6 Introduction to Stochastic Calculus

advertisement
Chapter 6 Fundamentals of
Stochastic Calculus
Stochastic Calculus: Introduction
Although stochastic and ordinary calculus share many common
properties, there are fundamental differences. The probabilistic nature of
stochastic processes distinguishes them from the deterministic functions
associated with ordinary calculus. Since stochastic differential equations
so frequently involve Brownian motion, second order terms in the Taylor
series expansion of functions become important, in contrast to ordinary
calculus where they can be ignored.
Differentials Of Stochastic Processes
In many respects, differentials of stochastic processes mirror differentials
of real-valued functions from ordinary calculus, sharing many of their
properties. However, there are also important differences. In ordinary
calculus, if Xt is a real-valued function of the real variable t, then the
derivative 𝑋𝑑′ exists for a large class of functions. If Xt is a stochastic
process, then it is usually not possible to well-define the derivative of Xt
with respect to t, at least for the class of stochastic processes that are
relevant in finance. This is because normally Xt involves Brownian
motion, and Brownian motion is not differentiable.
Differentials Of Stochastic Processes (Continued)
The differential of a stochastic process is the change of a stochastic
process Xt resulting from a small change in t. Define the differential dXt
of a stochastic process Xt to be a quantity that satisfies the property
𝑋𝑑+𝑑𝑑 − 𝑋𝑑 − 𝑑𝑋𝑑
lim
= 0.
𝑑𝑑→0
𝑑𝑑
The differential dXt is used to approximate Xt+dt –Xt, and any terms that
approach zero after dividing by dt as dt→0 can be ignored.
An Example Of A Real-Valued Differential From
Ordinary Calculus
Consider the (ordinary) real-valued function X(t) = t3, with t and y being
real variables. Then:
𝑋 𝑑 + 𝑑𝑑 − 𝑋 𝑑 = (𝑑 + 𝑑𝑑)3 − 𝑑 3 = 3𝑑 2 𝑑𝑑 + 3𝑑 𝑑𝑑
2
+ 𝑑𝑑
3
Notice that [3t(dt)2 + (dt)3]/dt = 3tdt + (dt)2 → 0 as dt → 0. This means
that for small values of dt, 3t(dt)2 + (dt)3 is much smaller than 3t2dt and
we can approximate X(t+dt)-X(t) by 3t2dt. So, the differential of X(t) = t3
is dX = 3t2dt.
An Example Of A Stochastic Differential
Now, consider the following stochastic process Xt = tZt, with Zt being
standard Brownian motion:
𝑋𝑑+𝑑𝑑 − 𝑋𝑑 = 𝑑 + 𝑑𝑑 𝑍𝑑+𝑑𝑑 − 𝑑𝑍𝑑
= 𝑑 𝑍𝑑+𝑑𝑑 − 𝑍𝑑 + 𝑍𝑑+𝑑𝑑 𝑑𝑑
= 𝑑 𝑍𝑑+𝑑𝑑 − 𝑍𝑑 + 𝑍𝑑+𝑑𝑑 𝑑𝑑 − 𝑍𝑑 𝑑𝑑 + 𝑍𝑑 𝑑𝑑
= 𝑑 𝑍𝑑+𝑑𝑑 − 𝑍𝑑 + 𝑍𝑑+𝑑𝑑 − 𝑍𝑑 𝑑𝑑 + 𝑍𝑑 𝑑𝑑
= 𝑑𝑑𝑍𝑑 + 𝑑𝑍𝑑 𝑑𝑑 + 𝑍𝑑 𝑑𝑑
An Example Of A Stochastic Differential (Continued)
Consider choosing dXt = tdZt + Ztdt as a differential of Xt. With this
choice, note that
𝑋𝑑+𝑑𝑑 − 𝑋𝑑 − 𝑑𝑋𝑑 𝑑𝑑𝑍𝑑 + 𝑑𝑍𝑑 𝑑𝑑 + 𝑍𝑑 𝑑𝑑 − 𝑑𝑑𝑍𝑑 − 𝑍𝑑 𝑑𝑑
=
𝑑𝑑
𝑑𝑑
𝑑𝑍𝑑 𝑑𝑑
=
= 𝑑𝑍𝑑
𝑑𝑑
Since dZt = Zt+dt – Zt, dZt is distributed normally with mean 0 and variance dt. Thus, with
probability approaching 1, the values of dZt must be on the order of the standard deviation 𝑑𝑑.
This means that the quotient above approaches 0 with probability 1 as dt → 0. This confirms that
dXt = tdZt + Ztdt is a differential of Xt.
Properties Of The Differential
1.
Linearity: If Xt and Yt are stochastic processes, and a and b are
constants, then d(aXt +bYt) = adXt +bdYt.
2. General Product Rule: If Xt and Yt are stochastic processes, then
d(XtYt) = XtdYt + YtdXt +dXtdYt.
3. Special Product Rule: Suppose that dXt = μtdt + σtdZt and dYt =ρtdt
where σt, μt, and ρt are stochastic processes, then d(XtYt) = XtdYt +
YtdXt.
Proofs Of Properties Of The Differential
As to property 1 above, note that
𝑑 π‘Žπ‘‹π‘‘ + π‘π‘Œπ‘‘ = π‘Žπ‘‹π‘‘+𝑑𝑑 + π‘π‘Œπ‘‘+𝑑𝑑 − π‘Žπ‘‹π‘‘ + π‘π‘Œπ‘‘
= π‘Ž 𝑋𝑑+𝑑𝑑 − 𝑋𝑑 + 𝑏 π‘Œπ‘‘+𝑑𝑑 − π‘Œπ‘‘ = π‘Žπ‘‘π‘‹π‘‘ + π‘π‘‘π‘Œπ‘‘
As to property 2, since Xt+dt – Xt = dXt and Yt+dt - Yt = dYt, then Xt+dt = Xt
+ dXt and Yt+dt = Yt+ dYt. Thus, the differential of XtYt is:
𝑑 𝑋𝑑 π‘Œπ‘‘ = 𝑋𝑑+𝑑𝑑 π‘Œπ‘‘+𝑑𝑑 − 𝑋𝑑 π‘Œπ‘‘ = 𝑋𝑑 + 𝑑𝑋𝑑 π‘Œπ‘‘ + π‘‘π‘Œπ‘‘ − 𝑋𝑑 π‘Œπ‘‘
= 𝑋𝑑 π‘‘π‘Œπ‘‘ + π‘Œπ‘‘ 𝑑𝑋𝑑 + 𝑑𝑋𝑑 π‘‘π‘Œπ‘‘
Proofs Of Properties Of The Differential (Continued)
To derive property 3, we make use of property 2. The term dXtdYt can be
ignored in the differential as we now demonstrate:
𝑑𝑋𝑑 π‘‘π‘Œπ‘‘ = πœ‡π‘‘ 𝑑𝑑 + πœŽπ‘‘ 𝑑𝑍𝑑 πœŒπ‘‘ 𝑑𝑑 = πœ‡π‘‘ πœŒπ‘‘ (𝑑𝑑)2 +πœŽπ‘‘ πœŒπ‘‘ 𝑑𝑍𝑑 𝑑𝑑.
We saw earlier that the values of dZt are on the order of 𝑑𝑑, which implies that the
values for σtρtdZtdt/dt = σtρtdZt are on the order of 𝑑𝑑. This implies that the term
σtρtdZtdt/dt approaches 0 as dt approaches 0. Obviously, μtρt(dt)2/dt = μtρtdt
approaches 0 as dt approaches 0. This proves that the term dXtdYt can be ignored in
the differential, which implies that d(XtYt) = XtdYt + YtdXt.
