Wti+1 − Wti are known
n−1
X
at time t, for any ti+1 ≤ t. It follows that the partial sum Sn = fti (Wti+1 − Wti ) is
i=0
Ft -predictable. The following result states that this is also valid in mean square:
Rt
Proposition 4.4.4 The Ito integral 0
fs dWs is Ft -predictable.
The following two results state that if the upper limit of an Ito integral is replaced by the
parameter t we obtain a continuous martingale.
Rt
Since s f (Wu , u) dWu contains only information between s and t, it is unpredictable given the
information set Fs , so
hZ t i
E f (Wu , u) dWu |Fs = 0.
s
Substituting in (4.4.4) yields the desired result.
Rt
Proposition 4.4.6 Consider the process Xt = 0 f (Ws , s) dWs . Then Xt is continuous, i.e.
for almost any state of the world ω ∈ Ω, the path t → Xt (ω) is continuous.
Proof: A rigorous proof is beyond the purpose of this book. We shall provide a rough sketch.
Assume the process f (Wt , t) satisfies E[f (Wt , t)2 ] < M , for some M > 0. Let t0 be fixed and
consider h > 0. Consider the increment Yh = Xt0 +h −Xt0 . Using the aforementioned properties
of the Ito integral we have
h Z t0 +h i
E[Yh ] = E[Xt0 +h − Xt0 ] = E f (Wt , t) dWt = 0
t0
h³ Z t0 +h ´2 i Z t0 +h
E[Yh2 ] = E f (Wt , t) dWt = E[f (Wt , t)2 ] dt
t0 t0
Z t0 +h
< M dt = M h.
t0
The process Yh has zero mean for any h > 0 and its variance tends to 0 as h → 0. Using a
convergence theorem yields that Yh tends to 0 in mean square, as h → 0. This is equivalent
with the continuity of Xt at t0 .
All properties of Ito integrals hold also for Winer integrals. The Wiener integral is a random
variable with mean zero
hZ b i
E f (t) dWt = 0
a
and variance
h³ Z b ´2 i Z b
E f (t) dWt = f (t)2 dt.
a a
However, in the case of Wiener integrals we can say something about the its distribution.
Rb
Proposition 4.5.1 The Wiener integral I(f ) = a f (t) dWt is a normal random variable with
mean 0 and variance Z b
V ar[I(f )] = f (t)2 dt := kf k2L2 .
a
69
Proof: Since increments Wti+1 − Wti are normally distributed with mean 0 and variance
ti+1 − ti , then
f (ti )(Wti+1 − Wti ) ∼ N (0, ti (ti+1 − ti )).
Since these random variables are independent, by the Central Limit Theorem (see Theorem
2.3.1), their sum is also normally distributed, with
n−1
X ³ n−1
X ´
Sn = f (ti )(Wti+1 − Wti ) ∼ N 0, f (ti )(Wti+1 − Wti ) .
i=0 i=0
³ Z b ´
N 0, f (t)2 dWt .
a
The previous convergence holds in distribution, and it still need to be shown in the mean square.
We shall omit this essential proof detail.
RT
Exercise 4.5.2 Show that the random variable X = 1 √1t dWt is normally distributed with
mean 0 and variance ln T .
RT √
Exercise 4.5.3 Let Y = 1 t dWt . Show that Y is normally distributed with mean 0 and
variance (T 2 − 1)/2.
Rt
Exercise 4.5.4 Find the distribution of the integral 0
et−s dWs .
Rt Rt
Exercise 4.5.5 Show that Xt = 0 (2t − u) dWu and Y = 0 (3t − 4u) dWu are Gaussian
processes with mean 0 and variance 73 t3 .
Rt bu
Exercise 4.5.6 Find all constants a, b such that Xt = 0
(a + t ) dWu is a Brownian motion
process.
For predictability reasons, the intermediate points are the left-handed limit to the endpoints of
each interval. Since the process Ft is non-anticipative, the random variables Fti− and Mti+1 −
Mti are independent.
The integral of Ft− with respect to Mt is the mean square limit of the partial sum Sn
Z T
ms-lim Sn = Ft− dMt ,
n→∞ 0
provided the limit exists. More precisely, this convergence means that
h³ Z b ´2 i
lim E Sn − Ft− dMt = 0.
n→∞ a
1
Using xy = [(x + y)2 − x2 − y 2 ], by letting x = Mti− and y = Mti+1 − Mti , we get (Where
2
does a minus go?)
1 2 1 1
Mti− (Mti+1 − Mti ) = M − M 2 − (Mti+1 − Mti )2 .
2 ti+1 2 ti 2
After pair cancelations we have
n−1 n−1 n−1
1X 2 1X 2 1X
Sn = Mti+1 − Mti − (Mti+1 − Mti )2
2 i=0 2 i=0 2 i=0
n−1
1 2 1X
= M tn − (Mti+1 − Mti )2
2 2 i=0
Since tn = T , we get
n−1
1 2 1X
Sn = MT − (Mti+1 − Mti )2 .
2 2 i=0
The second term on the right is the quadratic variation of Mt , using formula (2.8.9) yields that
Sn converges in mean square towards 12 MT2 − 12 NT , since N0 = 0.
Hence we have arrived at the following formula
Z T
1 2 1
Mt− dMt = M − NT .
0 2 T 2
The stochastic integral with respect to the compensated Poisson process Mt has in general
the following properties, which are left as an exercise to the reader
Proposition 4.6.2 We have
1. Linearity:
Z b Z b Z b
(αf + βg) dMt = α f dMt + β g dMt , α, β ∈ R;
a a a
2. Zero mean:
hZ b i
E f dMt = 0;
a
2. Isometry:
h³ Z b ´2 i hZ b i
E f dMt =E f 2 dNt ;
a a
Exercise 4.6.3 Let ω be a fixed state of the world and assume the sample path t → Nt (ω) has
a jump in the interval (a, b). Show that the Riemann-Stieltjes integral
Z b
Nt (ω) dNt
a
Exercise 4.6.4 Let Nt− denote the left-hand limit of Nt . Show that Nt− is predictable, while
Nt is not.
The previous exercises provide the reason why in the following we shall work with Mt− instead
Z b Z b
of Mt : the integral Mt dNt might not exist, while Mt− dNt does exist.
a a
Stochastic Differentiation
make sense in stochastic calculus. The only quantities allowed to be used are the infinitesimal
changes of the process, in our case dWt .
The infinitesimal change of a process
The change in the process Xt between instances t and t + ∆t is given by ∆Xt = Xt+∆t − Xt .
When ∆t is infinitesimally small, we obtain the infinitesimal change of a process Xt
dXt = Xt+dt − Xt .
d(c Xt ) = c dXt .
The verification follows from a straightforward application of the infinitesimal change formula
73
74
The proof is similar with the one for the sum rule.
The product rule
If Xt and Yt are two stochastic processes, then
The proof follows from Ito’s formula and shall be postponed for the time being.
When the process Yt is replaced by the deterministic function f (t), and Xt is an Ito diffusion,
then the previous formula becomes
³ X ´ f (t)dX − X df (t)
t t t
d = .
f (t) f (t)2
75
Applying the product rule and the fundamental relation (dWt )2 = dt, yields
d(Wt2 ) = Wt dWt + Wt dWt + dWt dWt = 2Wt dWt + dt.
Example 5.2.2 Show that
d(Wt3 ) = 3Wt2 dWt + 3Wt dt.
Applying the product rule and the previous exercise yields
d(Wt3 ) = d(Wt · Wt2 ) = Wt d(Wt2 ) + Wt2 dWt + d(Wt )2 dWt
= Wt (2Wt dWt + dt) + Wt2 dWt + dWt (2Wt dWt + dt)
= 2Wt2 dWt + Wt dt + Wt2 dWt + 2Wt (dWt )2 + dt dWt
= 3Wt2 dWt + 3Wt dt,
where we used (dWt )2 = dt and dt dWt = 0.
Example 5.2.3 Show that d(tWt ) = Wt dt + t dWt .
Using the product rule and t dWt = 0, we get
d(tWt ) = Wt dt + t dWt + dt dWt
= Wt dt + t dWt .
Rt
Example 5.2.4 Let Zt = 0
Wu du be the integrated Brownian motion. Show that
dZt = Wt dt.
The infinitesimal change of Zt is
Z t+dt
dZt = Zt+dt − Zt = Ws ds = Wt dt,
t
df (x) = f 0 (x)dx.
We shall work out a similar formula in the stochastic environment. In this case the deter-
ministic function x(t) is replaced by a stochastic process Xt . The composition between the
differentiable function f and the process Xt is denoted by Ft = f (Xt ). Since the increments
involving powers of dt2 or higher are neglected, we may assume that the same holds true for
the increment dXt , i.e., dXt = O(dt). Then the expression (5.3.1) becomes
¡ ¢ 1 ¡ ¢¡ ¢2
dFt = f 0 Xt dXt + f 00 Xt dXt . (5.3.2)
2
In the computation of dXt we may take into the account stochastic relations such as dWt2 = dt,
or dt dWt = 0.
Theorem 5.3.1 (Ito’s formula for diffusions) If Xt is an Ito diffusion, and Ft = f (Xt ),
then
h b(Wt , t) 00 i
dFt = a(Wt , t)f 0 (Xt ) + f (Xt ) dt + b(Wt , t)f 0 (Xt ) dWt . (5.3.3)
2
77
Proof: We shall provide a formal proof. Using the relations dWt2 = dt and dt2 = dWt dt = 0,
we have
³ ´2
(dXt )2 = a(Wt , t)dt + b(Wt , t)dWt
= a(Wt , t)2 dt2 + 2a(Wt , t)b(Wt , t)dWt dt + b(Wt , t)2 dWt2
= b(Wt , t)2 dt.
1 00
dFt = f (Wt )dt + f 0 (Wt ) dWt . (5.3.4)
2
Particular cases
1. If f (x) = xα , with α constant, then f 0 (x) = αxα−1 and f 00 (x) = α(α − 1)xα−2 , then (5.3.4)
becomes the following useful formula
1
d(Wtα ) = α(α − 1)Wtα−2 dt + αWtα−1 dWt .
2
1
d(ekWt ) = kekWt dWt + k 2 ekWt dt.
2
Exercise 5.3.3 Use the previous rules to find the following increments
a. d(Wt eWt )
b. d(3Wt2 + 2e5Wt )
2
c. d(et+Wt )
¡ ¢
d. d (t + Wt )n .
³1 Z t ´
e. d Wu du
t 0
³1 Z t ´
f. d α eWu du , where α is a constant.
t 0
In the case when the function f = f (t, x) is also time dependent, the analog of (5.3.1) is
given by
1
df (t, x) = ∂t f (t, x)dt + ∂x f (t, x)dx + ∂x2 f (t, x)(dx)2 + O(dx)3 + O(dt)2 . (5.3.5)
2
Substituting x = Xt yields
1
df (t, Xt ) = ∂t f (t, Xt )dt + ∂x f (t, Xt )dXt + ∂x2 f (t, Xt )(dXt )2 . (5.3.6)
2
If Xt is an Ito diffusion we obtain an extra-term in formula (5.3.3)
h b(Wt , t) 2 i
dFt = ∂t f (t, Xt ) + a(Wt , t)∂x f (t, Xt ) + ∂x f (t, Xt ) dt
2
+b(Wt , t)∂x f (t, Xt ) dWt . (5.3.7)
that yields
Z T
1
Mt− dMt = (M 2 − NT ).
0 2 T
The left-hand limit is used for predictability reasons, see section 4.6.
∂f ∂f ∂f
dFt = (t, Xt , Yt )dt + (t, Xt , Yt )dXt + (t, Xt , Yt )dYt
∂t ∂x ∂y
1 ∂2f 2 1 ∂2f
+ (t, X t , Y t )(dX t ) + (t, Xt , Yt )(dYt )2
2 ∂x2 2 ∂y 2
∂2f
+ (t, Xt , Yt )dXt dYt .
∂x∂y
Particular cases
In the case when Ft = f (Xt , Yt ), with Xt = Wt1 , Yt = Wt2 independent Brownian motions, we
have
∂f ∂f 1 ∂2f 1 ∂2f
dFt = dWt1 + dWt2 + 2
(dWt1 )2 + (dWt2 )2
∂x ∂y 2 ∂x 2 ∂y 2
1 ∂2f
+ dWt1 dWt2
2 ∂x∂y
∂f ∂f 1 ³ ∂2f ∂2f ´
= dWt1 + dWt2 + 2
+ 2 dt
∂x ∂y 2 ∂x ∂y
The expression
1 ³ ∂2f ∂2f ´
∆f = 2
+ 2
2 ∂x ∂y
is called the Laplacian of f . We can rewrite the previous formula as
∂f ∂f
dFt = dWt1 + dWt2 + ∆f dt
∂x ∂y
A function f with ∆f = 0 is called harmonic. The aforementioned formula in the case of
harmonic functions takes the very simple form
∂f ∂f
dFt = dWt1 + dWt2 .
∂x ∂y
80
Exercise 5.3.6 Use the previous formulas to find dFt in the following cases
(a) Ft = (Wt1 )2 + (Wt2 )2
(b) Ft = ln[(Wt1 )2 + (Wt2 )2 ].
p
Exercise 5.3.7 Consider the Bessel process Rt = (Wt1 )2 + (Wt2 )2 , where Wt1 and Wt2 are
two independent Brownian motions. Prove that
1 W1 W2
dRt = dt + t dWt1 + t dWt2 .
2Rt Rt Rt
Example 5.3.1 (The product rule) Let Xt and Yt be two processes. Show that
Example 5.3.2 (The quotient rule) Let Xt and Yt be two processes. Show that
³ X ´ Y dX − X dY − dX dY Xt
t t t t t t t
d = + 2 (dYt )2 .
Yt Yt2 Yt
Computing a stochastic integral starting from the definition of the Ito integral is a quite ineffi-
cient method. Like in the elementary Calculus, several methods can be developed to compute
stochastic integrals. In order to keep the analogy with the elementary Calculus, we have called
them Fundamental Theorem of Stochastic Calculus and Integration by Parts. The integration
by substitution is more complicated in the stochastic environment and we have considered only
a particular case of it, which we called The method of heat equation.
The integral on the left side can be computed as in the following. If consider the partition
0 = t0 < t1 < · · · < tn−1 < tn = t, then
Z t n−1
X
dXs = ms-lim (Xtj+1 − Xtj ) = Xt − Xa ,
a n→∞
j=0
The following result bares its name from the analogy with the similar result from elementary
Calculus.
Theorem 6.0.8 (The Fundamental Theorem of Stochastic Calculus) (i) For any a <
t, we have
³Z t ´
d f (s, Ws )dWs = f (t, Wt )dWt .
a
81
82
Hence dXt = dYt , or d(Xt − Yt ) = 0. Since the process Xt − Yt has zero increments, then
Xt − Yt = c, constant. Taking t = 0, yields
Z 0 ³W2
0 0´
c = X0 − Y0 = Ws dWs − − = 0,
0 2 2
and hence c = 0. It follows that Xt = Yt , which verifies the desired relation.
Example 6.0.4 Verify the formula
Z t ´ 1Z t
t³ 2
sWs dWs = Wt − 1 − Ws2 ds.
0 2 2 0
Rt t³ 2 ´
1
Rt
Consider the stochastic processes Xt = 0 sWs dWs , Yt = Wt − 1 , and Zt = 2 0
Ws2 ds.
2
The Fundamental Theorem yields
dXt= tWt dWt
1 2
dZt = W dt.
2 t
Applying Ito’s formula, see Exercise 5.3.4, we get
³t³ ´´ 1 ³t´
dYt = d Wt2 − 1 = d(tWt2 ) − d
2 2 2
1h 2
i 1
= (1 + Wt )dt + 2tWt dWt − dt
2 2
1 2
= W dt + tWt dWt .
2 t
We can easily see that
dXt = dYt − dZt .
This implies d(Xt − Yt + Zt ) = 0, i.e. Xt − Yt + Zt = c, constant. Since X0 = Y0 = Z0 = 0, it
follows that c = 0. This proves the desired relation.
83
Consider the function f (t, x) = 13 x3 − tx, and let Ft = f (t, Wt ). Since ∂t f = −x, ∂x f = x2 − t,
and ∂x2 f = 2x, then Ito’s formula provides
1
dFt = ∂t f dt + ∂x f dWt + ∂x2 f (dWt )2
2
2 1
= −Wt dt + (Wt − t) dWt + 2Wt dt
2
= (Wt2 − t)dWt .