Stochastic Integration
ο‚— If f(x) is a real-valued continuous function defined on the interval
a ≤ x ≤ b, the integral of f(x) is
𝑏
𝑛−1
𝑓 π‘₯ 𝑑π‘₯ = lim
βˆ†π‘₯→0
π‘Ž
𝑓 π‘₯𝑖 βˆ†π‘₯
𝑖=0
ο‚— Integration of a stochastic process Xt with respect to the real variable t
is defined as:
𝑏
𝑛−1
𝑋𝑑 𝑑𝑑 = lim
βˆ†π‘‘→0
π‘Ž
𝑋𝑑𝑖 βˆ†π‘‘
𝑖=0
Where Δt = (b – a)/n and ti = a + iΔt for i = 0,1,…,n
Stochastic Integration (Continued)
To define the integral of a stochastic process Xt with respect to a stochastic
process Yt, divide the interval a ≤ t ≤ b in the same way as above. Then
𝑏
𝑛−1
𝑋𝑑 π‘‘π‘Œπ‘‘ = lim
βˆ†π‘‘→0
π‘Ž
𝑋𝑑𝑖 π‘Œπ‘‘π‘–+1 − π‘Œπ‘‘π‘–
𝑖=0
Once again, whenever this integral exists, the result is a random variable, since it is a limiting sum
of random variables. Notice that our first definition of an integral is a special case of the second by
choosing Yt = t so that dYt = dt. Although Yt = t is a real-valued function, it is also a special case of a
stochastic process. This follows from the observation that for each fixed value of t, Xt = t with
probability 1.
Stochastic Integration (Continued)
Consider the following particular case of stochastic integration that will
arise frequently in this book. Suppose that σt and μt are stochastic
processes (note that σt and μt can be chosen so that each takes on single
constant values) and Zt is standard Brownian motion. Then
𝑏
𝑛−1
(πœŽπ‘‘ 𝑑𝑍𝑑 + πœ‡π‘‘ 𝑑𝑑) = lim
βˆ†π‘‘→0
π‘Ž
πœŽπ‘‘π‘– 𝑍𝑑𝑖+1 − 𝑍𝑑𝑖 + πœ‡π‘‘π‘– βˆ†π‘‘ ,
𝑖=0
where Δt = (b – a)/n, t0 = a, tn = b, and ti+1 – ti = Δt for i = 1, 2, …,n.
Contrasting Integration of Real-valued Functions with
Integration of Stochastic Processes
Integration of Real-valued Functions(1)
2
Integration of Stochastic Processes(2)
2
5𝑑𝑋𝑑 = 5𝑋𝑑 |20 = 5𝑑 2 |20
0
= 5(2)2 −5 0
2
5𝑑𝑍𝑑 = 5𝑍𝑑 |20
0
= 20
Where Xt = t2
= 5𝑍2 − 5𝑍0 = 5𝑍2
Where Zt is standard Brownian motion
The solution in the first illustration is a number. The solution in the second illustration is a random variable
since Z2 is a normally distributed random variable, with mean 0 and variance 2. Thus, 5Z2 is a normally
distributed random variable with mean 0 and variance (5)2(2) = 50.
Elementary Properties of Stochastic Integrals
ο‚— Integral of a Stochastic Differential
ο‚— Linearity
Integral of a Stochastic Differential
If Xt is a stochastic process, the integral of a stochastic differential dXt is :
𝑏
𝑑𝑋𝑑 = 𝑋𝑏 − π‘‹π‘Ž
π‘Ž
Proof:
𝑏
𝑑𝑋𝑑
π‘Ž
= lim
βˆ†π‘‘→0
𝑛−1
𝑖=0 (𝑋𝑑𝑖+1
− 𝑋𝑑𝑖 )
= lim (𝑋𝑑1 −𝑋𝑑0 ) + (𝑋𝑑2 −𝑋𝑑1 ) + β‹― + (𝑋𝑑𝑛−1 −𝑋𝑑𝑛−2 ) + (𝑋𝑑𝑛 −𝑋𝑑𝑛−1 )
βˆ†π‘‘→0
= lim 𝑋𝑑𝑛 + −𝑋𝑑𝑛−1 + 𝑋𝑑𝑛−1 + −𝑋𝑑𝑛−2 + 𝑋𝑑𝑛−2 + β‹― + −𝑋𝑑1 + 𝑋𝑑1 − 𝑋𝑑0
βˆ†π‘‘→0
= lim (𝑋𝑑𝑛 −𝑋𝑑0 ) = 𝑋𝑏 − π‘‹π‘Ž
βˆ†π‘‘→0
Linearity
If Xt, Yt, Ut, and Vt are stochastic processes, and C and D are real
numbers, then
𝑏
𝑏
(πΆπ‘ˆπ‘‘ 𝑑𝑋𝑑 + 𝐷𝑉𝑑 π‘‘π‘Œπ‘‘ ) = 𝐢
π‘Ž
Proof:
π‘ˆπ‘‘ 𝑑𝑋𝑑 + 𝐷
π‘Ž
𝑏
𝑉𝑑 π‘‘π‘Œπ‘‘
π‘Ž
𝑛−1
(πΆπ‘ˆπ‘‘ 𝑑𝑋𝑑 + 𝐷𝑉𝑑 π‘‘π‘Œπ‘‘ ) = lim
βˆ†π‘‘→0
π‘Ž
𝑛−1
= lim 𝐢
βˆ†π‘‘→0
𝑏
πΆπ‘ˆπ‘‘π‘– (𝑋𝑑𝑖+1 − 𝑋𝑑𝑖 ) + 𝐷𝑉𝑑𝑖 (π‘Œπ‘‘π‘–+1 − π‘Œπ‘‘π‘– )
𝑖=0
𝑛−1
π‘ˆπ‘‘π‘– (𝑋𝑑𝑖+1 − 𝑋𝑑𝑖 ) + lim 𝐷
𝑖=0
βˆ†π‘‘→0
𝑏
𝑉𝑑𝑖 (π‘Œπ‘‘π‘–+1 − π‘Œπ‘‘π‘– ) = 𝐢
𝑖=0
𝑏
π‘ˆπ‘‘ 𝑑𝑋𝑑 + 𝐷
π‘Ž
𝑉𝑑 π‘‘π‘Œπ‘‘
π‘Ž
Illustration
Suppose that the price of a security follows the following arithmetic Brownian
motion process with drift: dXt = 06dZt +.02dt. Further suppose that the price
of the security at time zero equals 20: X0 = 20. Now we seek to find Xt, the price
of the security at time t: First observe that
𝑑
𝑑
𝑑𝑋𝑠 =
0
.06𝑑𝑍𝑠 + .02𝑑𝑠 .
0
By the first property, the left side equals:
𝑑
𝑑𝑋𝑠 = 𝑋𝑠 |𝑑0 = 𝑋𝑑 −20,
0
Illustration (Continued)
By linearity and the first property, the right side equals:
𝑑
.06𝑑𝑍𝑠 + .02𝑑𝑠 = .06𝑍𝑠 + .02𝑠 |𝑑0
0
= .06𝑍𝑑 + .02𝑑 − .06𝑍0 = .06𝑍𝑑 + .02𝑑
Setting these results equal and solving for Xt, we find that Xt = 20+.02t
+.06Zt. The security price at time t equals its price at time zero plus .02
multiplied by the elapsed time plus .06 times the Brownian motion. The
price is a normally distributed random variable with mean 20+.02t and
variance (.06)2t
More on Defining Stochastic Integrals
𝑇
πΆπ‘œπ‘›π‘ π‘–π‘‘π‘’π‘Ÿ π‘‘β„Žπ‘’ π‘“π‘œπ‘™π‘™π‘œπ‘€π‘–π‘›π‘” π‘ π‘‘π‘œπ‘β„Žπ‘Žπ‘ π‘‘π‘–π‘ π‘–π‘›π‘‘π‘’π‘”π‘Ÿπ‘Žπ‘™:
0
𝑇
𝐸
𝑛−1
𝑍𝑑 𝑑𝑍𝑑 = 𝐸 lim
βˆ†π‘‘→0
0
𝑍𝑑 𝑑𝑍𝑑 . Its expected value is ∢
𝑛−1
𝑍𝑑𝑖 𝑍𝑑𝑖+1 − 𝑍𝑑𝑖
= lim
𝑖=0
βˆ†π‘‘→0
𝐸[𝑍𝑑𝑖 (𝑍𝑑𝑖+1 −𝑍𝑑𝑖 )]
𝑖=0
Since 𝑍𝑑𝑖 is Brownian motion in the time interval [0,ti] and 𝑍𝑑𝑖+1 − 𝑍𝑑𝑖 is Brownian motion in the time
interval [ti,ti+1], and these time intervals are non-overlapping, then they are independent random
variables. Thus:
𝐸[𝑍𝑑𝑖 (𝑍𝑑𝑖+1 −𝑍𝑑𝑖 )] = 𝐸[𝑍𝑑𝑖 ]𝐸[𝑍𝑑𝑖+1 − 𝑍𝑑𝑖 ] = 0
This result suggests the following more general theorem:
𝑇
𝐸
𝑍𝑑 𝑑𝑍𝑑 = 0
0
Significant Results based on Stochastic Integration
ο‚— For any integrable stochastic process ft that is adapted to the same
filtration {ℱ𝑑 } as standard Brownian motion Zt, we have
𝑇
𝐸
𝑓𝑑 𝑑𝑍𝑑 | ℱ𝑠 = 0
𝑠
whenever s < T
ο‚— If dXt = ftdZt, then Xt is a martingale, where ft is an integrable stochastic
process that is adapted to the same filtration {ℱ𝑑 } as a standard
Brownian motion Zt.
Itô Isometry
If ft is a square integrable stochastic process, then
2
𝑇
𝐸
𝑓𝑑 𝑑𝑍𝑑 )
0
𝑇
𝐸 𝑓𝑑2 𝑑𝑑.
=
0
Characteristics of Integrals of Real-Valued Functions
with Respect to Brownian Motion
If f(s) is a real-valued function of s, then for each t the random variable Xt
defined by:
𝑑
𝑋𝑑 =
𝑓 𝑠 𝑑𝑍𝑠
0
has a normal distribution with mean 0 and variance
𝑑
[𝑓 𝑠 ]2 𝑑𝑠
0
Proof
We prove this last theorem by noting that the random variable Xt can be
approximated by the following sum:
𝑋𝑑 ≈
𝑛−1
𝑖=0 𝑓(𝑠𝑖 )(𝑍𝑠𝑖+1
where 0 = s0 < s1 < …< sn = t, and si+1 - si =t/n.
− 𝑍𝑠𝑖 )
Since 𝑍𝑠𝑖+1 − 𝑍𝑠𝑖 for i from 0 to n-1 are pairwise independent random
variables, each with mean 0, then the sum above is a random variable
that has a normal distribution with mean 0 and variance
𝑑
𝑛−1
𝑓 𝑠𝑖
𝑖=0
2
[𝑓 𝑠 ]2 𝑑𝑠
0
By the fundamental theorem of calculus, in the limit as n approaches ∞ the variance approaches
above.
Converting Certain Stochastic Processes to Martingales
Martingale processes and deterministic processes are normally much
easier to analyze and manipulate than other stochastic processes
containing both drift and random elements. One major reason for this is
that working with martingales free us from having to calculate unknown
risk-adjusted discount rates and risk premiums in the valuation process.
One type of procedure to convert drift processes to martingales involves
changing the probability measure.
The Radon-Nikodym Derivative
ο‚— The Radon-Nikodym process is useful for converting physical
probability measures to equivalent martingales. It is often easier to
price a security under a particular measure, such as an equivalent
measure that produces a martingale than it is to price a security under a
physical probability that is not a martingale. This process is essentially
converting from a pricing framework in which investors have risk
preferences to risk-neutral pricing framework.
ο‚— The Radon-Nikodym derivative is a "bridge" that converts one
probability measure to an equivalent measure.
The Radon-Nikodym Derivative
Let β„™ and β„š be two probability measures on (,β„±) such that β„š ~ β„™ (β„š is
π‘‘β„š
equivalent to β„™) on β„±. Then there exists a random variable
known as the
𝑑ℙ
Radon-Nikodym derivative of β„š with respect to β„™ on β„± that is defined by the
property:
πœ”∈πœ™
π‘‘β„š
πœ” 𝑑 β„™ ω = β„š()
𝑑ℙ
for every measurable set  οƒŽ β„±. β„š() denotes the probability that the event 
occurs with respect to the probability measure β„š.
Properties of the Radon-Nikodym Derivative
1.
q() =
2.
π‘‘β„š
𝑑ℙ
3.
4.
p() for any ο·οƒŽο— as long as p() > 0
 ο‚³ 0 ∀ ο·οƒŽ 
πœ”∈٠𝑝
π‘‘β„š
𝑑ℙ
π‘‘β„š
𝑑ℙ
ω
π‘‘β„š
𝑑ℙ
ω =1
is a measurable function on the σ-algebra β„±
The notation p() refers to the β„™ probability that the outcome  occurs,
and an analogous interpretation of q().
Properties of Radon-Nikodym Derivative (continued)
If the probability measures are continuous, then the sum in property 3
is replaced with an integral and p() is replaced with 𝑝 π‘₯ 𝑑π‘₯ where
p(x) is the density function for the probability space β„™.
2. If the sample space  has at most a countable number of elements and
π‘‘β„š
π‘ž(πœ”)
p(ω) > 0 for every ω οƒŽ Ω, then
πœ” =
for every  οƒŽ .
1.
𝑑ℙ
𝑝(πœ”)
3. If the probability measures β„™ and β„š are continuous with continuous
density functions, say p(x) and q(x) > 0, then the Radon-Nikodym
π‘‘β„š
π‘ž(π‘₯)
derivative is simply
π‘₯ =
.
𝑑ℙ
𝑝(π‘₯)
The Radon-Nikodym Derivative t()
ο‚— The Radon-Nikodym derivative t() of β„š with respect to β„™ on a
filtration {ℱ𝑑 } at time t is defined as
πœ‰π‘‘ πœ” 𝑑 β„™ ω = β„š 
for every measurable set  οƒŽ ℱ𝑑
πœ”∈πœ™
ο‚— In the case that the sample space Ω is finite, it is sufficient to define the
Radon-Nikodym derivative πœ‰π‘‘ for any outcome ω Ο΅ Ω:
π‘ž  = πœ‰π‘‘  𝑝 
The Radon-Nikodym Derivative t() (Continued)
ο‚— For a given stochastic process Xt that is a price of a security, we will choose β„™
as the physical probability measure and β„š is the risk-neutral probability
measure. Intuitively, the Radon-Nikodym derivative t() is the ratio of the
risk-neutral probability (or density) to the physical probability (or density)
associated with the filtration ℱ𝑑 at time t.
ο‚— The Radon-Nikodym derivative for each fixed time t is a random variable
itself because it is a measurable function on the probability space β„™. It tells us
how we transform β„™ to obtain an equivalent probability measure β„š.
Furthermore, πœ‰π‘‘ is a stochastic process since it is also a function of t. If Xt is a
martingale with respect to β„š, we call β„š either the risk neutral probability
measure or the equivalent martingale measure.