This formula is to be used when integrating a product between a function of t and a function
of the Brownian motion Wt , for which an antiderivative is known. The following two particular
cases are important and useful in applications.
1. If g(Wt ) = Wt , the aforementioned formula takes the simple form
Z b ¯t=b Z b
¯
f (t) dWt = f (t)Wt ¯ − f 0 (t)Wt dt. (6.0.2)
a t=a a
Z T
Application 1 Consider the Wiener integral IT = t dWt . From the general theory, see
0
Proposition 4.5.1, it is known that I is a random variable normally distributed with mean 0
and variance Z T
T3
V ar[IT ] = t2 dt = .
0 3
Recall the definition of integrated Brownian motion
Z t
Zt = Wu du.
0
Formula (6.0.2) yields a relationship between I and the integrated Brownian motion
Z T Z T
IT = t dWt = T WT − Wt dt = T WT − ZT ,
0 0
and hence IT + ZT = T WT . This relation can be used to compute the covariance between IT
and ZT .
Cov(IT + ZT , IT + ZT ) = V ar[T WT ] ⇐⇒
V ar[IT ] + V ar[ZT ] + 2Cov(IT , ZT ) = T 2 V ar[WT ] ⇐⇒
T 3 /3 + T 3 /3 + 2Cov(IT , ZT ) = T 3 ⇐⇒
Cov(IT , ZT ) = T 3 /6,
where we used that V ar[ZT ] = T 3 /3. The processes It and Zt are not independent. Their
correlation coefficient is 0.5 as the following calculation shows
Cov(IT , ZT ) T 3 /6
Corr(IT , ZT ) = ³ ´1/2 = T 3 /3
V ar[IT ]V ar[ZT ]
= 1/2.
x2
Application 2 If let g(x) = 2 in formula (6.0.3), we get
Z b
Wt2 ¯¯b 1
Wt dWt = ¯ − (b − a).
a 2 a 2
It worth noting that letting a = 0 and b = T we retrieve a formula proved by direct methods
in a previous chapter
Z T
W2 T
Wt dWt = T − .
0 2 2
Next we shall deduct inductively from (6.0.3) an explicit formula for the stochastic inte-
RT n+1
gral 0 Wtn dWt , for n natural number. Letting g(x) = xn+1 in (6.0.3) and denoting In =
RT n
0
Wt dWt we obtain the recursive formula
1 n+1
In+1 = W n+2 − In , n ≥ 1.
n+2 T 2
85
Since
n!
n(n − 1) · · · (n − k + 1) =
,
(n − k)!
the aforementioned formula leads to the explicit formula
Z T X n−2
1 n! n! T
Wtn dWt = WTn+1 + (−1)k+1 k+1 W n−k + (−1)n n .
0 n+1 2 (n − k)! T 2 2
k=0
Z T
1 3 1 2 T
Wt2 dWt = W − W + . (6.0.4)
0 3 T 2 T 2
Z T
1 4 1 3 3 3
Wt3 dWt = WT − WT + 2 WT2 − 2 T. (6.0.5)
0 4 2 2 2
86
Application 3
RT
Choosing f (t) = eαt and g(x) = cos x, we shall compute the stochastic integral 0 eαt cos Wt dWt
using the formula of integration by parts
Z T Z T
αt
e cos Wt dWt = eαt (sin Wt )0 dWt
0 0
¯T Z T Z
1 T αt
αt ¯ αt 0
= e sin Wt ¯ − (e ) sin Wt dt − e (cos Wt )00 dt
0 0 2 0
Z T Z
1 T αt
= eαT sin WT − α eαt sin Wt dt + e sin Wt dt
0 2 0
³ Z
1 ´ T αt
= eαT sin WT − α − e sin Wt dt.
2 0
1
The particular case α = 2 leads to the following exact formula of a stochastic integral
Z T
t T
e 2 cos Wt dWt = e 2 sin WT . (6.0.6)
0
RT
In a similar way, we can obtain an exact formula for the stochastic integral 0
eβt sin Wt dWt
as follows
Z T Z T
eβt sin Wt dWt = − eβt (cos Wt )0 dWt
0 0
¯T Z T Z T
¯ 1
= −eβt cos Wt ¯ + β eβt cos Wt dt − eβt cos Wt dt.
0 0 2 0
1
Taking β = 2 yields the closed form formula
Z T
t T
e 2 sin Wt dWt = 1 − e 2 cos WT . (6.0.7)
0
is
Z T
t T
e 2 +iWt dWt = i(1 − e 2 +iWT ).
0
This formula is of theoretical value. In practice, the term dXt dYt needs to be computed using
the rules Wt2 = dt, and dt dWt = 0.
Exercise 6.0.12 (a) Let T > 0. Show the following relation using integration by parts
Z T Z T
2Wt 1 − Wt2
dWt = ln(1 + Wt2 ) − ds.
0 1 + Wt2 0 (1 + Wt2 )2
(b) Show that for any real number x the following double inequality holds
1 1 − x2
− ≤ ≤ 1.
8 (1 + x2 )2
T
− ≤ E[ln(1 + Wt2 )] ≤ T.
8
(e) Use Jensen’s inequality to get
Example 6.0.6 Find all solutions of the equation (6.0.8) of the type
ϕ(t, x) = a(t) + b(x).
Since the right side is a function of x only, while the right side is a function of variable t,
the only case when the previous equation is satisfied is when both sides are equal to the same
constant C. This is called a separation constant. Therefore a(t) and b(x) satisfy the equations
1 00
a0 (t) = −C, b (x) = C.
2
Integrating yields a(t) = −Ct + C0 and b(x) = Cx2 + C1 x + C2 . It follows that
ϕ(t, x) = C(x2 − t) + C1 x + C3 ,
Example 6.0.7 Find all solutions of the equation (6.0.8) of the type
ϕ(t, x) = a(t)b(x).
ϕ(t, x) = a(t)b(x) = c1 x + c0 , c0 , c1 ∈ R
In particular, the functions x, x2 − t, ex−t/2 , e−x−t/2 , et/2 sin x and et/2 cos x, or any linear
combination of them are solutions of the heat equation (6.0.8). However, there are other
solutions which are not of the previous type.
90
Exercise 6.0.13 Prove that ϕ(t, x) = 13 x3 − tx is a solution of the heat equation (6.0.8).
2
Exercise 6.0.14 Show that ϕ(t, x) = t−1/2 e−x /(2t)
is a solution of the heat equation (6.0.8)
for t > 0.
Theorem 6.0.15 Let ϕ(t, x) be a solution of the heat equation (6.0.8) and denote f (t, x) =
∂x ϕ(t, x). Then
Z b
f (t, Wt ) dWt = ϕ(b, Wb ) − ϕ(a, Wa ).
a
Choose the solution of the heat equation (6.0.8) given by ϕ(t, x) = x2 − t. Then f (t, x) =
∂x ϕ(t, x) = 2x. Theorem 6.0.15 yields
Z T Z T ¯T
¯
2Wt dWt = f (t, Wt ) dWt = ϕ(t, x)¯ = WT2 − T.
0 0 0
Consider the function ϕ(t, x) = 31 x3 − tx, which is a solution of the heat equation (6.0.8), see
Exercise 6.0.13. Then f (t, x) = ∂x ϕ(t, x) = x2 − t. Applying Theorem 6.0.15 yields
Z T Z T ¯T
¯ 1
(Wt2 − t) dWt = f (t, Wt ) dWt = ϕ(t, Wt )¯ = WT3 − T WT .
0 0 0 3
91
2
From Exercise 6.0.14 we have that ϕ(t, x) = t−1/2 e−x /(2t) is a solution of the homogeneous
2
heat equation. Since f (t, x) = ∂x ϕ(t, x) = −t−3/2 xe−x /(2t) , applying Theorem 6.0.15 yields to
the desired result. The reader can easily fill in the details.
Integration techniques will be used when solving stochastic differential equations in the next
chapter.
Exercise 6.0.22 Find the value of the stochastic integrals
Z 1 √
(a) et cos( 2Wt ) dWt
0
Z 3
(b) e2t cos(2Wt ) dWt
0
Z 4 √
(c) e−t+ 2Wt
dWt .
0
Exercise 6.0.23 Let ϕ(t, x) be a solution of the following non-homogeneous heat equation with
time-dependent and uniform heat source G(t)
1
∂t ϕ + ∂x2 ϕ = G(t).
2
Denote f (t, x) = ∂x ϕ(t, x). Then
Z b Z b
f (t, Wt ) dWt = ϕ(b, Wb ) − ϕ(a, Wa ) − G(t) dt.
a a
where the last integral is taken in the Ito sense. Relation (7.1.2) is taken as the definition for
the stochastic differential equation (7.1.1), so the definition of stochastic differential equations
is fictions. However, since it is convenient to use stochastic differentials informally, we shall
approach stochastic differential equations by analogy with the ordinary differential equations,
and try to present the same methods of solving equation in the new stochastic environment.
The functions a(t, Wt , Xt ) b(t, Wt , Xt ) are called drift rate and volatility. A process Xt is
called a solution for the stochastic equation (7.1.1) if it satisfies the equation. In the following
we shall start with an example.
Example 7.1.1 (The Brownian Bridge) Let a, b ∈ R. Show that the process
Z t
1
Xt = a(1 − t) + bt + (1 − t) dWs , 0 ≤ t < 1
0 1 − s
is a solution of the stochastic differential equation
b − Xt
dXt = dt + dWt , 0 ≤ t < 1, X0 = a.
1−t
We shall perform a routine verification to show that Xt is a solution. First we compute the
b − Xt
quotient :
1−t
93
94
Z t
1
b − Xt = b − a(1 − t) − bt − (1 − t) dWs
0 1−s
Z t
1
= (b − a)(1 − t) − (1 − t) dWs ,
0 1 − s
Using
³Z t
1 ´ 1
d dWs = dWt ,
0 1−s 1−t
the product rule yields
Z t ³Z t 1 ´
1
dXt = a d(1 − t) + bdt + d(1 − t) dWs + (1 − t)d dWs
0 1−s 0 1−s
³ Z t ´
1
= b−a− dWs dt + dWt
0 1 − s
b − Xt
= dt + dWt ,
1−t
where the last identity comes from (7.1.3). We just verified that the process Xt is a solution of
the given stochastic equation. The question of how this solution was obtained in the first place,
is the subject of study for the next few sections.
d
E[Xt ] = E[a(t, Wt , Xt )].
dt
We note that Xt is not differentiable, but its expectation E[Xt ] is. This equation can be solved
exactly in a few particular cases.
d
1. If a(t, Wt , Xt ) = a(t), then dt E[Xt ] = a(t) with the exact solution E[Xt ] = X0 +
Rt
0
a(s) ds.
95
2. If a(t, Wt , Xt ) = α(t)Xt + β(t), with α(t) and β(t) continuous deterministic functions.
Then
d
E[Xt ] = α(t)E[Xt ] + β(t),
dt
which is a linear differential equation in E[Xt ]. Its solution is given by
³ Z t ´
E[Xt ] = eA(t) X0 + e−A(s) β(s) ds , (7.2.5)
0
Rt
where A(t) = 0 α(s) ds. It worth noting that the expectation E[Xt ] does not depend on the
volatility term b(t, Wt , Xt ).
Example 7.2.1 If dXt = (2Xt + e2t )dt + b(t, Wt , Xt )dWt , then
Exercise 7.2.3 State the previous result in the particular case when a(x) = sin x, with 0 ≤
x ≤ π.
E[Xt ] = eA(t) X0
Z t
V ar[Xt ] = e2A(t) e−A(s) b2 (s) ds,
0
Rt
where A(t) = 0
α(s) ds.
Proof: The expression of E[Xt ] follows directly from formula (7.2.5) with β = 0. In order to
compute the second moment we first compute
where we used Ito’s formula. If let Yt = Xt2 , the previous equation becomes
¡ ¢ p
dYt = 2α(t)Yt + b2 (t) dt + 2b(t) Yt dWt .
Applying formula (7.2.5) with α(t) replaced by 2α(t) and β(t) by b2 (t), yields
³ Z t ´
2A(t)
E[Yt ] = e Y0 + e−2A(s) b2 (s) ds ,
0
Remark 7.2.5 We note that the previous equation is of linear type. This shall be solved
explicitly in a future section.
The mean and variance for a given stochastic process can be computed by working out the
associated stochastic equation. We shall provide next a few examples.
Example 7.2.2 Find the mean and variance of ekWt , with k constant.
97
kWt
If let f (t) = E[e ], then differentiating the previous relations yields the differential equation
1 2
f 0 (t) = k f (t)
2
2
with the initial condition f (0) = E[ekW0 ] = 1. The solution is f (t) = ek t/2
, and hence
2
E[ekWt ] = ek t/2
.
The variance is
2 2
V ar[ekWt ] = E[e2kWt ] − (E[ekWt ])2 = e4k t/2
− ek t
2 2
= ek t (ek t − 1).
We shall set up a stochastic differential equation for Wt eWt . Using the product formula and
Ito’s formula yields
Let f (t) = E[Wt eWt ]. Using E[eWs ] = et/2 , the previous integral equation becomes
Z t
1
f (t) = ( f (s) + es/2 ) ds,
0 2
98
(2k)! k
E[Wt2k ] = t , E[Wt2k+1 ] = 0.
2k k!
In particular, E[Wt4 ] = 3t2 , E[Wt6 ] = 15t3 .
From Ito’s formula we have
n(n − 1) n−2
d(Wtn ) = nWtn−1 dWt + Wt dt.
2
Integrate and get Z Z
t
n(n − 1) t n−2
Wtn = n Wsn−1 dWs + Ws ds.
0 2 0
Since the expectation of the first integral on the right side is zero, taking the expectation yields
the following recursive relation
Z
n(n − 1) t
E[Wtn ] = E[Wsn−2 ] ds.
2 0
Using the initial values E[Wt ] = 0 and E[Wt2 ] = t, the method of mathematical induction
implies that E[Wt2k+1 ] = 0 and E[Wt2k ] = (2k)!
2k k!
tk .
Exercise 7.2.7 Find E[sin Wt ].
From Ito’s formula
1
d(sin Wt ) = cos Wt dWt − sin Wt dt,
2
then integrating yields
Z t Z t
1
sin Wt = cos Ws dWs − sin Ws ds.
0 2 0
Exercise 7.2.9 Use the previous exercise and the definition of expectation to show that
Z ∞
2 π 1/2
(a) e−x cos x dx = 1/4 ;
−∞ e
Z ∞ r
−x2 /2 2π
(b) e cos x dx = .
−∞ e
Not in all cases can the mean and the variance be obtained directly from the stochastic
equation. In these cases we need more powerful methods that produce closed form solutions.
In the next sections we shall discuss several methods of solving stochastic differential equation.
Rt
Using the property of Wiener integrals, 0 b(s) dWs is Gaussian distributed with mean 0 and
Rt
variance 0 b2 (s) ds. Then Xt is Gaussian (as a sum between a predictable function and a
100
Gaussian), with
Z t Z t
E[Xt ] = E[X0 + a(s) ds + b(s) dWs ]
0 0
Z t Z t
= X0 + a(s) ds + E[ b(s) dWs ]
0 0
Z t
= X0 + a(s),
0
Z t Z t
V ar[Xt ] = V ar[X0 + a(s) ds + b(s) dWs ]
0 0
hZ t i
= V ar b(s) dWs
0
Z t
= b2 (s) ds,
0
Exercise 7.3.2 Solve the following stochastic differential equations for t ≥ 0 and determine
the mean and the variance of the solution
(a) dXt = cos t dt − sin t dWt , X0 = 1.
√
(b) dXt = et dt + t dWt , X0 = 0.
t 3/2
(c) dXt = 1+t 2 dt + t dWt , X0 = 1.
If the drift and the volatility depend on both variables t and Wt , the stochastic differential
equation
dXt = a(t, Wt )dt + b(t, Wt )dWt , t≥0
defines an Ito diffusion. Integrating yields the solution
Z t Z t
Xt = X0 + a(s, Ws ) ds + b(s, Ws ) dWs .
0 0
There are several cases when both integrals can be computed explicitly.
dXt = dt + Wt dWt , X0 = 1.
Rt
Let Zt = 0 Ws ds denote the integrated Brownian motion process. Integrating the equation
between 0 and t and using (6.0.4), yields
Z s Z t Z t
Xt = dXs = (Ws − 1)ds + Ws2 dWs
0 0 0
1 1 t
= Zt − t + Wt3 − Wt2 −
3 2 2
1 3 1 2 t
= Zt + Wt − Wt − .