Illustration 1: The Radon-Nikodym Derivative and 2 Coin
Tosses
Suppose that at time 0 we pay 2.2 to toss a coin twice, with time 1 payoffs given as follows: SHH
= 4, SHT = 3, STH = 2 and STT = 1, with probabilities given by the following physical probability
measure β„™ and risk-neutral probability measure β„š:
p(HH) = .25; p(HT) = .25; p(TH) = .25; p(TT) = .25
q(HH) = .16; q(HT) = .24; q(TH) = .24; q(TT) = .36.
Time 1 values for the Radon-Nikodym derivative are calculated as follows:
πœ‰ 𝐻𝐻 =
π‘ž(𝐻𝐻)
𝑝(𝐻𝐻)
πœ‰ 𝑇𝐻 =
π‘ž(𝑇𝐻)
𝑝(𝑇𝐻)
=
=
.16
.25
.24
.25
= .64
πœ‰ 𝐻𝑇 =
π‘ž(𝐻𝑇)
𝑝(𝐻𝑇)
= .96
πœ‰ 𝑇𝑇 =
π‘ž(𝑇𝑇)
𝑝(𝑇𝑇)
=
=
.24
.25
.36
.25
= .96
= 1.44.
Illustration 1 (Continued)
Expected values for time 1 payoffs under probability measures β„™ and β„š
are calculated as follows:
𝐸ℙ 𝑆 = 𝑖 𝑆𝑖 𝑝𝑖 = 4ο‚΄. 25 + 3ο‚΄. 25 + 2ο‚΄. 25 + 1ο‚΄. 25 = 2.5
πΈβ„š 𝑆 =
𝑆𝑖 π‘žπ‘– =
𝑖
𝑆𝑖 πœ‰π‘– 𝑝𝑖
𝑖
= 4ο‚΄. 64ο‚΄. 25 + 3ο‚΄. 96ο‚΄. 25 + 2ο‚΄. 96ο‚΄. 25 + 1ο‚΄1.44ο‚΄. 25 = 2.2
Note that the process S is a martingale to risk-neutral measure β„š; that is, we pay at time zero for
the toss exactly its expected time 1 payoff in this one time period scenario.
Illustration 1 (Continued)
In this illustration, we changed our probability measure while changing both its
expected value and variance. The variances of the two distributions are:
𝐸ℙ 𝑆 − 𝐸ℙ 𝑆
= 4 − 2.5
2
=
2 ο‚΄. 25
πΈβ„š 𝑆 − πΈβ„š 𝑆
2
𝑆𝑖 −𝐸ℙ 𝑆
2𝑝
𝑖
𝑖
+ 3 − 2.5 2 ο‚΄. 25 + 2 − 2.5 2 ο‚΄. 25 + 1 − 2.5 2 ο‚΄. 25 = 1.25
=
𝑆𝑖 − πΈβ„š 𝑆
𝑖
2
π‘žπ‘–
= 4 − 2.2 2 ο‚΄. 16 + 3 − 2.2 2 ο‚΄. 24 + 2 − 2.2 2 ο‚΄. 24 + 1 − 2.2 2 ο‚΄. 36 = 1.2.
Illustration 2: Calculating Risk-Neutral Probabilities and the
Radon-Nikodym Derivative
Suppose that we pay S0 = 6 to participate in a gamble, with potential
payoffs given as follows: S1 = 3, S2 = 7 and S3 = 11, with probabilities given
by the following physical probability measure β„™:
p1 = 1 3; p2 = 1 3; p3 = 1 3 .
Obviously, 𝐸ℙ 𝑆 = 7; the variance equals 𝐸ℙ [ 𝑆 − 𝐸ℙ 𝑆 2 ] =102 3. How
would we obtain the risk-neutral probability measure β„š from physical
probability measure β„™ without knowing the risk-neutral probabilities in
advance?
Illustration 2 (Continued)
Consider the following three statements, which state that the risk-neutral
probabilities must sum to one, the expected value of the β„š distribution must
equal S0 = 6 (it is an equivalent martingale to β„™) and the variance of both
distributions will equal 102 3:
π‘žπ‘– = π‘ž1 + π‘ž2 + π‘ž3 = 1
𝑖
𝑆𝑖 π‘žπ‘– = 3π‘ž1 + 7π‘ž2 + 11π‘ž3 = πΈβ„š 𝑆 = 6
𝑖
πΈβ„š 𝑆 − πΈβ„š 𝑆
2
=
𝑆𝑖 − πΈβ„š 𝑆
𝑖
2
π‘žπ‘– = 3 − 6 2 π‘ž1 + 7 − 6 2 π‘ž2 + 11 − 6 2 π‘ž3 = 10 2 3
Illustration 2 (Continued)
We solve the system simultaneously as follows:
1
1 1 1 π‘ž1
6
3 7 11 π‘ž2 =
10 2 3
9 1 25 π‘ž3
1
π‘ž1
1.28125 −.1875 .03125
.48958333
6
= π‘ž2 = .27083333
.1875
.125
−.0625
π‘ž3
−.46875 .0625 .03125 10 2 3
.23958333
Thus, values of the Radon-Nikodym derivative are 1, 2 and 3 are:
πœ‰1 =
π‘ž1
.489583
π‘ž2
.270833
π‘ž3
.239583
=
= 1.46875 πœ‰2 =
=
= .8125 πœ‰3 =
=
= .71875
𝑝1
.333333
𝑝2
.333333
𝑝3
.333333
The Radon-Nikodym Derivative And Binomial Pricing
Consider a binomial model over one time period with the price of a security potentially
increasing from price S0 to uS0 with probability pu and decreasing to price dS0 with
probability pd = 1 – pu. The expected value of the price of the stock at time 1 is simply:
𝐸ℙ 𝑆1 = 𝑒𝑆0 𝑝𝑒 + 𝑑𝑆0 𝑝𝑑
However, it might well be that Eβ„™ [S1] ≠ S0, which means that the security price is not a
martingale. However, we can use the equivalent probabilities for pricing:
such that
1−𝑑
π‘žπ‘’ =
and π‘žπ‘‘ = 1 − π‘žπ‘’
𝑒−𝑑
Eβ„š[S1] = S0 = uS0qu + dS0qd
After a change in the numeraire, these arbitrage-free probabilities will ensure consistent pricing in
which πΈβ„š 𝑆1 = 𝑆0 in terms of the riskless bond.
The Radon-Nikodym Derivative And Binomial Pricing
(Continued)
It can be very useful to represent the change from the physical probabilities β„™ = {p1, p2}
to the equivalent probabilities β„š = {q1, q2} by means of the Radon-Nikodym derivative
π‘‘β„š
, and use these derivatives to represent the expected value of the security S1 with
𝑑ℙ
respect to the probabilities from space β„š in terms of the appropriate expectation with
respect to the probabilities from space β„™. In our illustration above, we would have:
π‘žπ‘’
π‘‘β„š
𝑑ℙ
=
𝑝𝑒
π‘žπ‘‘
𝑝𝑑
𝑖𝑓 𝑆1 = 𝑒𝑆0
𝑖𝑓 𝑆1 = 𝑑𝑆0.
Observe that
𝐸ℙ
π‘‘β„š
π‘žπ‘’
π‘žπ‘‘
𝑆1 = 𝑒𝑆0 𝑝𝑒 + 𝑑𝑆0 𝑝𝑑 = π‘žπ‘’ 𝑒𝑆0 + π‘žπ‘‘ 𝑑𝑆0 = πΈβ„š 𝑆1
𝑑ℙ
𝑝𝑒
𝑝𝑑
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment
Recall from Chapter 5 our 2-time period binomial model illustration
where we converted from 1-period physical probabilities (all upjumps had
physical probabilities of .6) to 1-period risk-neutral probabilities (all upjumps had risk neutral probabilities of .75). This 2-time period model had
the sample space Ω = {uu,ud,du,dd}.
The time 2 physical probability measure β„™ was defined as p(uu)= (.6)2 =
.36, p(ud) = (.6)(.4) = .24, p(du)= (.4)(.6) = .24, and p(dd)= (.4)2 = .16.
The risk neutral probability measure β„š was defined as q(uu)= (.75)2 =
.5625, q(ud) = (.75)(.25) = .1875, q(du) = (.25)(.75) = .1875, and q(dd) =
(.25)2 = .0625
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
At time 2, we have complete knowledge of the possible outcomes in the
sample space Ω, so that p2() =p() and q2() = q() for every ο·οƒŽ . The
Radon-Nikodym derivative ξ2 with respect to the σ-algebra β„±2 is:
π‘ž2 (𝑒𝑒) .5625
π‘ž2 (𝑒𝑑) .1875
πœ‰2 𝑒𝑒 =
=
= 1.5625, πœ‰2 𝑒𝑑 =
=
= .78125
𝑝2 (𝑒𝑒)
.36
𝑝2 (𝑒𝑑)
.24
π‘ž2 (𝑑𝑒) .1876
π‘ž2 (𝑑𝑑) ,0625
πœ‰2 𝑑𝑒 =
=
= .78125, πœ‰2 𝑑𝑑 =
=
= .390625.
𝑝2 (𝑑𝑒)
.24
𝑝2 (𝑑𝑑)
.16
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
One can use the values of 2(ω) to compute the probability of any event 
οƒŽ β„±2 with respect to probability measure β„š in terms of the probability
measure β„™. For example, suppose that a given event  ={uu,ud,dd}. Then,
β„š() for  ={uu,ud,dd} is:

β„š  =
πœ‰2 πœ” 𝑝2 πœ” = 1.5625 .36 + .78125 .24 + .390625 .16
πœ”πœ–
= .5625 + .1875 + .0625 = .8125
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
Next, we find the Radon Nikodym derivative 1() with respect to the σ-algebra
β„±1 . Here, we only have knowledge up to time 1. We can only know whether
there has been an upjump or downjump at time 1. We don’t know what will
happen at time 2. So, we can’t distinguish uu from ud, nor du from dd. So the
σ-algebra for possible events at time 1 is:
β„±1 = {, 𝑒𝑒, 𝑒𝑑 , 𝑑𝑒, 𝑑𝑑 , }.
The event {uu,ud} means that an upjump (u) occurred at time 1, but we don’t
know what the jump will be at time 2. The event {du,dd} means that an
downjump (d) occurred at time 1, and either a u or a d can occur at time 2.
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
So, the only possible distinguishable outcomes that can occur at time 1 is
either u or d. The probabilities for these two outcomes with respect to
probability measures β„™ and β„š are: p1(u) = .6, p1(d) = .4, q1(u) = .75, and
q1(d) = .25. The Radon Nikodym derivative πœ‰1 πœ” with respect to β„±1 is:
π‘ž1 (𝑒) .75
π‘ž1 (𝑑) .25
πœ‰1 𝑒 =
=
= 1.25 π‘Žπ‘›π‘‘ πœ‰1 𝑑 =
=
= .625
𝑝1 (𝑒)
.6
𝑝1 (𝑑)
.4
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
Expected values for time 1 stock prices (recall that S0 = 10, u = 1.5 and d = .5) under
probability measures β„™ and β„š are calculated as follows:
𝐸ℙ 𝑆1 =
𝑆1 πœ” 𝑝1 (πœ”) = (1.5ο‚΄10) .6 + (.5ο‚΄10)(, 4) = 11
πœ”∈{𝑒,𝑑}
πΈβ„š 𝑆1 =
𝑆1 πœ” π‘ž1 (πœ”) =
πœ”∈{𝑒,𝑑}
𝑆1 (πœ”)πœ‰1 πœ” 𝑝1 (πœ”)
πœ”∈{𝑒,𝑑}
= 1.5ο‚΄10 1.25 (.6) + .5ο‚΄10 .625 (.4) = 12.50
Note that while probability measure β„™ prices the stock expected value in terms of currency, probability measure β„š prices using
riskless bonds (with face value $1) as the numeraire. The price of the stock under risk-neutral measure β„š is 12.5 bonds, each worth
$0.8 at time zero. Discounting these bonds reveals that the monetary value of the stock is $0.8ο‚΄12.50 = $10.00
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
Similarly, we compute expected values for time 2 stock prices under
probability measures β„™ and β„š as follows (note that p(uu) = .36, p(ud) =
p(du) = .24. p(dd) = .16, u2 = 2.25, ud = du = .75 and d2 = .25):
𝐸ℙ 𝑆2 =
𝑆2 πœ” 𝑝2 (πœ”) = 1.5ο‚΄1.5ο‚΄10 .36 + 1.5ο‚΄. 5ο‚΄10 .24
πœ”∈Ω
+ .5ο‚΄1.5ο‚΄10 .24 + (.5ο‚΄. 5ο‚΄10)(.16) = 1.21.
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
πΈβ„š 𝑆2 =
𝑆2 πœ” π‘ž2 (πœ”) =
πœ”∈Ω
𝑆2 πœ” πœ‰2 (πœ”)𝑝2 (πœ”)
πœ”∈Ω
(1.25)2 .36
= 1.5ο‚΄1.5ο‚΄10
+ .5ο‚΄1.5ο‚΄10 .625
2
+ 1.5ο‚΄. 5ο‚΄10 (.78125) .24
.24 + .5ο‚΄. 5ο‚΄10 .390625 .16 = 15.625.
where 2(u,u) = 1.252, 2(u,d) = 2(d,u) = 1.25ο‚΄.625 and 2(d,d) = .6252.
These results are depicted in the figure on the next slide. The price of the stock under
risk-neutral measure β„š is 15.625 bonds, each worth $0.64 (there are 2 time periods
before payoff) at time zero. Discounting these bonds reveals that the monetary value of
the stock is $0.64ο‚΄15.625 = $10.00. Thus, at time 2, the discounted value of the stock
once again exhibits the martingale property with respect to β„š
Radon-Nikodym Derivatives In A Multiple Period Binomial
Environment (Continued)
u
Time 0
uuS0 = 22.5, 2(uu) = 1.5625
ud
udS0 = 7.5, 2(ud) = .78125
du
duS0 = 7.5, 2(du) = .78125
dd
ddS0 = 2.5, 2(dd) = .390625
uS0 = 15
1(u)=1.25
S0=10
d
uu
dS0 = 5
1(d)=.625
Time 1
Time 2
Radon-Nikodym Derivatives In A Multiple Period
Binomial Environment (Continued)
These concepts can be extended to multiple time periods in the binomial situation as
well as to cases when there are more than two possible choices at each stage. Suppose
ps,t() is the physical probability that, given the stock price Ss at time s, it will have the
value St() at time t for any outcome . Let qs,t() be the equivalent probability that,
given the stock price Ss at time s, it will have the value St() at time t for any outcome ,
so that the martingale property is satisfied. To make this precise, define the Radonπ‘ž ()
π‘‘β„š
Nikodym derivative () = 𝑠,𝑑 . This gives:
𝑑ℙ
𝑝𝑠,𝑑 ()
𝑛
π‘žπ‘ ,𝑑 ()
π‘‘β„šs,t
𝐸ℙ
𝑆 𝑆 =
𝑆𝑠,𝑑 𝑝𝑠,𝑑 =
𝑆𝑠,𝑑 π‘žπ‘ ,𝑑 () = πΈβ„š 𝑆𝑑 𝑆𝑠 = 𝑆𝑠
𝑑ℙs,t 𝑑 𝑠
𝑝𝑠,𝑑 ()
πœ”
In this case, the Radon-Nikodym derivative
𝑖=1
π‘‘β„šπ‘ ,𝑑
𝑑ℙ𝑠,𝑑
is the multiplicative factor of the stochastic process St that
converts the expectation on a probability space β„™ to a probability space β„š, so that the stochastic process St
becomes a martingale in the probability space β„š. We point out that the Radon-Nikodym derivative is a function
of the times s and t
Radon-Nikodym Derivatives And Multiple Time Periods
Suppose that there is a physical probability p0,1 to go from a given price S0 at
time 0 to a particular price S1 at time 1. Of course, ΔS1 = S1 – S0 is a random
variable and there can be a countable number of possible values of ΔS1 with
each particular choice of its value associated with a particular probability p0,1.