3 2 2
Example 7.3.3 Solve the stochastic differential equation
Integrating yields
Z t Z t
2
Xt = s ds + es/2 cos Ws dWs
0 0
3
t
= + et/2 sin Wt , (7.3.7)
3
where we used (6.0.11). Even if the process Xt is not Gaussian, we can still compute its mean
and variance. By Ito’s formula we have
1
d(sin Wt ) = cos Wt dWt − sin Wt dt
2
Integrating between 0 and t yields
Z t Z t
1
sin Wt = cos Ws dWs − sin Ws ds,
0 2 0
where we used that sin W0 = sin 0 = 0. Taking the expectation in the previous relation yields
hZ t i 1Z t
E[sin Wt ] = E cos Ws dWs − E[sin Ws ] ds.
0 2 0
From the properties of the Ito integral, the first expectation on the right side is zero. Denoting
µ(t) = E[sin Wt ], we obtain the integral equation
Z
1 t
µ(t) = − µ(s) ds.
2 0
Differentiating yields the differential equation
1
µ0 (t) = − µ(t)
2
with the solution µ(t) = ke−t/2 . Since k = µ(0) = E[sin W0 ] = 0, it follows that µ(t) = 0.
Hence
E[sin Wt ] = 0.
102
Taking the expectation and using that Ito integrals have zero expectation, yields
Z t
E[cos 2Wt ] = 1 − 2 E[cos 2Ws ] ds.
0
If denote m(t) = E[cos 2Wt ], the previous relation becomes an integral equation
Z t
m(t) = 1 − 2 m(s) ds.
0
with the solution m(t) = ke−2t . Since k = m(0) = E[cos 2W0 ] = 1, we have m(t) = e−2t .
Substituting in (7.3.8) yields
et et − e−t
V ar[Xt ] = (1 − e−2t ) = = sinh t.
2 2
In conclusion, the solution Xt has the mean and the variance given by
t3
E[Xt ] = , V ar[Xt ] = sinh t.
3
and then find the distribution of the solution Xt and its mean and variance.
103
Dividing by et/2 , integrating between 0 and t, and using formula (6.0.10) yields
Z t Z t
−s/2
Xt = e ds + e−s/2+Ws dWs
0 0
= 2(1 − e−t/2 ) + e−t/2 eWt − 1
= 1 + e−t/2 (eWt − 2).
The process Xt has the distribution of a sum between the predictable function 1 − 2e−t/2 and
the log-normal process e−t/2+Wt .
Then ∂t f = et x2 + T 0 (t) and ∂x2 f = 2et . Substituting in the first equation yields
et (1 + x2 ) = et x2 + T 0 (t) + et .
This implies T 0 (t) = 0, or T = c constant. Hence f (t, x) = x + et x2 + c, and Xt = f (t, Wt ) =
Wt + et Wt2 + c. Since X0 = 0, it follows that c = 0. The solution is Xt = Wt + et Wt2 .
Example 7.4.2 Find the solution of
¡ ¢
dXt = 2tWt3 + 3t2 (1 + Wt ) dt + (3t2 Wt2 + 1)dWt , X0 = 0.
The coefficient functions are a(t, x) = 2tx3 + 3t2 (1 + x) and b(t, x) = 3t2 x2 + 1. The associated
system is given by
1
2tx3 + 3t2 (1 + x) = ∂t f (t, x) + ∂x2 f (t, x)
2
3t2 x2 + 1 = ∂x f (t, x).
Integrate partially in the second equation yields
Z
f (t, x) = (3t2 x2 + 1) dx = t2 x3 + x + T (t).
Then ∂t f = 2tx3 + T 0 (t) and ∂x2 f = 6t2 x, and plugging in the first equation we get
1
2tx3 + 3t2 (1 + x) = 2tx3 + T 0 (t) + 6t2 x.
2
After cancelations we get T 0 (t) = 3t2 , so T (t) = t3 + c. Then
f (t, x) = t2 x3 + x + 3t2 = t2 (x3 + 1) + x + c.
The solution process is given by Xt = f (t, Wt ) = t2 (Wt3 + 1) + Wt + c. Using X0 = 0 we get
c = 0. Hence the solution is Xt = t2 (Wt3 + 1) + Wt .
The next result deals with a closeness-type condition.
105
Theorem 7.4.1 If the stochastic differential equation (7.4.9) is exact, then the coefficient func-
tions a(t, x) and b(t, x) satisfy the condition
1
∂x a = ∂t b + ∂x2 b. (7.4.12)
2
Proof: If the stochastic equation is exact, there is a function f (t, x) satisfying the system
(7.4.10–7.4.10). Differentiating the first equation of the system with respect to x yields
1
∂x a = ∂t ∂x f + ∂x2 ∂x f.
2
Substituting b = ∂x f yields the desired relation
Remark 7.4.2 The equation (7.4.12) has the meaning of a heat equation. The function b(t, x)
represents the temperature measured at x at the instance t, while ∂x a is the density of heat
sources. The function a(t, x) can be regarded as the potential from which the density of heat
sources is derived by taking the gradient in x.
It worth noting that equation (7.4.12) is a just necessary condition for exactness. This
means that if this condition is not satisfied, then the equation is not exact. In this case we
need to try a different method to solve the equation.
Example 7.4.3 Is the stochastic differential equation
dXt = (1 + Wt2 )dt + (t4 + Wt2 )dWt
exact?
Collecting the coefficients, we have a(t, x) = 1 + x2 , b(t, x) = t4 + x2 . Since ∂x a = 2x, ∂t b = 4t3 ,
and ∂x2 b = 2, the condition (7.4.12) is not satisfied, and hence the equation is not exact.
³ X ´ f (t)dX − X df (t)
t t t
d = .
f (t) f (t)2
For instance, if a stochastic differential equation can be written as
dXt = f 0 (t)Wt dt + f (t)dWt ,
the product rule brings the equation in the exact form
³ ´
dXt = d f (t)Wt ,
dXt = d(tWt2 ),
Since d(e2t ) = 2e2t dt and d(Wt2 ) = 2Wt dWt + dt, the previous relation becomes
Divide by t6
t3 dXt − Xt d(t3 )
= t−5 dt + dWt .
(t3 )2
Applying the quotient rule yields
³X ´ ³ t−4 ´
t
d 3 = −d + dWt .
t 4
Integrating between 1 and t, yields
Xt t−4
= − + Wt − W1 + c
t3 4
so
1
Xt = ct3 − + t3 (Wt − W1 ), c ∈ R.
4t
Using X1 = 0 yields c = 1/4 and hence the solution is
1³ 3 1´
Xt = t − + t3 (Wt − W1 ), c ∈ R.
4 t
Exercise 7.5.1 Solve the following stochastic differential equations by the inspection method
(a) dXt = (1 + Wt )dt + (t + 2Wt )dWt , X0 = 0;
2 3
(b) t dXt = (2t − Wt )dt + tdWt , X1 = 0;
−t/2 1
(c) e dXt = 2 Wt dt + dWt , X0 = 0;
(d) dXt = 2tWt dWt + Wt2 dt,
X0 = 0;
³ ´ √
1
(e) dXt = 1 + 2√t Wt dt + t dWt , X1 = 0.
Integrating yields
Z t Z t
−A(t) −A(s)
e Xt = X0 + e β(s) ds + e−A(s) b(s, Ws ) dWs
0 0
³Z t Z t ´
Xt = X0 eA(t) + eA(t) e−A(s) β(s) ds + e−A(s) b(s, Ws ) dWs .
0 0
The first integral in the previous parenthesis is a Riemann integral, and the latter one is an
Ito stochastic integral. Sometimes, in practical applications these integrals can be computed
explicitly.
When b(t, Wt ) = b(t), the latter integral becomes a Wiener integral. In this case the solution
Xt is Gaussian with mean and variance given by
Z t
E[Xt ] = X0 eA(t) + eA(t) e−A(s) β(s) ds
0
Z t
V ar[Xt ] = e2A(t) e−2A(s) b(s)2 ds.
0
Another important particular case is when α(t) = α 6= 0, β(t) = β are constants and
b(t, Wt ) = b(t). The equation in this case is
¡ ¢
dXt = αXt + β dt + b(t)dWt , t ≥ 0,
Integrating yields
Z t Z t
e−t/2 Xt = X0 + e−s/2 ds + es/2 cos Wt dWs
0 0
is given by Z t
Xt = m + (X0 − m)e−t + α es−t dWs . (7.6.14)
0
Xt is a Gaussian with mean and variance given by
E[Xt ] = m + (X0 − m)e−t
α2
V ar[Xt ] = (1 − e−2t ).
2
Proof: Adding Xt dt to both sides and multiplying by the integrating factor et we get
d(et Xt ) = met dt + αet dWt ,
which after integration yields
Z t
et Xt = X0 + m(et − 1) + α es dWs ,
0
and hence
Z t
Xt = X0 e−t + m − e−t + αe−t es dWs
0
Z t
= m + (X0 − m)e−t + α es−t dWs .
0
Since Xt is the sum between a predictable function and a Wiener integral, using Proposition
4.5.1 it follows that Xt is a Gaussian, with
h Z t i
E[Xt ] = m + (X0 − m)e−t + E α es−t dWs = m + (X0 − m)e−t
0
h Z t i Z t
V arXt = V ar α es−t dWs = α2 e−2t e2s ds
0 0
2t
e −1 1
= α2 e−2t = α2 (1 − e−2t ).
2 2
The variance also tends to zero exponentially, lim V ar[Xt ] = 0. According to Proposition
t→∞
3.3.1, the process Xt tends to m in the mean square sense.
Proposition 7.6.3 (The Brownian Bridge) For a, b ∈ R fixed, the stochastic differential
equation
b − Xt
dXt = dt + dWt , 0 ≤ t < 1, X0 = a
1−t
has the solution
Z t
1
Xt = a(1 − t) + bt + (1 − t) dWs , 0 ≤ t < 1. (7.6.15)
0 1 − s
E[Xt ] = a(1 − t) + bt
V ar[Xt ] = V ar[Ut ] = t(1 − t).
It worth noting that the variance is maximum at the midpoint t = (b − a)/2 and zero at the
end points a and b.
with α constant. This is the equation which is known in physics to model the linear noise.
Dividing by Xt yields
dXt
= α dWt
Xt
Switch to the integral form Z Z
dXt
= α dWt ,
Xt
and integrate “blindly” to get ln Xt = αWt + c, with c integration constant. This leads to the
“pseudo-solution”
Xt = eαWt +c .
The nomination “pseudo” stands for the fact that Xt does not satisfy the initial equation. We
shall find a correct solution by letting the parameter c to be a function of t. In other words,
we are looking for a solution of the following type
where the function c(t) is subject to be determined. Using Ito’s formula we get
dXt = d(eWt +c(t) ) = eαWt +c(t) (c0 (t) + α2 /2)dt + αeαWt +c(t) dWt
= Xt (c0 (t) + α2 /2)dt + αXt dWt .
Substituting the last term from the initial equation (7.7.17) yields
Example 7.7.1 Use the method of variation of parameters to solve the equation
dXt = Xt Wt dWt .
Dividing by Xt convert the differential equation into the equivalent integral form
Z Z
1
dXt = Wt dWt .
Xt
The right side is a well-known stochastic integral given by
Z
W2 t
Wt dWt = t − + C.
2 2
The left side will be integrated “blindly” according to the rules of elementary Calculus
Z
1
dXt = ln Xt + C.
Xt
Equating the last two relations and solving for Xt we obtain the “pseudo-solution”
Wt2
− 2t +c
Xt = e 2 ,
with c constant. In order to get a correct solution, we let c to depend on t and Wt . We shall
assume that c(t, Wt ) = a(t) + b(Wt ), so we are looking for a solution of the form
Wt2
− 2t +a(t)+b(Wt )
Xt = e 2 .
Example 7.7.2 Use the method of variation of parameters to solve the stochastic differential
equation
dXt = µXt dt + σXt dWt ,
with µ and σ constants.
After dividing by Xt we bring the equation into the equivalent integral form
Z Z Z
dXt
= µ dt + σ dWt .
Xt
Integrate on the left “blindly” and get
ln Xt = µt + σWt + c,
where c is an integration constant. We arrive at the following “pseudo-solution”
Xt = eµt+σWt +c .
Assume the constant c is replaced by a function c(t), so we are looking for a solution of the
form
Xt = eµt+σWt +c(t) . (7.7.19)
Apply Ito’s formula and get
¡ σ2 ¢
dXt = Xt µ + c0 (t) + dt + σXt dWt .
2
Subtracting the initial equation yields
¡ σ2 ¢
c0 (t) + dt = 0,
2
2 2
which is satisfied for c0 (t) = − σ2 , with the solution c(t) = − σ2 t + k, k ∈ R. Substituting into
(7.7.19) yields the solution
σ2 σ2 σ2
Xt = eµt+σWt − 2 t+k
= e(µ− 2 )t+σWt +k
= X0 e(µ− 2 )t+σWt
.
Integrating yields
Z t
ρt Xt = ρ0 X0 + r ρs ds
0
Exercise 7.8.1 Let α be a constant. Solve the following stochastic differential equations by the
method of integrating factor
(a) dXt = αXt dWt ;
(b) dXt = Xt dt + αXt dWt ;
1
(c) dXt = dt + αXt dWt , X0 > 0.
Xt
Exercise 7.8.2 Let X R tt be the solution of the stochastic equation dXt = σXt dWt , with σ
constant. Let At = 1t 0 Xs dWs be the stochastic average of Xt . Find the stochastic equation
followed by At and the mean and variance of At .
116
We shall look for a solution of the type Xt = f (Wt ). Ito’s formula yields
1
dXt = f 0 (Wt )dWt + f 00 (Wt )dt.
2
Equating the coefficients of dt and dWt in the last two equations yields
Theorem 7.9.1 (Existence and Uniqueness) Consider the stochastic differential equation
The first condition says that the drift and volatility increase no faster than a linear function
in x. The second conditions states that the functions are Lipschitz in the second argument.
Chapter 8
Martingales
is an Ft -martingale.
117
118
Z t
Example 8.1.2 Let Xt = v(s) dWs be a process as in Example 8.1.1. Then
0
Z t
Mt = Xt2 − v 2 (s) ds
0
is an Ft -martingale.
The process Xt satisfies the stochastic equation dXt = v(t)dWt . By Ito’s formula
is an Ft -martingale.
is an Ft -martingale for 0 ≤ t ≤ T .
Z t Z
1 t 2
Consider the process Ut = u(s) dWs − u (s) ds. Then
0 2 0
119
1
dUt = u(t)dWt − u2 (t)dt
2
(dUt )2 = u(t)dt.
hZ t i £
Z t ¤
E u(τ )Mτ dWτ |Fs = E u(τ )Mτ dWτ = 0,
s s
and hence Z t
E[Mt |Fs ] = E[Ms + u(τ )Mτ dWτ |Fs ] = Ms .
s
Remark 8.1.4 The condition that u(s) is continuous on [0, T ] can be relaxed by asking only
u ∈ L2 [0, T ]. It worth noting that the conclusion still holds if the function u(s) is replaced by a
stochastic process u(t, ω) satisfying Novikov’s condition
£ 1 RT 2 ¤
E e 2 0 u (s,ω) ds < ∞.
The previous process has a distinguished importance in the theory of martingales and it will
be useful when proving the Girsanov theorem.
Definition 8.1.5 Let u ∈ L2 [0, T ] be a deterministic function. Then the stochastic process
Rt Rt
u(s) dWs − 12 u2 (s) ds
Mt = e 0 0
Rt
Let Zt = 0
Ws ds be the integrated Brownian motion. Then
Rt Rt
d dWs − 12 s2 ds
Mt = e 0 0
t3
= etWt − 6 −Zt
is an Ft -martingale.
Example 8.1.6 Let Xt be a solution of dXt = u(t)dt + dWt , with u(s) bounded function.
Consider the exponential process
Rt Rt
u(s) dWs − 12 u2 (s) ds
M t = e− 0 0 .
Then Yt = Mt Xt is an Ft -martingale.
In Example 8.1.3 we obtained dMt = −u(t)Mt dWt . Then
dMt dXt = −u(t)Mt dt.
Product rule yields
dYt = Mt dXt + Xt dMt + dMt dXt
¡ ¢
= Mt u(t)dt + dWt − Xt u(t)Mt dWt − u(t)Mt dt
¡ ¢
= Mt 1 − u(t)Xt dWt .