Assume that the change in the price of the security at time t, ΔSt, is independent
of the change of the price ΔSs at any other time s. Thus, the price of the stock Sn
at time n is a sum of n independent random variables ΔS1, ΔS2,…, ΔSn. The
probability p0,n that the price takes on the value Sn by going through a particular
choice of price changes ΔS1, ΔS2,…, ΔSn equals
𝑛
𝑝0,𝑛 = 𝑝0,1 𝑝1,2 … 𝑝𝑛−1,𝑛 =
𝑝𝑖−1,𝑖
𝑖=1
Radon-Nikodym Derivatives And Multiple Time Periods
(Continued)
Now, suppose that qi-1,i is the equivalent probability measure to replace pi-1,i such that
the price Si has the martingale property going from time i-1 to time i: Eβ„š[Si|Si-1] = Si-1.
The equivalent probability measure that the price goes from S0 to Sn equals
π‘ž0,𝑛 = π‘ž0,1 π‘ž1,2 … π‘žπ‘›−1,𝑛 =
𝑛
𝑖=1 π‘žπ‘–−1,𝑖
=
𝑛 π‘žπ‘–−1,𝑖
𝑝𝑖−1,𝑖
𝑖=1 𝑝
𝑖−1,𝑖
=
𝑛 π‘žπ‘–−1,𝑖
𝑖=1 𝑝
𝑖−1,𝑖
𝑛
𝑖=1 𝑝𝑖−1,𝑖
=
𝑛 π‘žπ‘–−1,𝑖
𝑖=1 𝑝
𝑖−1,𝑖
𝑝0,𝑛 = πœ‰π‘› 𝑝0,𝑛
This shows that the Radon-Nikodym derivative to use to change the measure from time
0 to time n is
𝑛
πœ‰π‘› =
𝑖=1
π‘žπ‘–−1,𝑖
.
𝑝𝑖−1,𝑖
Choosing the measure β„š appropriately, we can arrange that Sn has the martingale
property:
Eβ„š[Sn|S0] = S0
Illustration: The Binomial Pricing Model
Suppose the physical probability measure β„™ is associated with a given
value p and p≠(1- d)/(u-d). Then, as we showed earlier, the risk neutral
probability measure β„š is associated with the value q=(1- d)/(u-d), where q
replaces p in the equations for the binomial pricing model. This will give
Eβ„š[Sn|Sm]=Sm for any m < n. Next, we find the Radon-Nikodym derivative
associated with this change of measure from β„™ to β„š at time n. The original
probability measure β„™i-1,i for the price change from time i-1 to i associated
with the value p is
ℙ𝑖−1,𝑖
𝑝 𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘–π‘›π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑒
=
1 − 𝑝 𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘‘π‘’π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑑.
Illustration: The Binomial Pricing Model (Continued)
The risk-neutral probability measure β„ši-1,i for the price change from time i-1 to
time i associated with the value q is
β„šπ‘–−1,𝑖
π‘ž 𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘–π‘›π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑒
=
1 − π‘ž 𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘‘π‘’π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑑
where q=(1- d)/(u-d).
The Radon-Nikodym derivative for the change of measure is
π‘‘β„šπ‘–−1,𝑖
𝑑ℙ𝑖−1,𝑖
π‘ž
𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘–π‘›π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑒 π‘“π‘Ÿπ‘œπ‘š π‘‘π‘–π‘šπ‘’ 𝑖 − 1 π‘‘π‘œ 𝑖
𝑝
= 1−π‘ž
𝑖𝑓 π‘‘β„Žπ‘’ π‘π‘Ÿπ‘–π‘π‘’ π‘‘π‘’π‘π‘Ÿπ‘’π‘Žπ‘ π‘’π‘  𝑏𝑦 π‘‘β„Žπ‘’ π‘“π‘Žπ‘π‘‘π‘œπ‘Ÿ 𝑑 π‘“π‘Ÿπ‘œπ‘š π‘‘π‘–π‘šπ‘’ 𝑖 − 1 π‘‘π‘œ 𝑖.
1−𝑝
Illustration: The Binomial Pricing Model (Continued)
Since the changes in the price over the different time intervals are independent of one
another, then the Radon-Nikodym derivative (n) of the price changes from time 0 to n
equals
𝑛
π‘‘β„šπ‘–−1,𝑖
πœ‰π‘› =
𝑑ℙ𝑖−1,𝑖
𝑖=1
If there were exactly k increases and n-k decreases in the price from time 0 to n then
π‘˜
𝑛−π‘˜
π‘ž
1−π‘ž
πœ‰π‘› =
𝑝
1−𝑝
This is the Radon-Nikodym derivative for change of measure for the binomial pricing
model.
Illustration: The Binomial Pricing Model (Continued)
The probability of obtaining any one particular outcome of k upjumps and n-k
downjumps equals pk(1-p)n-k in probability measure β„™ and equals qk(1-q)n-k in
probability measure β„š. Notice that n provides the correct multiplicative factor to the
change the measure from β„™ to β„š since
π‘ž
π‘˜
𝑛−π‘˜
πœ‰π‘› 𝑝 (1 − 𝑝)
=
𝑝
π‘˜
1−π‘ž
1−𝑝
𝑛−π‘˜
π‘π‘˜ (1 − 𝑝)𝑛−π‘˜ = π‘žπ‘˜ (1 − π‘ž)𝑛−π‘˜
So, the probability of obtaining exactly k upjumps and n-k downjumps in the price
from time 0 to n equals
𝑛 π‘˜
with respect to probability measure β„™, and equals
𝑝 (1 − 𝑝)𝑛−π‘˜
π‘˜
𝑛 π‘˜
π‘ž (1 − π‘ž)𝑛−π‘˜
π‘˜
with respect to probability measureβ„š
Change Of Normal Density
Consider a random variable X with the normal distribution X ~ N(,
1). Denote its density function under probability measure β„™ as:
𝑓ℙ π‘₯ =
1
2πœ‹
𝑒
−(π‘₯−)2 /2
For any continuous function F, the expectation of the random
variable F(X) is given by:
∞
1
−(π‘₯−)2 /2
𝐸 𝐹(𝑋) =
𝐹(π‘₯)𝑒
𝑑π‘₯
β„™
2πœ‹
−∞
If F(X) = X, we recognize this as the mean 𝐸ℙ 𝑋 = πœ‡. If F(X)=(X-μ)2, we recognize this as the variance
𝐸ℙ (𝑋 − πœ‡)2 = 1.