Integrating between s and t yields
Z t ¡ ¢
Yt = Ys + Mτ 1 − u(τ )Xτ dWτ .
s
Rt ¡ ¢
Since s
Mτ 1 − u(τ )Xτ dWτ is independent of Fs ,
hZ t ¡ ¢ i hZ t ¡ ¢ i
E Mτ 1 − u(τ )Xτ dWτ |Fs = E Mτ 1 − u(τ )Xτ dWτ = 0,
s s
and hence
E[Yt |Fs ] = Ys .
1
Exercise 8.1.7 Prove that (Wt + t)e−Wt − 2 t is an Ft -martingale.
Exercise 8.1.8 Let h be a continuous function. Using the properties of the Wiener integral
and lognormal random variables, show that
h Rt i 1
Rt 2
E e 0 h(s) dWs = e 2 0 h(s) ds .
Exercise 8.1.9 Use the previous exercise to show that for any t > 0
Rt
u(s)2 ds
(a) E[Mt ] = 1 (b) E[Mt2 ] = e 0 .
Exercise 8.1.10 Let Ft = σ{Wu ; u ≤ t}. Show that the following processes are Ft -martingales:
(a) et/2 cos Wt ;
(b) et/2 sin Wt .
Research Topic: Given a function f , find a function g such that g(t)f (Wt ) is an Ft -martingale.
121
We have not used the upper script until now since there was no doubt which probability measure
was used. In this section we shall use also another probability measure given by
dQ = MT dP,
Q(A) > 0, A 6= Ø;
Z
Q(Ω) = MT dP = E P [MT ] = E P [MT |F0 ] = M0 = 1,
Ω
which shows that Q is a probability on F, and hence (Ω, F, Q) becomes a probability space.
The following transformation of expectation from the probability measure Q to P will be useful
in the sequel. If X is a random variable
Z Z
E Q [X] = X(ω) dQ(ω) = X(ω)MT (ω) dP (ω)
Ω Ω
= E P [XMT ].
The following result will play a central role in proving Girsanov’s theorem:
We shall prove this identity by showing that both terms are equal to Xs Ms . Since Xs is
Fs -predictable and Mt is a martingale, the right side term becomes
Let s < t. Using tower property (see Proposition 1.10.4, part 3), the left side term becomes
£ ¤ £ ¤
E P [Xt MT |Fs ] = E P E P [Xt MT |Ft ]|Fs = E P Xt E P [MT |Ft ]|Fs
£ ¤
= E P Xt Mt |Fs = Xs Ms ,
where we used that Mt and Xt Mt are martingales and Xt is Ft -predictable. Hence (8.2.3) holds
and Xt is an Ft -martingale w.r.t. the probability measure Q.
E Q [Xt2 ] = t.
Rt
Proof: Denote U (t) = 0
u(s) ds. Then
From Exercise 8.1.9 (a) we have E P [MT ] = 1. In order to compute E P [Wt MT ] we use the
tower property and the martingale property of Mt
Taking the expectation and using the property of Ito integrals we have
Z t Z t
E[Wt Mt ] = − u(s)E[Ms ] ds = − u(s) ds = −U (t). (8.2.6)
0 0
124
Proof: In order to prove that Xt is a Brownian motion on the probability space (Ω, F, Q)
we shall apply Lévy’s characterization theorem, see Theorem 2.1.4. Hence it suffices to show
that Xt is a Wiener process. Lemma 8.2.1 implies that the process Xt satisfies the following
properties:
1. X0 = 0;
2. Xt is continuous in t;
3. Xt is a square integrable Ft -martingale on the space (Ω, F, Q).
The only property we still need to show is
4. E Q [(Xt − Xs )2 ] = t − s, s < t.
The rest of the proof is dedicated to the verification of relation 4. Using Lemma 8.2.2, the
martingale property of Wt , the additivity and the tower property of expectations, we have
Corollary 8.2.4 Let Wt be a Brownian motion on the probability space (Ω, F, P ). Then the
process
Xt = λt + Wt , 0≤t≤T
is a Brownian motion on the probability space (Ω, F, Q), where
1 2
dQ = e− 2 λ T −λWT
dP.
126
Part II
Applications to Finance
127
Chapter 9
Elementary Calculus provides powerful methods to model an ideal world. However, the real
world is imperfect, and in order to study it, one needs to employ the methods of Stochastic
Calculus.
Given the initial investment M (0) = M0 , the account balance at time t is given by the solution
of the aforementioned equation, which is M (t) = M0 ert .
The model in the real world
In the real world the interest rate r is not constant. It may be assumed constant only for a very
small amount of time, like one day or one week. The interest rate changes unpredictably in time,
which means that it is a stochastic process. This can be modeled in several different ways. For
instance, we may assume that the interest rate at time t is given by the continuous stochastic
process rt = r + σWt , where σ > 0 is a constant that controls the volatility of the rate, and
Wt is a Brownian motion process. The process rt represents a diffusion that starts at r0 = r,
with constant mean E[rt ] = r and variance proportional with the time elapsed, V ar[rt ] = σ 2 t.
With this change in the model, the account balance at time t becomes a stochastic process Mt
that satisfies the stochastic equation
dMt = (r + σWt )Mt dt, t ≥ 0. (9.1.2)
129
130
= M0 ert+σZt ,
Rt
where Zt = 0 Ws ds is the integrated Brownian motion process. Since the moment generating
2 3
function of Zt is m(σ) = E[eσZt ] = eσ t /6 (see Exercise 2.3.3), we obtain
2 3
E[eσWs ] = eσ t /6
;
2 3 2 3
σWs
V ar[e ] = m(2σ) − m(σ) = eσ t /3
(eσ t /3
− 1).
Conclusions
We shall make a few interesting remarks. If M (t) and Mt represent the balance at time t in
the ideal and real worlds, respectively, then
2 3
E[Mt ] = M0 ert eσ t /6
> M0 ert = M (t).
This means that we expect to have more money in the account in the real world rather than
in the idealistic world. It looks that we make money on an investment if the interest rate
is stochastic. In a similar way a bank can expect to make money when lending money at a
stochastic interest rate. This inequality is due to the convexity of the exponential function. If
Xt = rt + σ 2 t3 /6, then Jensen’s inequality yields
Integrating we obtain Z t
e−rt Mt = M0 + α e−rs dWs .
0
Hence the solution is
Z t
rt
Mt = M0 e + α er(t−s) dWs . (9.2.4)
0
This is called the Ornstein-Uhlenbeck process. Since the last term is a Wiener integral, by
Proposition (7.3.1) we have that Mt is Gaussian with the mean
Z t
rt
E[Mt ] = M0 e + E[α er(t−s) dWs ] = M0 ert
0
and variance Z
£ t ¤ α2 2rt
V ar[Mt ] = V ar α er(t−s) dWs = (e − 1).
0 2r
It worth noting that the expected balance is equal to the real world balance M0 ert . The variance
for t small is approximately equal to α2 t.
If the constant α is replaced by an unpredictable function α(t, Wt ), the equation becomes
This process is not Gaussian. Its mean and variance are given by
E[Mt ] = M0 ert
Z t
V ar[Mt ] = e2r(t−s) E[α2 (t, Wt )] ds.
0
√ √
In the particular case when α(t, Wt ) = e 2rWt , using Application 6.0.18 with λ = 2r, we
can work out for (9.2.5) an explicit form of the solution
Z t √
rt
Mt = M0 e + er(t−s) e 2rWt dWs
0
Z t √
rt rt
= M0 e + e e−rs e2 rWt dWs
0
rt 1 ¡ −rs+√2rWt
rt
¢
= M0 e + e √ e −1
2r
1 ¡ √ ¢
= M0 ert + √ e 2rWt − ert .
2r
The previous model for the interest rate is not a realistic one because it allows for negative
rates. The Brownian motion hits −r after a finite time with probability 1. However, it might
be a good stochastic model for the evolution of population, since in this case the rate can be
negative.
132
1.25
1.24
1.23
1.22
1.21
Equilibrium level
1000 2000 3000 4000 5000 6000
Figure 9.1: A simulation of drt = a(b − rt )dt + σdWt , with r0 = 1.25, a = 3, σ = 1%, b = 1.2.
The drift rate and volatility of the spot rate rt do not depend explicitly on the time t. There
are several classical choices for m(rt ) and σ(rt ) that will be discuses in the sequel.
with a, b, σ constants.
Assuming the spot rates are deterministic, we take σ = 0 and obtain the ODE
rt = b + (r0 − b)e−at .
This implies that the rate rt is pulled towards level b at the rate a. This means that, if r0 > b,
then rt is decreasing towards b, and if r0 < b, then rt is increasing towards the horizontal
asymptote b. The term σdWt in Vasicek’s model adds some “white noise” to the process. In
the following we shall find an explicit formula for the spot rate rt in the stochastic case.
Since the spot rate rt is the sum between the predictable function r0 e−at + be−at (eat − 1) and
a multiple of a Wiener integral, from Proposition (4.5.1) follows that rt is Gaussian, with
12.0
11.5
10.5
Vasicek's model
Equilibrium level
10.0
Figure 9.2: Comparison between a simulation in CIR and Vasicek models, with parameter
values a = 3, σ = 15%, r0 = 12, b = 10. Note that CIR process tends to be more volatile than
Vasicek’s.
Proof: Since lim E[rt ] = lim (b+(r0 −b)e−at ) = b and lim V ar[rt ] = 0, applying Proposition
t→∞ t→∞ t→∞
3.3.1 we get ms-lim rt = b.
t→∞
Since rt is normally distributed, the Vasicek’s model has been criticized because it allows for
negative interest rates and unbounded large rates. See Fig.9.1 for a simulation of the short-term
interest rates for the Vasicek model.
Exercise 9.4.3 Find the probability that rt is negative. What happens with this probability
when t → ∞? Find the rate of change of this probability.
yields the differential equation µ (t) = ab − aµ(t). Multiply by the integrating factor eat and
0
obtain
d(eat µ(t)) = abeat .
Integrating and using that µ(0) = r0 provides the solution
Hence ¡ ¢
lim E[rt ] = lim b + e−at (r0 − b) = b,
t→∞ t→∞
Differentiate again
µ02 (t) = (2ab + σ 2 )µ(t) − 2aµ2 (t).
Solving as a linear differential equation in µ2 (t) yields
Exercise 9.4.4 Use a similar method to find a recursive formula for the moments of a CIR
process.
136
In this model θ(t) is the average direction in which rt moves and it is considered independent
of rt , while σ is the standard deviation of the short rate The solution process is Gaussian and
is given by
Z t
rt = r0 + θ(s) ds + σWt .
0
If F (0, t) denotes the forward rate at time t as seen at time 0, it is known that θ(t) = Ft (0, t) +
σ 2 t. In this case the solution becomes
σ2
rt = r0 + F (0, t) + t + σWt .
2
with a and σ constants. We can solve the equation multiplying by the integrating factor eat
Since the first two terms are deterministic and the last is a Wiener integral, the process rt is
Gaussian.
The θ(t) function can be calculated from the term structure as
σ2
θ(t) = ∂t F (0, t) + aF (0, t) + (1 − e−2at ).
2a
Then
Z t Z t Z t Z t
as as as σ2
θ(s)e ds = ∂s F (0, s)e ds + a F (0, s)e ds + (1 − e−2as )eas ds
0 0 0 2a 0
σ2 ¡ ¢
= F (0, t)eat − r0 + cosh(at) − 1 ,
a2
137
This implies that ut is Gaussian and hence rt = eut is log-normal for each t. Using u0 = ln r0
and σ(t)
u0 σ(t)
e σ(0) σ(t) = e σ(0) ln r0 = r0σ(0) ,
we obtain the following explicit formula for the spot rate
σ(t) Rt θ(s)
rt = r0σ(0) eσ(t) 0 σ(s)
ds σ(t)Wt
e .
Since σ(t)Wt is normally distributed with mean 0 and variance σ 2 (t)t, the log-normal variable
eσ(t)Wt has
2
E[eσ(t)Wt ] = eσ (t)t/2
2 2
V ar[eσ(t)Wt ] = eσ(t) t (eσ(t) t − 1).
Hence
σ(t) Rtθ(s)
ds σ 2 (t)t/2
E[rt ] = r0σ(0) eσ(t) 0 σ(s) e
2σ(t) Rt θ(s)
ds σ(t)2 t 2
V ar[rt ] = r0σ(0)
e2σ(t) 0 σ(s) e (eσ(t) t − 1).
The price of a stock can be modeled by a continuous stochastic process which is the sum
between a predictable and an unpredictable part. However, this type of model does not take
into account market crashes. If those are to be taken into consideration, the stock price needs
to contain a third component which models unexpected jumps. We shall discuss these models
in the following.
The parameters µ and σ are positive constants which represent the drift and volatility of the
stock. This equation has been solved in Example 7.7.2 by applying the method of variation of
parameters. The solution is
σ2
St = S0 e(µ− 2 )t+σWt
, (10.1.2)
where S0 denotes the price of the stock at time t = 0. It worth noting that the stock price is
Ft -adapted, positive, and it has a log-normal distribution. Using Exercise 1.6.1, the mean and
variance are
E[St ] = S0 eµt (10.1.3)
2
V ar[St ] = S02 e2µt (eσ t − 1). (10.1.4)
139
140
1.15
1.10
1.05
Figure 10.1: Two distinct simulations for the stochastic equation dSt = 0.15St dt + 0.07St dWt ,
with S0 = 1.
Exercise 10.1.1 Let Fu be the information set at time u. Find E[St |Fu ] and V ar[St |Fu ] for
u ≤ t.
Exercise 10.1.2 Find the stochastic process followed by ln St . What are the values of E[ln(St )]
and V ar[ln(St )]? What happens when s = t?
Exercise 10.1.3 Find the stochastic differential equations associated with the following pro-
cesses
1
(a) (b) Stn (c) ln St (d) (St − 1)2 .
St
2
Exercise 10.1.4 Show that E[St2 ] = S02 e(2µ+σ )t
. Find a similar formula for E[Stn ], with n
positive integer.
The next result deals with the probability that the stock price reaches certain barrier before
another barrier.
Theorem 10.1.6 Let Su and Sd be fixed, such that Sd < S0 < Su . The probability that the
stock price St hits the upper value Su before the lower value Sd is
dγ − 1
p= ,
dγ − uγ
where Su /S0 = u, Sd /S0 = d, and γ = 1 − 2µ/σ 2 .
Proof: Let Xt = mt + Wt . Theorem 3.1.13 provides
e2mβ − 1
P (Xt goes up to α before down to − β) = . (10.1.5)
e2mβ − e−2mα
141
dγ − 1
P (St goes up to Su before down to Sd ) = ,
dγ − uγ
³ S ´1− 2µ2
0 σ
P (St hits Su ) = .
Su
We shall solve the equation using the method of integrating factors presented in section 7.8.
Multiplying by the integrating factor
Rt Rt
σ(s) dWs + 12 σ 2 (s) ds
ρt = e − 0 0
the equation (10.2.7) becomes d(ρt St ) = ρt µ(t)St dt. Substituting Yt = ρt St yields the deter-
ministic equation dYt = µ(t)Yt dt with solution
Rt
µ(s) ds
Yt = Y0 e 0 .
Proposition 10.2.1 The solution St is Ft -adapted and log-normally distributed, with mean
and variance given by
Rt
µ(s) ds
E[St ] = S0 e 0
¡ Rt 2
Rt ¢
V ar[St ] = S02 e2 0 µ(s) ds e 0 σ (s) ds − 1 .
Rt Rt
Proof: Let Xt = 0 (µ(s) − 12 σ 2 (s)) ds + 0 σ(s) dWs . Since Xt is a sum between a predictable
integral function and a Wiener integral, it is normally distributed, see Proposition 4.5.1, with
Z t
¡ 1 ¢
E[Xt ] = µ(s) − σ 2 (s) ds
0 2
hZ t i Z t
V ar[Xt ] = V ar σ(s) dWs = σ 2 (s) ds.
0 0
Using Exercise 1.6.1, the mean and variance of the log-normal random variable St = S0 eXt are
given by
Rt 2 Rt Rt
(µ− σ2 ) ds+ 12 σ 2 ds
E[St ] = S0 e 0 = S0 e 0 µ(s) ds
0
Rt 1 2
Rt 2 ¡ Rt 2 ¢
V ar[St ] = S02 e2 0 (µ− 2 σ ) ds+ 0 σ ds e 0 σ − 1
Rt ¡ Rt 2 ¢
= S02 e2 0 µ(s) ds e 0 σ (s) ds − 1 .
E[St ] = S0 eµt
2
V ar[St ] = S02 e2µt (eσ t − 1).
It follows that the continuously sampled arithmetic average of stock prices between 0 and t is
given by
Z
1 t
At = Su du.
t 0
This integral can be computed explicitly just in the case µ = 0, see Application 6.0.18. It worth
noting that At is neither normal nor log-normal, fact which makes the price of Asian options
on arithmetic averages hard to evaluate.
144
Rt
Let It = 0 Su du. By the Fundamental Theorem of Calculus implies dIt = St dt. Then the
quotient rule yields
³I ´ dIt t − It dt St t dt − It dt 1
t
dAt = d = 2
= 2
= (St − At )dt,
t t t t
i.e. the continuous arithmetic average At satisfies
1
dAt = (St − At )dt.
t
If At < St , the right side is positive and hence dAt > 0, i.e. the average At goes up. Similarly,
if At > St , then the average At goes down. This shows that the average At tends to trace the
stock values St .
By L’Hospital’s rule we have
It
A0 = lim
= lim St = S0 .
t t→0
t→0
S02 2
E[It St ] = [e(2µ+σ )t − eµt ].
µ + σ2
with the initial condition g(0) = 0. This can be solved as a linear differential equation in g(t)
multiplying by the integrating factor e−µt . The solution is
S02 2
g(t) = 2
[e(2µ+σ )t − eµt ].
µ+σ
(ii) Since At and It are proportional, it suffices to show that It and St are not independent.
This follows from part (i) and the fact
S02 µt
E[It St ] 6= E[It ]E[St ] = (e − 1)eµt .
µ
Next we shall find E[It2 ]. Using dIt = St dt, then (dIt )2 = 0 and hence Ito’s formula yields
eµt − 1
E[At ] = S0 , t>0
µt
( )
2 h e(2µ+σ )t − 1 eµt − 1 i (eµt − 1)2
2
S02
V ar[At ] = − − .
t2 µ + σ 2 2µ + σ 2 µ µ2
Exercise 10.3.3 Find approximative formulas for E[At ] and V ar[At ] for t small, up to the
order O(t2 ).
Therefore the continuously sampled geometric average of stock prices between instances 0 and
t is given by
1
Rt
ln Su du
Gt = e t 0 . (10.3.12)
Theorem 10.3.4 Gt has a log-normal distribution, with the mean and variance given by
σ2
) 2t
E[Gt ] = S0 e(µ− 6
σ2 ¡ σ2 t ¢
V ar[Gt ] = S02 e(µ− 6 )t
e 3 −1 .
Proof: Using
³ σ2
´ σ2
ln Su = ln S0 e(µ− 2 )u+σWu = ln S0 + (µ − )u + σWu ,
2
then taking the logarithm yields
Z th i Z t
1 σ2 σ2 t σ
ln Gt = ln S0 + (µ − )u + σWu du = ln S0 + (µ − ) + Wu du.
t 0 2 2 2 t 0
Rt
Since the integrated Brownian motion Zt = 0
Wu du is Gaussian with Zt ∼ N (0, t3 /3), it
follows that ln Gt has a normal distribution
³ σ2 t σ2 t ´
ln Gt ∼ N ln S0 + (µ − ) , . (10.3.13)
2 2 3
147
This implies that Gt has a log-normal distribution. Using Exercise 1.6.1, we obtain
1 σ2 t σ2 t σ2
) 2t
E[Gt ] = eE[ln Gt ]+ 2 V ar[ln Gt ] = eln S0 +(µ− 2 ) 2 + 6 = S0 e(µ− 6 .
¡ ¢
V ar[Gt ] = e2E[ln Gt ]+V ar[ln Gt ] eV ar[ln Gt ] − 1
σ2 ¡ σ2 t ¢ σ2 ¡ σ2 t ¢
= e2 ln S0 +(µ− 6 )t e 3 − 1 = S02 e(µ− 6 )t e 3 − 1 .
σ2
Rt
) 2t + σt
Gt = S0 e(µ− 2 0
Wu du
.
An important consequence of the fact that Gt is log-normal is that Asian options on geometric
averages have closed-form solutions.
Exercise 10.3.6 (a) Show that ln Gt satisfies the stochastic differential equation
1³ ´
d(ln Gt ) = ln St − ln Gt dt.
t
(b) Show that Gt satisfies
1 ³ ´
dGt = Gt ln St − ln Gt dt.
t
Consider tk = kt
n . Then the continuously sampled harmonic average is obtained by taking the
limit n → ∞ in the aforementioned relation
n t t
lim n = lim n =Z t ,
n→∞ X 1 n→∞ X 1 t 1
du
S(tk ) S(tk ) n 0 Su
k=1 k=1
t
Ht = Z t .
1
du
0 Su
Z t
t 1
We may also write Ht = , where It = du satisfies
It 0 Su
1 ³1´ 1
dIt = dt, I0 = 0, d =− dt.
St It St It2
148
Ht ≤ Gt ≤ At .
Exercise 10.3.10 The stochastic average of stock prices between 0 and t is defined by
Z
1 t
Xt = Su dWu ,
t 0
where Wu is a Brownian motion process.
(i) Find dXt , E[Xt ] and V ar[Xt ].
St − S0
(ii) Show that σXt = Rt − µAt , where Rt = is the “raw average” and At =
Rt t
1
t 0
Su du is the continuous arithmetic average.
149
• the Brownian motion process Wt , which models regular events given by infinitesimal
changes in the price, and which is a continuous process;
• the Poisson process Nt , which is discontinuous and models sporadic jumps in the stock
price that correspond to shocks in the market.
Since E[dNt ] = λdt, the Poisson process Nt has a positive drift and we need to “compensate”
it by subtracting λt from Nt . The resulting process Mt = Nt − λt is a martingale, called the
compensated Poisson process, that models unpredictable jumps of size 1 at a constant rate λ.
It worth noting that the processes Wt and Mt involved in modeling the stock price are assumed
to be independent.
dSt
To set up the model, we assume the instantaneous return on the stock, , to be the sum
St
of the following three components:
• the predictable part µdt;
• the noisy part due to unexpected news σdWt ;
• the rare events part due to unexpected jumps ρdMt ,
where µ, σ and ρ are constants, corresponding to the drift rate of the stock, volatility and jump
size.
Adding yields
dSt
= µdt + σdWt + ρdMt .
St
Hence the dynamics of a stock price, subject to rare events, is modeled by the following stochas-
tic differential equation
dSt = µSt dt + σSt dWt + ρSt dMt . (10.4.15)
It worth noting that in the case of zero jumps, ρ = 0, the previous equation becomes the
classical stochastic equation (10.1.1).
Using that Wt and Mt are martingales, we have
This shows the unpredictability of the last two terms, i.e. given the information set Ft at time
t, it is not possible to predict any future increments in the next interval of time dt. The term
σSt dWt captures regularly events of insignificant size, while ρSt dMt captures rare events of
large size. The “rare events” term, ρSt dMt , incorporates jumps proportional with the stock
price and is given in terms of the Poisson process Nt as
The constant λ represents the rate at which the jumps of the Poisson process Nt occur. This is
the same as the rate of rare events in the market, and can be determined from historical data.
The following result provides an explicit solution for the stock price when rare events are taken
into account.
where
µ is the stock price drift rate;
σ is the volatility of the stock;
λ is the rate at which rare events occur;
ρ is the size of jump in the expected return when a rare events occur.
Proof: We shall look for a solution of the type St = S0 eΦ(t,Wt ,Mt ) , with Φ(t, Wt , Mt ) =
α(t) + β(Wt ) + γ(Mt ). Using (dMt )2 = dNt = dMt + λdt, see (2.8.13), applying Ito’s formula
we have
1 1 λ λ
µ = α0 (t) + β 0 (Wt )2 + β 00 (Wt ) + γ 0 (Mt )2 + γ 00 (Mt )
2 2 2 2
σ = β 0 (Wt )
1 1
ρ = γ 0 (Mt ) + γ 0 (Mt )2 + γ 00 (Mt )
2 2
From the second equation we get β(Wt ) = σWt , and substituting in the first one yields
1
α0 (t) = µ − σ 2 − λρ + λγ 0 (Mt ).
2
Since the left side term is a function of t and the right side is a function of Mt , it must be a
separation constant C such that
1
α0 (t) = C = µ − σ 2 − λρ + λγ 0 (Mt ). (10.4.16)
2
151
Then α(t) = Ct. In the following we shall find the value of the constant C. The second identity
of (10.4.16) yields γ 0 (Mt ) = κ, constant. Then the aforementioned formula for ρ yields
1
ρ = κ + κ2 .
2
Multiplying by 2 and completing the square we have
p
(κ + 1)2 = 1 + 2ρ =⇒ κ1,2 = −1 ± 1 + 2ρ.
Substituting in (10.4.16) yields
1 p
C1,2 = µ − σ 2 − λρ + λ(−1 ± 1 + 2ρ)
2
1 p
= µ − σ 2 + λ[± 1 + 2ρ − ρ − 1].
2
In order to pick the correct sign, if let ρ = 0, i.e. if assume the jumps have size zero, we need
to get α(t) = (µ − 12 σ 2 )t. This implies the positive sign in front of the square root. Then
1 p
C = µ − σ 2 + λ[ 1 + 2ρ − ρ − 1].
2
√
It worth noting that 1 + 2ρ − ρ − 1 > 0, and C > 0. It follows that
³ 1 p ´
α(t) = µ − σ 2 + λ[ 1 + 2ρ − ρ − 1] t
2
β(Wt ) = σWt
p
γ(Mt ) = ( 1 + 2ρ − 1)Mt ,
where we choose α(0) = β(0) = γ(0) = 0. Hence
³ 1 p ´ p
Φ(t, Wt , Mt ) = µ − σ 2 + λ[ 1 + 2ρ − ρ − 1] t + σWt + ( 1 + 2ρ − 1)Mt .
2
Risk-Neutral Valuation
The rate r is considered constant, but the method can be easily adapted for time dependent
rates.
In the following we shall present explicit computations for the most common European type
derivatives prices using the risk-neutral valuation method.
If let x = ln(ST /St ), using the log-normality of the stock price in a risk-neutral world
σ2
ST = St e(r− 2 )(T −t)+σWT −t
,
153
154
with notations
Z ∞ Z ∞
I1 = Kp(x) dx, I2 = St ex p(x) dx
ln(K/St ) ln(K/St )
2
x−(r− σ2 )(T −t)
With the substitution y = √
σ T −t
, the first integral becomes
Z ∞ Z ∞
1 y2
I1 = K p(x) dx = K √ e− 2 dy
ln(K/St ) −d2 2π
Z d2
1 y2
= K √ e− 2 dy = KN (d2 ),
−∞ 2π
where 2
ln(St /K) + (r − σ2 )(T − t)
d2 = √ ,
σ T −t
and Z u
1 2
N (u) = √ e−z /2
dz
2π −∞
where 2
√ ln(St /K) + (r + σ2 )(T − t)
d1 = d2 + σ T − t = √ .
σ T −t
Substituting back to (11.2.2) and then into (11.2.1) yields
c(t) = e−r(T −t) (I2 − I1 ) = e−r(T −t) [St er(T −t) N (d1 ) − KN (d2 )]
= St N (d1 ) − Ke−r(T −t) N (d2 ).
11.3 Cash-or-nothing
A financial security that pays 1 dollar if the stock price ST ≥ K and 0 otherwise, is called bet
contract, or cash-or-nothing contract. The payoff can be written as
½
1, if ST ≥ K
fT =
0, if ST < K.
Exercise 11.3.1 Let 0 < K1 < K2 . Find the price of a financial derivative which pays at
maturity $1 if K1 ≤ St ≤ K2 and zero otherwise. This is a “box-bet” and its payoff is given by
½
1, if K1 ≤ ST ≤ K2
fT =
0, otherwise.
Exercise 11.3.3 Find the price at time t of a derivative which pays at maturity
½ n
ST , if ST ≥ K
fT =
0, otherwise.
11.4 Log-contract
The financial security that pays at maturity fT = ln ST is called a log-contract. Since the stock
is log-normal distributed,
³ σ2 ´
ln ST ∼ N ln St + (µ − )(T − t), σ 2 (T − t) ,
2
the risk-neutral expectation at time t is
2
bt [fT ] = E[ln ST |Ft , µ = r] = ln St + (r − σ )(T − t),
E
2
and hence the price of the log-contract is given by
³ 2 ´
bt [fT ] = e−r(T −t) ln St + (r − σ )(T − t) .
ft = e−r(T −t) E
2
Exercise 11.4.1 Find the price at time t of a square log-contract whose payoff is given by
fT = (ln ST )2 .
11.5 Power-contract
The financial derivative which pays at maturity the nth power of the stock price, STn , is called
a power contract. Since ST has a log-normal distribution with
σ2
ST = St e(µ− 2 )(T −t)+σWT −t
,
the nth power of the stock, STn , is also log-normal distributed, with
σ2
fT = STn = Stn en(µ− 2 )(T −t)+nσWT −t
.
Then the expectation at time t in the risk-neutral world is
σ2
E[fT |Ft , µ = r] = Stn en(r− 2 )(T −t)
E[enσWT −t ]
σ2 2
1
σ 2 (T −t)
= Stn en(r− 2 )(T −t)
e2n
157
where hi,t is the price at time t of a derivative that pays at maturity hi,T . We shall successfully
use this method in the situation when the payoff fT can be decomposed in simpler payoffs, for
which we can evaluate directly the price of the associate derivative. In this case the price of
the initial derivative, ft , is obtained as a combination of the prices of the more easy to valuate
derivatives.
The reason underlying the aforementioned superposition principle is the linearity of the
expectation operator Ê
n
X
ft = b T ] = e−r(T −t) E[
e−r(T −t) E[f b ci hi,T ]
i=1
n
X n
X
= e−r(T −t) b i,T ] =
ci E[h ci hi,t .
i=1 i=1
158
This principle is also connected with the absence of arbitrage opportunities in the market.
Consider two portfolios of derivatives with equal values at the maturity time T
n
X m
X
ci hi,T = aj gj,T .
i=1 j=1
If think this common value to be the payoff of a derivative, fT , then by the aforementioned
principle, the portfolios have the same value at any prior time t ≤ T
n
X m
X
ci hi,t = aj gj,t .
i=1 j=1
The last identity can also result from the absence of arbitrage opportunities in the market. If
there is a time t at which the identity fails, then buying the cheaper portfolio and selling the
more expensive one will lead to an arbitrage profit.
The superposition principle can be used to price package derivatives such as spreads, strad-
dles, strips, straps and strangles. We shall deal with these in the sequel.
with ½ ½
ST , if x ≥ K 1, if x ≥ K
h1,T = h2,T =
0, if x < K, 0, if x < K.
These are the payoffs of an asset-or-nothing and of a cash-or-nothing derivatives. From section
11.3 and Exercise 11.3.2 we have h1,t = St N (d1 ), h2,t = e−r(T −t) N (d2 ). By superposition we
get the price of a call at time t
h2,t = St N (−d1 ), t ≤ T.
pT = Kh1,T − h2,T
and use the superposition principle to find the price pt of the put.
159
fT = AT − K.
For instance, if the asset is electric energy or natural gas, it makes sense to make a deal
on the average price of the asset, since their prices are volatile and sometimes can become
unaffordable expensive during the winter season.
The risk-neutral expectation E bt [AT ] is obtained from formula (10.3.9) by letting µ = r and
replacing t by T − t and S0 by St . Since K is a constant, the price of the contract at time t is
given by
It worth noting that the price of an asian forward contract is always cheaper than the price
of an usual forward contract on the asset. To see that, we substitute x = r(T − t) in the
inequality e−x > 1 − x, x > 0, to get
1 − e−r(T −t)
< 1.
r(T − t)
1 − e−r(T −t)
St − e−r(T −t) K < St − e−r(T −t) K.
r(T − t)
Since the left side is the price of an asian forward contract, while the right side is the price of
an usual forward contract, we obtain the desired inequality.
Exercise 11.9.1 Find the price of a contract that pays at maturity date the difference between
the asset price and its arithmetic average, fT = ST − AT .
For theoretical purposes, one may consider asian forward contracts on geometric average.
This is a derivative that pays at maturity the difference fT = GT − K, where GT is the
continuous geometric average of the asset price between 0 and T and K is a fixed delivery
price.
Substituting µ = r, S0 by St and t by T − t in the first relation provided by Theorem 10.3.4,
the risk-neutral expectation of GT as of time t is
2
bt [GT ] = St e 21 (r− σ6
E )(T −t)
.
160
The value at time t of a derivative that pays at maturity the geometric average, gT = GT , is
1 σ2
bt [GT ] = St e− 2 (r+
gt = e−r(T −t) E 6 )(T −t)
.