Change Of Normal Density (Continued)
Suppose that we were to multiply both sides of the density function
2 /2
−πœ‡π‘₯+πœ‡
𝑓ℙ π‘₯ by 𝑒
:
πœ‡2
πœ‡2
π‘₯−πœ‡ 2
1
𝑒 −πœ‡π‘₯+ 2 𝑓ℙ π‘₯ = 𝑒 −πœ‡π‘₯+ 2 βˆ™
𝑒− 2
2πœ‹
1 − π‘₯ 2 − πœ‡2 + 2πœ‡π‘₯ +πœ‡π‘₯+ πœ‡2
1 −π‘₯ 2
2
2 =
=
𝑒 2 2
𝑒 2
2πœ‹
2πœ‹
2
By multiplying both side of our density function in β„™-space by 𝑒 −πœ‡π‘₯+πœ‡ /2 , we produced a
new probability measure for our rand variable X such that the new random variable still
has normal distribution with variance 1, but with a mean of 0 as we will see in next slide.
Change Of Normal Density (Continued)
Now let β„š be a new measure , and its density function is:
π‘“β„š π‘₯ =
1
2πœ‹
π‘₯2
𝑒− 2
For any continuous function F, the expectation of F(X) with respect to β„š :
πΈβ„š 𝐹(𝑋) =
1
2πœ‹
∞
2 /2
−
π‘₯
𝐹(π‘₯)𝑒
𝑑π‘₯
−∞
If F(X)=X, we recognize this as the mean πΈβ„š 𝑋 = 0 . If F(X)=(X-μ)2, we recognize this as the
variance πΈβ„š (𝑋 − πœ‡)2 = 1.
Change Of Normal Density (Continued)
we define a change of measure for a continuous probability space as follows:
π‘“β„š π‘₯ =
1
2πœ‹
π‘₯2
𝑒− 2
=
2 /2
−

π‘₯+

𝑒
𝑓ℙ
π‘‘β„š
π‘₯ = πœ‰ π‘₯ 𝑓ℙ π‘₯ =
𝑓ℙ π‘₯
𝑑ℙ
Here, we have defined the Radon-Nikodym derivative for our change of
measure:
2

π‘‘β„š
π‘₯ = πœ‰ π‘₯ = 𝑒 −π‘₯+ 2
𝑑ℙ
Illustration: Change Of Normal Density
Suppose that we were to begin with a random variable X with mean μ
under probability measure β„™, and we seek to transform this measure to an
equivalent probability measure β„š with a mean of zero by employing the
change of measure process that we just worked through. More
specifically, suppose that we wish to change a measure with distribution
N(.5, 1) to another with distribution N(0, 1)
The normal density p(x) associated with value x, along with RadonNikodym derivatives and equivalent densities q(x) are computed as
follows:
𝑓ℙ π‘₯ =
1
2πœ‹
𝑒−
π‘₯−.5 2 /2
2 /2
and πœ‰ π‘₯ = 𝑒 −.5π‘₯+.5
π‘ž π‘₯ = π‘“β„š π‘₯ = 𝑓ℙ π‘₯ πœ‰ π‘₯
Illustration: Change Of Normal Density (Continued)
Change of Normal Density
More General Shifts
Now consider a normal density function for a random variable X ~ N(, 2)
under probability measure β„™:
𝑓ℙ π‘₯ =
1
𝜎 2πœ‹
2
1 π‘₯−
−2 𝜎
𝑒
Suppose that we wish to shift our mean left ( > 0) or right ( < 0) by some
arbitrary constant . By multiplying both sides of our density function in β„™πœ†π‘₯ ο­πœ†
πœ†2
− 2+ 2 − 2
space by 𝑒 𝜎 𝜎 2𝜎 , we will produce a new probability measure for our
random variable X, which has a normal distribution with mean - and
unchanged variance 2. (see next slide)
More General Shifts (Continued)
We will call this new measure β„š, and its density function is:
π‘“β„š π‘₯ =
1
𝜎 2πœ‹
2
1 π‘₯−(−πœ†)
−
𝜎
𝑒 2
π‘‘β„š
= πœ‰ π‘₯ 𝑓ℙ π‘₯ =
𝑓ℙ π‘₯
𝑑ℙ
We have defined the Radon-Nikodym derivative for our change of measure:
π‘‘β„š
π‘₯ =πœ‰ π‘₯ =𝑒
𝑑ℙ
πœ†π‘₯ ο­πœ† πœ†2
− 2+ 2 − 2
𝜎
𝜎
2𝜎
This change of measure allows us to shift the mean of a normal distribution without
affecting its variance.
Change Of Brownian Motion
First, consider the process Xt = Zt + μt, where Zt is a standard Brownian motion
and μ ≥ 0 is a constant or drift term. Denote β„™ as the probability space for this
process. We wish to find a change of measure that changes the mean μ while
maintaining the variance of β„™. Observe that for any fixed t, Xt ~ N(μt,t). To
simplify the notation and generalize the process, consider a random variable X
~ N(μ, σ2) in probability space β„™. We saw in the previous section that if X ~
N(μ, σ2), then the change in measure
π‘‘β„š
=𝑒
𝑑ℙ
πœ†π‘₯ ο­πœ†
πœ†2
− 2+ 2 − 2
𝜎
𝜎
2𝜎
results in the random variable X having the distribution 𝑋~ N(μ - λ,σ2) in
probability space β„š. Note that we renamed the random variable X to be 𝑋 in the
new probability space β„š.
Change Of Brownian Motion (Continued)
Next, making the replacements: X→ Zt +μt, σ2→ t, μ→ μt, and λ→ λt,
suggests that the stochastic process Zt + μt in probability space β„™ becomes
the process 𝑍𝑑 + πœ‡ − πœ† 𝑑 in the probability space β„š with the change of
measure given by the Radon-Nikodym derivative at time t:
πœ‰π‘‘ = 𝑒
(πœ†π‘‘)2
πœ†π‘‘ 𝑍𝑑 +πœ‡π‘‘ πœ‡π‘‘ πœ†π‘‘
−
+
−
𝑑
𝑑
2𝑑
=
1
−πœ†π‘π‘‘ − πœ†2 𝑑
2
𝑒
𝑍𝑑 is Brownian motion in the probability space β„š.
The Cameron-Martin-Girsanov Theorem
The Cameron-Martin-Girsanov Theorem provides conditions under
which probability measures for continuous-time stochastic processes can
be converted to risk-neutral measures. More generally, the CameronMartin-Girsanov Theorem can be used to convert one probability
measure to another of the same type with the same variance but with a
different drift.
Multiple Time Periods: Discrete Case To Continuous
Case
Suppose Xt follows a stochastic process in continuous time t such that its
differential satisfies dXt = dZt + μt dt where Zt is Brownian motion with
respect to some probability measure β„™ and μt is a continuous real-valued
function of t. We solve for Xt as follows:
𝑑
𝑋𝑑 − 𝑋0 =
𝑑
𝑑𝑍𝑠 + πœ‡π‘  𝑑𝑠
0
𝑋𝑑 = 𝑋0 + 𝑍𝑑 +
πœ‡π‘  𝑑𝑠
0
In this case, we say that the price Xt follows a Brownian motion process with variable drift
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
The above continuous process can be approximated by a discrete process in the
following way. Divide the time interval [0,t] into n equal subintervals [ti-1,ti]
with ti – ti-1 =Δt =t/n for each i = 1,2,…,n. This implies that ti = ti/n. Define
βˆ†π‘‹π‘‘π‘– = 𝑋𝑑𝑖 − 𝑋𝑑𝑖−1 . This implies that
𝑋𝑑 = 𝑋0 + βˆ†π‘‹π‘‘1 + βˆ†π‘‹π‘‘2 + β‹― + βˆ†π‘‹π‘‘π‘›
For each βˆ†π‘‹π‘‘π‘– we have
𝑑𝑖
βˆ†π‘‹π‘‘π‘– = 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 +
πœ‡π‘  𝑑𝑠 ≈ 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘
𝑑𝑖−1
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Thus, we can approximate the process Xt by:
𝑛
𝑋𝑑 ≈ 𝑋0 +
( 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘).