Then the value of the forward contract on the geometric average becomes
1 σ2
ft = gt + e−r(T −t) K = St e− 2 (r+ 6 )(T −t)
− e−r(T −t) K. (11.9.3)
Exercise 11.9.2 Show that the contract (11.9.3) is cheaper than a usual forward contract on
the asset.
Lemma 11.10.1 The value at time t of a derivative, which pays at maturity $1 if the geometric
average GT ≥ K and 0 otherwise, is given by
where 2
ln St − ln K + (µ − σ2 ) T 2−t
de2 = p .
σ (T − t)/3
which was obtained by replacing S0 by St and t by T − t in formula (10.3.13). Let p(x) be the
probability density of the random variable XT
1 σ2
) T 2−t ]2 /(
2σ 2 (T −t)
p(x) = √ p e−[x−ln St −(µ− 2 3 )
. (11.10.4)
2πσ (T − t)/3
yields
Z ∞ Z df2
bt [hT ] = 1 2 1 2
E √ e−y /2 dy = √ e−y /2 dy
2π −df2 2π −∞
= N (de2 ),
where 2
ln St − ln K + (r − σ2 ) T 2−t
de2 = p .
σ (T − t)/3
Discounting to the free interest rate yields the price at time t
bt [hT ] = e−r(T −t) N (de2 ).
ht = e−r(T −t) E
The following result deals with the price of a geometric average-or-nothing derivative.
Lemma 11.10.2 The value at time t of a derivative, which pays at maturity GT if GT ≥ K
and 0 otherwise, is given by the formula
σ2
St N (de1 ),
1
gt = e− 2 (r+ 6 )(T −t)
where 2
ln St − ln K + (r + σ6 ) T 2−t
de1 = p .
σ (T − t)/3
Proof: Since the payoff can be written as
½ ½ X
GT , if GT ≥ K e T, if XT ≥ ln K
gT = =
0, if GT < K 0, if XT < ln K,
Z ∞ Z ∞
bt [g ] =
E gT (x)p(x) dx = ex p(x) dx,
T
−∞ ln K
where p(x) is given by (11.10.4), with µ is replaced by r. Using the substitution (11.10.5) and
completing the square yields
Z ∞
bt [g ] = 1 σ2
) T 2−t ]2 /(
2σ 2 (T −t)
E T
√ p ex e−[x−ln St −(r− 2 3 )
dx
2πσ (T − t)/3 ln K
Z ∞ √
1 1 σ2 1 2
= √ St e 2 (r− 6 )(T −t) e− 2 [y−σ (T −t)/3] dy
2π f
−d2
If let
p
de1 = de2 + σ (T − t)/3
2
ln St − ln K + (r − σ2 ) T 2−t p
= p + σ (T − t)/3
σ (T − t)/3
2
ln St − ln K + (r + σ6 ) T 2−t
= p ,
σ (T − t)/3
p
the previous integral becomes after substituting z = y − σ (T − t)/3
Z ∞
1 σ2 1 2 σ2
e− 2 z dz = St e 2 (r− 6 )(T −t) N (de1 )
1 1
√ St e 2 (r− 6 )(T −t)
2π f
−d 1
The value of the derivative at time t is obtained by discounting at the interest rate r
σ2
N (de1 )
1
gt = bt [g ] = e−r(T −t) St e 2 (r−
e−r(T −t) E 6 )(T −t)
T
2
= e − 21 (r+ σ6 )(T −t)
St N (de1 ).
Proposition 11.10.3 The value at time t of a geometric average price call option is
σ2
St N (de1 ) − Ke−r(T −t) N (de2 ).
1
ft = e− 2 (r+ 6 )(T −t)
fT = gT − KhT ,
with ½ ½
GT , if GT ≥ K 1, if GT ≥ K
gT = hT =
0, if GT < K, 0, if GT < K,
163
applying the superposition principle and Lemmas 11.10.1 and 11.10.2 yields
ft = gt − Kht
σ2
St N (de1 ) − Ke−r(T −t) N (de2 ).
1
= e− 2 (r+ 6 )(T −t)
Exercise 11.10.4 Find the value at time t of a geometric average price put option.
Arithmetic average price options There is no simple closed-form solution for a call or
for a put on the arithmetic average At . However, there is an approximate solution based on
computing exactly the first two moments of the distribution of At , and applying the risk-neutral
valuation assuming that the distribution is log-normal with the same two moments. This idea
was developed by Turnbull and Wakeman (1991), and it works pretty well for volatilities up to
about 20%.
The following result provides the mean and variance of a normal distribution in terms of
the first two moments of the associated log-normal distribution.
Proposition 11.10.5 Let Y be a log-normal distributed random variable, having the first two
moments given by
m1 = E[Y ], m2 = E[Y 2 ].
Then ln Y has the normal distribution ln Y ∼ N (µ, σ 2 ), with
m2 m2
µ = ln √ 1 , σ 2 = ln · (11.10.6)
m2 m21
Proof: Using Exercise 1.6.1 we have
σ2 2
m1 = E[Y ] = eµ+ 2 , m2 = E[Y 2 ] = e2µ+2σ .
eµT − 1
m1 = E[IT ] = S0 ;
µ
2S02 h e(2µ+σ )T − 1 eµT − 1 i
2
m2 = E[IT2 ] = − .
µ + σ2 2µ + σ 2 µ
using Proposition 11.10.5 it follows that ln At is normal distributed, with
³ m2 m2 ´
ln AT ∼ N ln √ 1 − ln t, ln 2 . (11.10.7)
m2 m1
164
Relation (11.10.7) represents the normal approximation of ln At . We shall price the arithmetic
average price call under this condition.
In the next two exercises we shall assume the distribution of AT given by the log-normal
distribution (11.10.7).
Exercise 11.10.6 Using a method similar with the one used in Lemma 11.10.1, show that an
approximate value at time 0 of a derivative, which pays at maturity $1 if the arithmetic average
AT ≥ K and 0 otherwise, is given by
h0 = e−rT N (dˇ2 ),
with √
ln(m21 / m2 ) − ln K − ln t
dˇ2 = p , (11.10.8)
ln(m2 /m21 )
where in the expressions of m1 and m2 we replaced µ by r.
Exercise 11.10.7 Using a method similar with the one used in Lemma 11.10.2, show that the
approximate value at time 0 of a derivative, which pays at maturity AT if AT ≥ K and 0
otherwise, is given by the formula
1 − e−rT
a0 = S0 N (dˇ1 ),
r
where
m2
ln √m12 + ln t − ln K
dˇ1 = ln(m2 /m21 ) + p , (11.10.9)
ln(m2 /m21 )
where in the expressions of m1 and m2 we replaced µ by r.
In the particular case, when the jump size is ρ = 0, or the rate of occurring of jumps λ = 0, we
obtain the familiar result
ft = St − e−r(T −t) K.
166
Chapter 12
Martingale Measures
Heuristically speaking, this means that given all information in the market at time u, Fu , the
expected price of any future stock price is the price of the stock at time u, i.e Su . This does
not make any sense, since in this case the investor would prefer investing the money in a bank
at risk-free interest rate, rather than buying a stock with zero return. Then (12.1.1) does not
hold. The next result shows how to fix this problem.
167
168
Since St is Ft -predictable, so will be e−µt St . Using formula (10.1.2), taking out the predictable
part yields
1 2
E[St |Fu ] = E[S0 e(µ− 2 σ )t+σWt
|Fu ]
(µ− 12 σ 2 )u+σWu (µ− 12 σ 2 )(t−u)+σ(Wt −Wu )
= E[S0 e e |Fu ]
(µ− 21 σ 2 )(t−u)+σ(Wt −Wu )
= E[Su e |Fu ]
(µ− 12 σ 2 )(t−u) σ(Wt −Wu )
= Su e E[e |Fu ]. (12.1.3)
which is equivalent to
E[e−µt St |Fu ] = e−µu Su .
The conditional expectation E[St |Fu ] can be expressed in terms of the conditional density
function as Z
E[St |Fu ] = St p(St |Fu ) dSt , (12.1.4)
This way, dP (x) = p(x|Fu ) dx becomes a martingale measure for e−µt St . Since the rate of
return µ might not be known from the beginning, and it depends on each particular stock, a
meaningful question would be:
Under what martingale measure does the discounted stock price, Mt = e−rt St , become a mar-
tingale?
The constant r denotes, as usual, the risk-free interest rate.
169
where Eb denotes the expectation with respect to the requested martingale measure. The
previous relation can be also written as
b t |Fu ] = Su ,
e−r(t−u) E[S u < t.
This states that the discounted expectation at the risk-free interest rate for the time interval
t − u is the price of the stock, Su . Since this does not involve any of the riskiness of the stock,
we might think of it as an expectation in the risk-neutral world. The aforementioned formula
can be written in the compound mode as
b t |Fu ] = Su er(t−u) ,
E[S u < t. (12.1.5)
This can be obtained from the conditional expectation E[St |Fu ] = Su eµ(t−u) by substituting
µ = r and replacing E by E, b which corresponds to the definition of the expectation in a
risk-neutral world. Therefore, the evaluation of derivatives in section 11 is done by using
the aforementioned martingale measure under which e−rt St is a martingale. Next we shall
determine this measure explicitly.
can be written as 2
1
St = S0 eµt eσWt − 2 σ t .
1 2
Then e−µt St = S0 eσWt − 2 σ t is an exponential process. By Example 8.1.3, particular case 1,
this process is an Ft -martingale, where Ft = σ{Wu ; u ≤ t} is the information available in the
market until time t. Hence e−µt St is a martingale, which is a result proved also by Proposition
12.1.1. The probability space where this martingale lives is (Ω, F, P ).
In the following we shall change the rate of return µ into the risk-free rate r and change the
probability measure such that the discounted stock price becomes a martingale. The discounted
stock price can be expressed in terms of the Brownian motion with drift
ct = µ − r t + Wt
W (12.1.6)
σ
as in the following
1 2 c 1 2
e−rt St = e−rt S0 eµt eσWt − 2 σ t
= e σ Wt − 2 σ t .
µ−r ct is a Brownian
If let λ = in Corollary 8.2.4 of Girsanov’s theorem, it follows that W
σ
motion on the probability space (Ω, F, Q), where
1 µ−r 2
dQ = e− 2 ( σ ) T −λWT dP.
170
c 1 2
As an exponential process, eσWt − 2 σ t becomes a martingale on this space. Consequently e−rt St
is a martingale process w.r.t. the probability measure Q. This means
where E Q [ · |Fu ] denotes the conditional expectation in the measure Q, and it is given by
1 µ−r 2
E Q [Xt |Fu ] = E P [Xt e− 2 ( σ ) T −λWT |Fu ].
The measure Q is called the equivalent martingale measure, or the risk-neutral measure. The
expectation taken with respect to this measure is called the expectation in the risk-neutral
world. Customarily we shall use the notations
b −rt St ] =
E[e E Q [e−rt St ]
bu [e−rt St ] =
E E Q [e−rt St |Fu ]
b −rt St ] = E
It worth noting that E[e b0 [e−rt St ], since e−rt St is independent of the initial informa-
tion set F0 .
ct is contained in the following useful result.
The importance of the process W
Proposition 12.1.3 The probability measure that makes the discounted stock price, e−rt St , a
martingale changes the rate of return µ into the risk-free interest rate r, i.e
ct .
dSt = rSt dt + σSt dW
This will clear some unclarities that appear in practical applications when we need to choose
the right probability density.
ct w.r.t. P and Q will be denoted respectively by p , p and pb ,
The densities of Wt and W P Q P
pbQ .
Since Wt and Wct are Brownian motions on the spaces (Ω, F, P ) and (Ω, F, Q), respectively,
they have the following normal probability densities
1 − x2
pP (x) = √ e 2t = p(x);
2πt
1 − x2
pbQ (x) = √ e 2t = p(x).
2πt
The associated distribution functions are
Z Z x
P
FWt (x) = P (Wt ≤ x) = dP (ω) = p(u) du;
{Wt ≤x} −∞
Z Z x
Q
FW ct ≤ x) =
c (x) =
t
Q(W dQ(ω) = p(u) du.
ct ≤x}
{W −∞
Since the underlying Brownian motions are perfectly correlated, one may be tempted to think
that the stock prices S1 and S2 are the same. The following result shows that in general the
stock prices are positively correlated:
Proposition 12.3.1 The correlation coefficient between the stock prices S1 and S2 driven by
the same Brownian motion is
eσ1 σ2 t − 1
Corr(S1 , S2 ) = 2 2 > 0.
(eσ1 t − 1)1/2 (eσ2 t − 1)1/2
In particular, if σ1 = σ2 , then Corr(S1 , S2 ) = 1.
Proof: Since
1 2 1 2
S1 (t) = S1 (0)eµ1 t− 2 σ1 t eσ1 Wt , S2 (t) = S2 (0)eµ2 t− 2 σ2 t , eσ2 Wt ,
2
from Exercise 12.3.4 and formula E[ekWt ] = ek t/2
, yields
Cov(eσ1 Wt , eσ2 Wt )
Corr(S1 , S2 ) = Corr(eσ1 Wt , eσ2 Wt ) = p
V ar(eσ1 Wt )V ar(eσ2 Wt )
E[e(σ1 +σ2 )Wt ] − E[eσ1 Wt ]E[eσ2 Wt ]
= p
V ar(eσ1 Wt )V ar(eσ2 Wt )
1 2 1 2 1 2
e 2 (σ1 +σ2 ) t − e 2 σ1 t e 2 σ2 t
= 2 2 2 2
[eσ1 t (eσ1 t − 1)eσ2 t (eσ2 t − 1)]1/2
eσ 1 σ 2 t − 1
= 2 2 .
(eσ1 t − 1)1/2 (eσ2 t − 1)1/2
If σ1 = σ2 = σ then the previous formula obviously provides
2
eσ t − 1
Corr(S1 , S2 ) = σ2 t = 1,
e −1
i.e. the stocks are perfectly correlated if they have the same volatility.
Corollary 12.3.2 The stock prices S1 and S2 are positively strongly correlated for small values
of t:
Corr(S1 , S2 ) → 1 as t → 0.
This fact has the following financial interpretation. If some stocks are driven by the same
unpredictable news, when one stock increases, then the other tends to increase too, at least
for a small amount of time. In the case when some bad news affect an entire financial market,
the risk becomes systemic, and hence if one stock fails, all the other tend to decrease as well,
leading to a severe strain on the financial market.
173
1.0
0.8
0.6
0.4
0.2
20 40 60 80 100
eσ1 σ2 t −1
Figure 12.1: The correlation function f (t) = 2 2 in the case σ1 = 0.15,
(eσ1 t −1)1/2 (eσ2 t −1)1/2
σ2 = 0.40.
Corollary 12.3.3 The stock prices correlation gets weak as t gets large:
Corr(S1 , S2 ) → 0 as t → ∞.
Proof: It follows from the asymptotic correspondence
eσ1 σ2 t − 1 eσ1 σ2 t (σ −σ )2
− 1 2 2 t
σ12 t 2 ∼ 2 2 = e → 0, t → 0.
(e − 1)1/2 (eσ2 t − 1)1/2 σ1 +σ2
e 2 t
It follows that in long run any two stocks tend to become uncorrelated, see Fig.12.1.
Exercise 12.3.4 If X and Y are random variables and α, β ∈ R, show that
½
Corr(X, Y ), if αβ > 0
Corr(αX, βY ) =
−Corr(X, Y ), if αβ < 0.
Exercise 12.3.5 Find the following
¡ ¢
(a) Cov dS1 (t), dS2 (t) ;
¡ ¢
(b) Corr dS1 (t), dS2 (t) .
Proof: Eliminating the term dWt from equations (12.3.7) − (12.3.8) yields
σ2 σ1
dS1 − dS2 = (µ1 σ2 − µ2 σ1 )dt. (12.4.10)
S1 S2
σ2 (t) σ1 (t)
Consider the portfolio P (t) = θ1 (t)S1 (t) − θ2 (t)S2 (t), with θ1 (t) = and θ2 (t) = .
S1 (t) S2 (t)
Using the properties of self-financing portfolios, we have
µ1 σ2 − µ2 σ1 = rθ1 S1 − rθ2 S2 ,
dS1 = ct
rS1 dt + σ1 S1 dW
dS2 = ct
rS2 dt + σ2 S2 dW
ct is the same in both equations
where the risk-neutral process dW
ct = µ1 − r µ2 − r
dW dt + dWt = dt + dWt .
σ1 σ2
= −re −rt
St dt + e −rt ct )
(rSt dt + σSt dW
ct ).
= e−rt (rSt dt + σSt dW
175
and taking the risk-neutral expectation w.r.t. the information set Fu yields
Z t
b −rt St |Fu ] = E[e
E[e b −ru Su + cs | Fu ]
σe−rs Ss dW
u
Z
£ t −rs ¤
= e−ru Su + E b σe Ss dW cs |Fu
u
Z t
£ ¤
= e−ru Su + E b cs
σe−rs Ss dW
u
= e−ru Su ,
Z t
since cs is independent of Fu . It follows that e−rt St is an Ft -martingale in the
σe−rs Ss dW
u
risk-neutral world. The following fundamental result can be shows using a similar proof as the
one encountered previously:
which assures that e−rt ft is a Zmartingale, since Wcs is a Brownian motion process. Using that
t
∂f s c
e−ru fu is Fu -predictable, and e−rs σS dWs is independent of the information set Fu , we
u ∂S
have Z
£ t −rs ∂fs ¤
b −rt
E[e ft |Fu ] = e −ru b
fu + E e σS cs = e−ru fu .
dW
u ∂S
Exercise 12.5.4 Use risk-neutral valuation to find the price of a derivative that pays at ma-
turity the following payoffs:
(a) fT = T ST ;
RT
(b) fT = 0 Su du;
RT
(c) fT = 0 Su dWu .
Chapter 13
Black-Scholes Analysis
∂u ∂ 2 u
− = 0, (13.1.1)
∂τ ∂x2
called the heat equation. If the initial heat distribution is known and is given by u(0, x) = f (x),
then we have an initial value problem for the heat equation.
Solving this equation involves a convolution between the initial temperature f (x) and the
fundamental solution of the heat equation G(τ, x), which will be defined shortly.
1 x2
G(τ, x) = √ e− 4τ , τ > 0,
4πτ
177
178
1.2
Τ = 0.10
1.0
GHΤ, xL 0.8
Τ = 0.24
0.6
0.4
Τ = 0.38
0.2 Τ = 0.50
x
-2 -1 1 2
Figure 13.1: The function G(τ, x) tends to the Dirac measure δ(x) as τ & 0, and flattens out
as τ → ∞.
for any smooth function with compact support ϕ. Consequently, we also have
Z
ϕ(x)δ(x − y)dx = ϕ(y).
R
One can think of δ(x) as a measure with infinite value at x = 0, zero for the rest of the values
and with the integral equal to 1, see Fig.13.1.
The physical significance of the fundamental solution G(τ, x) is that it describes the heat
evolution in the infinite rod after an initial heat impulse of infinite size applied at x = 0.
is given by the convolution between the fundamental solution and the initial temperature
Z
u(τ, x) = G(τ, y − x)f (y) dy, τ > 0.
R
179
Since the limit and the integral commute2 , using the properties of Dirac measure, we have
Z
u(0, x) = lim u(τ, x) = lim G(τ, z)f (x + z) dz
τ &0 τ &0 R
Z
= δ(z)f (x + z) dz = f (x).
R
The market participant holds aj units of stock Sj and bk units in derivative Fk . The coefficients
are positive for long positions and negative for short positions. For instance, a portfolio given
by P = 2F − 3S means that we buy 2 securities and sell 3 units of stock (a position with 2
securities long and 3 stocks short).
Substituting back in (13.3.4), and collecting the predictable and unpredictable parts, yields
h ∂P ³ ∂P ∂P ∂F ´ ∂P ³ ∂F 1 ∂2F ´
dP = + µS + + + σ2 S 2 2
∂t ∂S ∂F ∂S ∂F ∂t 2 ∂S
1 2 2³ ∂2P ∂ 2 P ³ ∂F ´2 ´i
+ σ S + dt
2 ∂S 2 ∂F 2 ∂S
³ ∂P ∂P ∂F ´
+σS + dWt . (13.3.7)
∂S ∂F ∂S
Looking at the unpredictable component, we have the following result:
dP
Proposition 13.3.1 The portfolio P is risk-less if and only if = 0.
dS
Proof: A portfolio P is risk-less if and only if its unpredictable component is identically zero,
i.e.
∂P ∂P ∂F
+ = 0.
∂S ∂F ∂S
Since the total derivative of P is given by
dP ∂P ∂P ∂F
= + ,
dS ∂S ∂F ∂S
dP
the previous relation becomes = 0.
dS
dP
Definition 13.3.2 The amount ∆P = is called the delta of the portfolio P .
dS
The previous result can be reformulated by saying that a portfolio is risk-less if and only if its
delta vanishes. In practice this can hold only for a short amount of time, so the portfolio need
to be re-balanced periodically. The process of making a portfolio risk-less involves a procedure
called delta hedging, through which the portfolio’s delta becomes zero or very close to this
value.
Assume P is a risk-less portfolio, so
dP ∂P ∂P ∂F
= + = 0. (13.3.8)
dS ∂S ∂F ∂S
Then equation (13.3.7) simplifies to
h ∂P ∂P ³ ∂F 1 ∂2F ´
dP = + + σ2 S 2 2
∂t ∂F ∂t 2 ∂S
1 2 2³ ∂2P ∂ 2 P ³ ∂F ´2 ´i
+ σ S + dt (13.3.9)
2 ∂S 2 ∂F 2 ∂S
Comparing with (13.3.3) yields
∂P ∂P ³ ∂F 1 ∂2F ´ 1 ³ ∂2P ∂ 2 P ³ ∂F ´2 ´
+ + σ2 S 2 2 + σ2 S 2 + = rP. (13.3.10)
∂t ∂F ∂t 2 ∂S 2 ∂S 2 ∂F 2 ∂S
This equation works under the general hypothesis that P = P (t, S, F ) is a risk-free financial
instrument that depends on time t, stock S and derivative F .
182
• no transaction costs.
∂F ∂F 1 ∂2F
+ rS + σ 2 S 2 2 = rF. (13.4.11)
∂t ∂S 2 ∂S
Proof: The equation (13.3.10) works under the general hypothesis that P = P (t, S, F ) is
a risk-free financial instrument that depends on time t, stock S and derivative F . We shall
consider P to be the following particular portfolio
P = F − λS.
This means taking a long position in derivative and a short position in λ units of stock (assuming
λ positive). The partial derivatives in this case are
∂P ∂P ∂P
= 0, = 1, = −λ,
∂t ∂F ∂S
∂2P ∂2P
= 0, = 0.
∂F 2 ∂S 2
∂F
From the risk-less property (13.3.8) we get λ = . Substituting in equation (13.3.10) yields
∂S
∂F 1 ∂2F ∂F
+ σ 2 S 2 2 = rF − rS ,
∂t 2 ∂S ∂S
which is equivalent with the desired equation.
However, the Black-Scholes equation is derived most often in a less rigorous way. This is
based on the assumption that the number λ = ∂F ∂S , which appears in the formula of the risk-less
portfolio P = F −λS, is considered constant for the time interval ∆t. If consider the increments
over the time interval ∆t
∆Wt = Wt+∆t − Wt
∆S = St+∆t − St
∆F = F (t + ∆t, St + ∆S) − F (t, S),
183
∆S = µS∆t + σS∆Wt .
Since both increments ∆F and ∆S are driven by the same uncertainly source, ∆Wt , we can
eliminate it by multiplying the latter equation by ∂F
∂S and subtract it from the former
∂F ³ ∂F 1 ∂2F ´
∆F − (t, S)∆S = (t, S) + σ 2 S 2 2 (t, S) ∆t.
∂S ∂t 2 ∂S
The left side can be regarded as the increment ∆P , of the portfolio
∂F
P =F − S.
∂S
This portfolio is risk-less because its increments are totally deterministic, so it must also satisfy
∆P = rP ∆t. The number ∂F ∂S is assumed constant for small intervals of time ∆t. Even if
this assumption is not rigorous enough, the procedure still leads to the right equation. This is
obtained by equating the coefficients of ∆t in the last two equations
∂F 1 ∂2F ³ ∂F ´
(t, S) + σ 2 S 2 2 (t, S) = r F − S ,
∂t 2 ∂S ∂S
which is equivalent to the Black-Scholes equation.
Example 13.6.1 (i) It is easy to show that F = S is a solution for the Black-Scholes equation.
Hence the stock is a tradable derivative.
(ii) If K is a constant, then F = ert K is a tradable derivative.
(iii) If S is the stock price, then F = eS is not a tradable derivative, since F it does not satisfy
equation (13.4.11).
Proposition 13.6.3 The general form of a traded derivative, which does not depend explicitly
on time, is given by
2
F (S) = C1 S + C2 S −2r/σ , (13.6.12)
with C1 , C2 constants.
Proof: If the derivative depends solely on the stock, F = F (S), then the Black-Scholes equation
becomes the ordinary differential equation
dF 1 d2 F
rS + σ 2 S 2 2 = rF. (13.6.13)
dS 2 dS
This is an Euler-type equation, which can be solved by using the substitution S = ex . The
d d
derivatives and are related by the chain rule
dS dx
d dS d
= .
dx dx dS
dS dex d d
Since = = ex = S, it follows that = S . Using product rule,
dx dx dx dS
d2 d ¡ d ¢ d 2 d
2
= S S = S + S ,
dx2 dS dS dS dS 2
and hence
d2 d2 d
S2 = − .
dS 2 dx2 dx
185
1 2 d2 G(x) 1 dG(x)
σ + (r − σ 2 ) = rG(x).
2 dx2 2 dx
where G(x) = G(ex ) = F (S). The associated indicial equation
1 2 2 1
σ α − (r − σ 2 )α = r
2 2
has solutions α1 = 1, α2 = −r/σ 2 , so the general solution has the form
r
G(x) = C1 ex + C2 e− σ2 x ,
Exercise 13.6.4 Show that the price of a forward contract, which is given by F (t, S) = S − Ke−r(T −t) ,
satisfies the Black-Scholes equation, i.e. a forward contract is a tradable derivative.
2
ln(St /K) + (r − σ2 )(T − t)
d2 = √
σ T −t
√
d1 = d2 + σ T − t.
∂F 1 ∂2F ∂F
+ σ 2 S 2 2 = rF − rS ,
∂t 2 ∂S ∂S
equation (13.3.10) becomes
∂P ∂P ³ ∂F ´ 1 2 2 ³ ∂ 2 P ∂ 2 P ³ ∂F ´2 ´
+ rF − rS + σ S 2
+ = rP.
∂t ∂F ∂S 2 ∂S ∂F 2 ∂S
Using the risk-less condition (13.3.8)
∂P ∂P ∂F
=− , (13.7.14)
∂S ∂F ∂S
186
∂P ∂P ∂P 1 h ∂2P ∂ 2 P ³ ∂F ´2 i
+ rS + +rF + σ2 S 2 + = rP. (13.7.15)
∂t ∂S ∂F 2 ∂S 2 ∂F 2 ∂S
In the following we shall find an equivalent expression for the last term on the left side. Differ-
entiate in (13.7.14) with respect to F yields
∂2P ∂ 2 P ∂F ∂P ∂ 2 F
= − −
∂F ∂S ∂F 2 ∂S ∂F ∂F ∂S
2
∂ P ∂F
= − 2 ,
∂F ∂S
where we used
∂2F ∂ ∂F
= = 1.
∂F ∂S ∂S ∂F
∂F
Multiplying by implies
∂S
∂ 2 P ³ ∂F ´2 ∂ 2 P ∂F
2
=− .
∂F ∂S ∂F ∂S ∂S
Substituting in the aforementioned equation yields
∂P ∂P ∂P 1 h ∂2P ∂ 2 P ∂F i
+ rS + +rF + σ2 S 2 − = rP. (13.7.16)
∂t ∂S ∂F 2 ∂S 2 ∂F ∂S ∂S
Application 13.7.1 If a risk-less investment P has the variable S and F separable, i.e. it is
the sum P (S, F ) = f (F ) + g(S), with f and g smooth functions, then
2
P (S, F ) = F + c1 S + c2 S −2r/σ .
Since P has separable variables, the mixed derivative term vanishes, and the equation (13.7.16)
becomes
σ 2 2 00
Sg 0 (S) + S g (S) − g(S) = f (F ) − F f 0 (F ).
2r
There is a separation constant C such that
f (F ) − F f 0 (F ) = C
2
σ 2 00
Sg 0 (S) + S g (S) − g(S) = C.
2r
Dividing the first equation by F 2 yields the exact equation
³1 ´0 C
f (F ) = − 2 ,
F F
187
2
with the solution f (F ) = c0 F + C. To solve the second equation, let κ = σ2r . Then the
substitution S = ex leads to the ordinary differential equation with constant coefficients
κλ2 + (1 − κ)λ − 1 = 0
has the solutions λ1 = 1, λ2 = − κ1 . The general solution is the sum between the particular
solution hp (x) = −C and the solution of the associated homogeneous equation, which is h0 (x) =
1
c1 ex + c2 e− κ x . Then
1
h(x) = c1 ex + c2 e− κ x − C.
Going back to the variable S, we get the general form of g(S)
2
g(S) = c1 S + c2 S −2r/σ − C,
with c1 , c2 constants. Since the constant C cancels by addition, we have the following formula
for the risk-less investment with separable variables F and S
2
P (S, F ) = f (F ) + g(S) = c0 F + c1 S + c2 S −2r/σ .
Dividing by c0 , we may assume c0 = 1. We shall find the derivative F (t, S) which enters the
previous formula. Substituting in (13.7.14) yields
∂F 2r 2
− = c1 − 2 c2 S −1−2r/σ ,
∂S σ
which after partial integration in S gives
2
F (t, S) = −c1 S − c2 S −2r/σ + φ(t),
where the integration constant φ(t) is a function of t. The sum of the first two terms is the
derivative given by formula (13.6.12). The remaining function φ(t) has also to satisfy the Black-
Scholes equation, and hence it is of the form φ(t) = c3 ert , with c3 constant. Then the derivative
F is given by
2
F (t, S) = −c1 S − c2 S −2r/σ + c3 ert .
It worth noting that substituting in the formula of P yields P = c3 ert , which agrees with the
formula of a risk-less investment.
Exercise 13.7.2 Find the function g(S) such that the product P = F g(S) is a risk-less invest-
ment, with F = F (t, S) derivative. Find the expression of the derivative F in terms of S and
t.
dg(S) σ 2 2 d2 g(S)
S + S = 0.
dS 2r dS 2
Substituting S = ex , and h(x) = g(ex ) = g(S) yields
³ 2r ´
h00 (x) + − 1 h0 (x) = 0.
σ2
188
2r 1− 2r
σ2
g(S) = h(ln S) = C1 + C2 e(1− σ2 ) ln S = C1 + C2 S .
∂F ∂F 1 ∂2F
+ rS + σ2 S 2 2 = rF
∂t ∂S 2 ∂S
F (T, ST ) = f T (ST ).
This means the solution is known at the final time T and we need to find its expression at any
time t prior to T , i.e.
ft = F (t, St ), 0 ≤ t < T.
First we shall transform the equation into an equation with constant coefficients. Substi-
tuting S = ex , and using the identities
∂ ∂ ∂2 ∂2 ∂
S = , S2 2
= −
∂S ∂x ∂S ∂x2 ∂x
the equation becomes
∂V 1 ∂2V 1 ∂V
+ σ 2 2 + (r − σ 2 ) = rV,
∂t 2 ∂x 2 ∂x
where V (t, x) = F (t, ex ). Using the time scaling τ = 12 σ 2 (T − t), chain rule provides
∂ ∂τ ∂ 1 ∂
= = − σ2 .
∂t ∂t ∂τ 2 ∂τ
2r
Denote k = . Substituting in the aforementioned equation yields
σ2
∂W ∂2W ∂W
= + (k − 1) − kW, (13.8.17)
∂τ ∂x2 ∂x
189
where W (τ, x) = V (t, x). Next we shall get rid of the last two terms on the right side of the
equation by using a crafted substitution.
Consider W (τ, x) = eϕ u(τ, x), where ϕ = αx + βτ , with α, β constants that will be deter-
mined such that the equation satisfied by u(τ, x) has on the right side only the second derivative
in x. Since
∂W ³ ∂u ´
= eϕ αu +
∂x ∂x
∂2W ³ ∂u ∂u ´
= eϕ α2 u + 2α +
∂x2 ∂x ∂x2
∂W ³ ∂u ´
= eϕ βu + ,
∂τ ∂τ
substituting in (13.8.17), dividing by eϕ and collecting the derivatives yields
∂u ∂2u ¡ ¢ ∂u ¡ 2 ¢
= 2
+ 2α + k − 1 + α + α(k − 1) − k − β u = 0
∂τ ∂x ∂x
∂u
The constants α and β are chosen such that the coefficients of ∂x and u vanish
2α + k − 1 = 0
2
α + α(k − 1) − k − β = 0.
Solving yields
k−1
α = −
2
(k + 1)2
β = α2 + α(k − 1) − k = − .
4
The function u(τ, x) satisfies the heat equation
∂u ∂2u
=
∂τ ∂x2
with the initial condition expressible in terms of fT
From the general theory of heat equation, the solution can be expressed as the convolution
between the fundamental solution and the initial condition
Z ∞
1 (y−x)2
u(τ, x) = √ e− 4τ u(0, y) dy
−∞ 4πτ
The previous substitutions yield the following relation between F and u
Z ∞
1 (y−x)2
F (t, ex ) = eϕ(τ,x) u(τ, x) = eϕ(τ,x) √ e− 4τ u(0, y) dy
−∞ 4πτ
Z ∞
1 (y−x) 2
= eϕ(τ,x) √ e− 4τ e−αy F (T, ey ) dy.
−∞ 4πτ
√
With the substitution y = x = s 2τ this becomes
Z ∞ √ √
x ϕ(τ,x) 1 s2
F (t, e ) = e √ e− 2 −α(x+s 2τ ) F (T, ex+s 2τ ) dy.
−∞ 2π
s2 √ 1³ k − 1 √ ´2 (k − 1)2 τ k−1
− − α(x + s 2τ ) = s− 2τ + + x,
2 2 2 4 2
Using
(k+1)2 (k−1)2
e− 4 τ
e 4 τ
= e−kτ = e−r(T −t) ,
1
(k − 1)τ = (r − σ 2 )(T − t),
2
√
after the substitution z = x + s 2τ we get
Z ∞
x −r(T −t) 1 1 (z−x−(k−1)τ )
2
1
F (t, e ) = e √ e− 2 2τ F (T, ez ) √ dz
2π −∞ 2τ
Z ∞ [z−x−(r− 1 σ 2 )(T −t)]2
1 − 2
= e−r(T −t) p e 2σ 2 (T −t) F (T, ez ) dz.
2
2πσ (T − t) −∞
Example 13.9.1 (a) The price of an European call option is the solution F (t, S) of the Black-
Scholes equation satisfying
fT (ST ) = max(ST − K, 0).
(b) The price of an European put option is the solution F (t, S) of the Black-Scholes equation
with the final condition
fT (ST ) = max(K − ST , 0).
(c) The value of a forward contract is the solution the Black-Scholes equation with the final
condition
fT (ST ) = ST − K.
It is worth noting that the superposition principle discussed in section 11 can be explained
now by the fact that the solution space of the Black-Scholes equation is a linear space. This
means that a linear combination of solutions is also a solution.
Another interesting feature of the Black-Scholes equation is its independence of the stock
drift rate µ. Then its solutions must have the same property. This explains why in the risk-
neutral valuation the value of µ does not appear explicitly in the solution.
Asian options satisfy similar Black-Scholes equations, with small differences, as we shall see
in the next section.
When S → 0, the option does not get exercised, so the initial boundary condition is
F (t, 0) = 0.
192
20
15
10 S - K
10 20 30 40 50 60
Figure 13.2: The graph of the option price before maturity in the case K = 40, σ = 30%,
r = 8%, and T − t = 1.
³ ∂P∂P ∂F ´
σS + dWt
∂S ∂F ∂S
³ ∂P ∂P ∂F ´ 1 2 2 ³ ∂P ∂ 2 F ∂2P ∂2P ∂2F ´
+ρS + + ρ S + + dMt .
∂S ∂F ∂S 2 ∂F ∂S 2 ∂S 2 ∂F 2 ∂S 2
The risk-less condition for the portfolio P is obtained when the coefficients of dWt and dMt
vanish
∂P ∂P ∂F
+ = 0 (13.11.23)
∂S ∂F ∂S
2 2 2 2
∂P ∂ F ∂ P ∂ P∂ F
2
+ 2
+ = 0. (13.11.24)
∂F ∂S ∂S ∂F 2 ∂S 2
These relations can be further simplified. If differentiate in (13.11.23) with respect to S
∂2P ∂P ∂ 2 F ∂ 2 P ∂F
+ = − .
∂S 2 ∂F ∂S 2 ∂S∂F ∂S
substituting in (13.11.24) yields
∂2P ∂2F ∂ 2 P ∂F
2 2
= . (13.11.25)
∂F ∂S ∂S∂F ∂S
Differentiating in (13.11.23) with respect to F we get
∂2P ∂ 2 P ∂F ∂P ∂ 2 F
= − 2
−
∂F ∂S ∂F ∂S ∂F ∂F ∂S
∂ 2 P ∂F
= − 2 , (13.11.26)
∂F ∂S
194
since
∂2F ∂ ³ ∂F ´
= .
∂F ∂S ∂S ∂F
∂F
Multiplying (13.11.26) by yields
∂S
∂ 2 P ∂F ∂ 2 P ³ ∂F ´2
=− 2 ,
∂F ∂S ∂S ∂F ∂S
and substituting in the right side of (13.11.25) leads to the equation
∂ 2 P h ∂ 2 F ³ ∂F ´2 i
+ = 0.
∂F 2 ∂S 2 ∂S
We arrived at the following result:
Proposition 13.11.2 Let F = F (t, S) be a derivative with the underlying asset S. The in-
vestment P = P (t, S, F ) is risk-less if and only if
∂P ∂P ∂F
+ = 0
∂S ∂F ∂S
∂ 2 P h ∂ 2 F ³ ∂F ´2 i
+ = 0.
∂F 2 ∂S 2 ∂S
There are two risk-less conditions because there are two unpredictable components in the incre-
ments of dP , one due to regular changes and the other dues to rare events. The first condition
dP
is equivalent with the vanishing total derivative, = 0, and corresponds to offsetting the
dS
regular risk.
∂2P ∂ 2 F ³ ∂F ´2
The second condition vanishes either if = 0 or if + = 0. In the first case P
∂F 2 ∂S 2 ∂S
∂F
is at most linear in F . For instance, if P = F −f (S), from the first condition yields f 0 (S) = .
∂S
∂F
In the second case, denote U (t, S) = ∂S . Then need to solve the partial differential equation
∂U
+ U 2 = 0.
∂t
Future research directions:
1. Solve the above equation
2. Find the predictable part of dP
3. Get an analog of the Black-Scholes in this case
4. Evaluate a call option in this case
5. Is the risk-neutral valuation still working and why?
Chapter 14
In this chapter we shall develop the Black-Scholes equation in the case of Asian derivatives and
we shall discuss the particular cases of options and forward contracts on weighted averages.
In the case of the later contracts we obtain closed form solutions, while for the former ones
we apply the reduction variable method to decrease the number of variables and discuss the
solution.
1
Example 14.0.3 (a) The uniform weight is obtained for ρ(t) = . In this case
T
Z T
1
Save = St dt
T 0
is the continuous arithmetic average of the stock on the time interval [0, T ].
195
196
2t
(b) The linear weight is obtained if ρ(t) = . In this case the weight is the time
T2
Z T
2
Save = 2 tSt dt.
T 0
kekt
(c) The exponential weight is obtained for ρ(t) = . If k > 0, the weight is increasing, so
ekT − 1
recent data are weighted more than old data; if k < 0, the weight is decreasing. The exponential
weighted average is given by
Z T
k
Save = kT ekt St dt.
e −1 0
(n)
Find the limit lim Save in the cases 0 < T < 1, T = 1, and T > 1.
n→∞
f (t) RT
In all previous examples ρ(t) = ρ(t, T ) = , with 0 f (t) dt = g(T ), so g 0 (T ) = f (T )
g(T )
and g(0) = 0. The average becomes
Z T
1 IT
Save (T ) = f (u)Su du = ,
g(T ) 0 g(T )
Rt
with It = 0
f (u)Su du satisfying dIt = f (t)St dt. From the product rule we get
à !
dIt g(t) − It dg(t) f (t) g 0 (t) It
dSave (t) = = St − dt
g(t)2 g(t) g(t) g(t)
f (t) ³ g 0 (t) ´
= St − Save (t) dt
g(t) f (t)
f (t) ³ ´
= St − Save (t) dt,
g(t)
It f (t)St f (t)
Save (0) = lim Save (t) = lim = lim 0 = S0 lim 0 = S0 ,
t&0 t&0 g(t) t&0 g (t) t&0 g (t)
Proposition 14.0.5 The weighted average Save (t) satisfies the stochastic differential equation
f (t)
dXt = (St − Xt )dt
g(t)
X0 = S0 .
197
and Proposition 14.0.5, an application of Ito’s formula together with the stochastic formulas
yields
∂F ∂F 1 ∂2F ∂F
dF = dt + dSt + (dSt )2 + dSave
∂t ∂St 2 ∂St2 ∂Save
³ ∂F ∂F 1 ∂2F f (t) ∂F ´
= + µSt + σ 2 St2 2 + (St − Save ) dt
∂t ∂St 2 ∂St g(t) ∂Save
∂F
+σSt dWt .
∂St
∂F
Let ∆F = ∂St . Consider the following portfolio at time t
P (t) = F − ∆F St ,
obtained by buying one derivative F and selling ∆F units of stock. The change in the portfolio
value during the time dt does not depend on Wt
dP = dF − ∆F dSt
³ ∂F 1 ∂2F f (t) ∂F ´
= + σ 2 St2 2 + (St − Save ) dt (14.1.1)
∂t 2 ∂St g(t) ∂Save
so the portfolio P is risk-less. Since no arbitrage opportunities are allowed, investing a value P
at time t in a bank at the risk-free rate r for the time interval dt yields
³ ∂F ´
dP = rP dt = rF − rSt dt. (14.1.2)
∂St
198
Equating (14.1.1) and (14.1.2) yields the following form of the Black-Scholes equation for Asian
derivatives on weighted averages
∂F ∂F 1 ∂2F f (t) ∂F
+ rSt + σ 2 St2 2 + (St − Save ) = rF.
∂t ∂St 2 ∂St g(t) ∂Save
∂V ∂V 1 ∂2V ∂V
+ rSt + σ 2 St2 2 + f (t)St = rV. (14.2.3)
∂t ∂St 2 ∂St ∂It
The payoff at maturity of an average strike call option can be written in the following form
where
It 1
Rt = , L(t, R) = max{1 − R, 0}.
St g(t)
Since at maturity the variable ST is separated from T and RT , we shall look for a solution of
equation (14.2.3) of the same type for any t ≤ T , i.e. V (t, S, I) = SG(t, R) . Since
∂V ∂G ∂V ∂G 1 ∂G
= S , =S = ;
∂t ∂t ∂I ∂R S ∂R
∂V ∂G ∂R ∂G
= G+S =G−R ;
∂S ∂R ∂S ∂R
∂2V ∂ ∂G ∂G ∂R ∂R ∂G ∂ 2 G ∂R
= (G − R )= − −R 2 ;
∂S 2 ∂S ∂R ∂R ∂S ∂S ∂R ∂R ∂S
∂2G I
= R 2 ;
∂R S
∂2V ∂2G
S2 = RI ·
∂S 2 ∂R2
RI
Substituting in (14.2.3) and using that = R2 , after cancelations yields
S
∂G 1 2 2 ∂ 2 G ∂G
+ σ R + (f (t) − rR) = 0. (14.2.4)
∂t 2 ∂R2 ∂R
199
This is a partial differential equation in only two variables, t and R. It can be solved explicitly
sometimes, depending of the form of the final condition G(T, RT ) and expression of the function
f (t).
In the case of a weighted average strike call option the final condition is
RT
G(T, RT ) = max{1 − , 0}. (14.2.5)
g(T )
Example 14.2.1 In the case of the arithmetic average the function G(t, R) satisfies the partial
differential equation
∂G 1 2 2 ∂ 2 G ∂G
+ σ R 2
+ (1 − rR) = 0
∂t 2 ∂R ∂R
RT
with the final condition G(T, RT ) = max{1 − , 0}.
T
Example 14.2.2 In the case of the exponential average the function G(t, R) satisfies the equa-
tion
∂G 1 2 2 ∂ 2 G ∂G
+ σ R 2
+ (kekt − rR) =0 (14.2.6)
∂t 2 ∂R ∂R
RT
with the final condition G(T, RT ) = max{1 − kT , 0}.
e −1
Neither of the previous two final condition problems can be solved explicitly.
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
20 40 60 80 100 20 40 60 80 100
(a) (b)
Figure 14.1: The profile of the solution H(t, R): (a) at expiration; (b) when there is T − t time
left before expiration.
It can be shown in the theory of partial differentia equations that the equation (14.2.4) together
with the final condition (14.2.5), see Fig.14.1(a), and boundary conditions (14.3.7) and (14.3.8)
has a unique solution G(t, R), see Fig.14.1(b).
There is no close form solution for the weighted average strike call option. Even in the most
simple case, when the average is arithmetic, the solution is just approximative, see section
11.10. In real life the price is worked out using the Monte-Carlo simulation. This is based
on averaging a large number, n, of simulations of the process Rt in the risk-neutral world, i.e.
RT,j
assuming µ = r. For each realization, the associated payoff GT,j = max{1− } is computed,
g(T )
with j ≤ n. Here RT,j represents the value of R at time T in the jth realization. The average
n
1X
GT,j
n j=1
is a good approximation of the payoff expectation E[GT ]. Discounting under the risk-free rate
we get the price at time t
³1 Xn ´
G(t, R) = e−r(T −t) GT,j .
n j=1
It worth noting that the term on the right is an approximation of the risk neutral conditional
b T |Ft ].
expectation E[G
When simulating the process Rt , it is convenient to know its stochastic differential equation.
Using
³1´ 1³ 2 ´
dIt = f (t)St dt, d = (σ − µ)dt − σdWt dt,
St St
the product rule yields
201
³I ´ ³ 1´
t
dRt = d = d It
St St
1 ³1´ ³1´
= dIt + It d + dIt d
St S St
³ t ´
2
= f (t)dt + Rt (σ − µ)dt − σdWt .
I0
The initial condition is R0 = S0 = 0, since I0 = 0.
Can we solve explicitly this equation? Can we find the mean and variance of Rt ?
We shall start by finding the mean E[Rt ]. The equation can be written as
Integrating yields
Z t Z t
2 2 2
e−(σ −µ)t
Rt = e−(σ −µ)u
f (u) du − σe−(σ −µ)u
Ru dWu
0 0
The first integral is deterministic while the second is an Ito integral. Using that the expectations
of Ito integrals vanish, we get
Z t
2 2
E[e−(σ −µ)t Rt ] = e−(σ −µ)u f (u) du
0
and hence
Z t
2 2
E[Rt ] = e(σ −µ)t
e−(σ −µ)u
f (u) du.
0
Substituting Yt = ρt Rt yields
¡ ¢
dYt = ρt f (t) + (σ 2 − µ)Yt dt,
202
Going back in the variable Rt = Yt /ρt , we obtain the following closed form expression
Z t
1 2
Rt = e(µ− 2 σ )(u−t)+σ(Wu −Wt )
f (u) du. (14.3.10)
0
V (t, St , It ) = St G(t, Rt )
Z T
1 £ −r(T −t) ¤
= St − It e + St f (u)e−r(T −u) du .
g(T ) t
f (u)
Using that ρ(u) = , going back to the initial variable Save (t) = It /g(t) yields
g(T )
Proposition 14.4.1 The value at time t of an Asian forward contract on a weighted average
with the weight function ρ(t), i.e. an Asian derivative with the payoff FT = ST − Save (T ), is
given by
It worth noting that the previous price can be written as a linear combination of St and
Save (t)
F (t, St , Save (t)) = α(t)St + β(t)Save (t),
where
Z T
α(t) = 1− ρ(u)e−r(T −u) du
t
Rt
g(t) −r(T −t) f (u) du −r(T −t)
β(t) = − e = − R 0T e .
g(T ) f (u) du
0
In the first formula ρ(u)e−r(T −u) is the discounted weight at time u, and α(t) is 1 minus the
total discounted weight between t and T . One can easily check that α(T ) = 1 and β(T ) = −1.
Exercise 14.4.2
Rt Find the value at time t of an Asian forward contract on an arithmetic av-
erage At = 0 Su du.
Exercise 14.4.3 (a) Find the value at time t of an Asian forward contract on an exponential
weighted average with the weight given by Example 14.0.3 (c).
(b) What happens if k = −r? Why?
Exercise 14.4.4 Find the value at time t of an Asian power contract with the payoff FT =
¡RT ¢n
0
Su du .
Hints and Solutions
205