𝑖=1
Since 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 ~𝑁 0, βˆ†π‘‘ , 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘ ~𝑁 πœ‡π‘‘π‘–−1 βˆ†π‘‘, βˆ†π‘‘ . This means that
the random variable 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘ has the density function
1
2πœ‹βˆ†π‘‘
(π‘₯𝑖 −πœ‡π‘‘π‘–−1 βˆ†π‘‘)2
2βˆ†π‘‘
𝑒−
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Now, label the probability measure associated with this density function
by β„™i. In section 6.3.6, we demonstrated that using the Radon-Nikodym
derivative
πœ‰π‘‘ =
1
−πœ†π‘π‘‘ − πœ†2 𝑑
2
𝑒
changed the random variable Zt + μt to the random variable 𝑍𝑑 + πœ‡ − πœ† 𝑑.
In this proof, we will only take into consideration the probability
distribution features of the processes and not attempt to justify their
additional continuity and independence properties.
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Making the replacements 𝑍𝑑 → 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 , πœ‡ → πœ‡π‘‘π‘–−1 , 𝑑 → βˆ†π‘‘, π‘Žπ‘›π‘‘ πœ† → πœ†π‘‘π‘–−1 , we
conclude that the change in measure from β„™i-1,i to β„ši-1,i with the Radon-Nikodym
derivative:
1
π‘‘β„šπ‘–−1,𝑖
−πœ†π‘‘π‘–−1 (𝑍𝑑𝑖 −𝑍𝑑𝑖−1 )−2πœ†2𝑑 βˆ†π‘‘
𝑖−1
=𝑒
𝑑ℙ𝑖−1,𝑖
will change the random variable 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘ to the random variable
𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + (πœ‡π‘‘π‘–−1 − πœ†π‘‘π‘–−1 )βˆ†π‘‘ where 𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 is Brownian motion with respect
to the probability measure β„ši-1,i. {𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘}𝑛𝑖=1 are independent
random variables since { 𝑑𝑖−1 , 𝑑𝑖 }𝑛𝑖=1 are non-overlapping intervals.
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Since the density function for a sum of independent random variables equals
the product of their density functions, then the density function for
𝑛
(𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘)
𝑖=1
is the product:
𝑛
𝑖=1
1
2πœ‹βˆ†π‘‘
𝑒
(π‘₯𝑖 −πœ‡π‘‘π‘–−1 βˆ†π‘‘)2
−
2βˆ†π‘‘
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Next, label the probability measure associated with this density function as β„™n.
Since {𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 }𝑛𝑖=1 are independent random variables, then the Radonπ‘‘β„š
Nikodym derivative 𝑛 to use to go from the random variable
𝑑ℙ𝑛
𝑛
𝑖=1(𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + πœ‡π‘‘π‘–−1 βˆ†π‘‘) with respect to the probability measure β„™n to the
random variable 𝑛𝑖=1[𝑍𝑑𝑖 − 𝑍𝑑𝑖−1 + (πœ‡π‘‘π‘–−1 − πœ†π‘‘π‘–−1 )βˆ†π‘‘] with respect to the
probability measure β„šn must be the product of the Radon-Nikodym derivatives
π‘‘β„šπ‘–−1,𝑖
at each time increment from ti-1,i to ti extending from 1 to n:
𝑑ℙ𝑖−1,𝑖
π‘‘β„šn
=
𝑑ℙn
𝑛
𝑖=1
π‘‘β„šπ‘–−1,𝑖
=
𝑑ℙ𝑖−1,𝑖
𝑛
1
−πœ†π‘‘π‘–−1 (𝑍𝑑𝑖 −𝑍𝑑𝑖−1 )− πœ†2𝑑 βˆ†π‘‘
2 𝑖−1
𝑒
𝑖=1
=𝑒
1 𝑛
2
− 𝑛
πœ†
(𝑍
−𝑍
)−
𝑑𝑖−1
𝑖=1 𝑑𝑖−1 𝑑𝑖
𝑖=1 πœ†π‘‘
2
𝑖−1
βˆ†π‘‘
Multiple Time Periods: Discrete Case To Continuous
Case (Continued)
Referring to the definitions of stochastic integration in Section 6.1.2, in
the limit as n → ∞, we obtain
πœ‰π‘‘ =
𝑑
1 𝑑
− 0 πœ†π‘  𝑑𝑍𝑠 − 2 0 πœ†2𝑠 𝑑𝑠
𝑒
.
This is the Radon-Nikodym derivative required for the change of measure
in the Cameron-Martin-Girsanov Theorem (see the following section) in
the event that μt and λt are continuous real-valued functions. This result
can be generalized to any integrable stochastic processes μt and λt, which
is the statement of the Cameron-Martin-Girsanov Theorem.
The Cameron-Martin-Girsanov Theorem
The Cameron-Martin-Girsanov Theorem applies to stochastic processes Xt that
are represented as Brownian motion with drift:
𝑑𝑋𝑑 = 𝑑𝑍𝑑 + πœ‡π‘‘ 𝑑𝑑
under probability measure β„™ with μt itself as a stochastic process and 𝑍𝑑 is
standard Brownian motion. The theorem states that, under certain technical
conditions (e.g., a finite variance), we can shift the drift so that dXt = 𝑑 𝑍𝑑 +
πœ‡π‘‘ − πœ†π‘‘ 𝑑𝑑 in an equivalent probability space β„š using the Radon-Nikodym
derivative
πœ‰π‘‘ =
𝑑
1 𝑑
− 0 πœ†π‘  𝑑𝑍𝑠 − 2 0 πœ†2𝑠 𝑑𝑠
𝑒
Illustration: Applying The Cameron-Martin-Girsanov
Theorem
Consider the process dXt = dZt +.05dt under some probability measure β„™ with X0 = 10.
Solving for Xt gives:
𝑑
𝑋𝑑 − 𝑋0 =
𝑑
𝑑𝑋𝑠 =
0
𝑑𝑍𝑠 + .05𝑑𝑠 = 𝑍𝑑 + .05𝑑,
0
so that Xt =Zt + .05t + X0. Because of the drift term .05t, Xt is not a martingale with
respect to the probability measure β„™. Using the Cameron-Martin-Girsanov Theorem,
we will be able to find an equivalent probability measure β„š, so that the process Xt with
respect to the probability measure β„š will be a martingale. Choose λ = .05 in the
Cameron-Martin-Girsanov Theorem so that the Radon-Nikodym derivative for the
change in measure will be
πœ‰π‘‘ =
𝑑
1 𝑑
− 0 .05𝑑𝑍𝑠 − 0 (.05)2 𝑑𝑠
2
𝑒
= 𝑒 −.05𝑍𝑑−.00125𝑑
Illustration: Applying The Cameron-Martin-Girsanov
Theorem (Continued)
By the Cameron-Martin-Girsanov Theorem, this will result in the process Xt taking the
form
𝑑𝑋𝑑 = 𝑑 𝑍𝑑 + .05 − .05 𝑑𝑑 = 𝑑 𝑍𝑑
where 𝑍𝑑 is Brownian motion with respect to the probability measure β„š. Since
obviously 𝑋𝑑 = 𝑍𝑑 + 𝑋0 , then 𝑋𝑑 is a martingale with respect to the probability measure
β„š. The density function of Xt with respect to measure β„™, the Radon Nikodym derivative
and the density function of 𝑋𝑑 with respect to β„š are given as follows:
𝑓ℙ π‘₯ =
π‘“β„š π‘₯ = 𝑓ℙ π‘₯ πœ‰π‘‘ π‘₯ =
1
2πœ‹
1
2πœ‹
𝑒
−
𝑒
− π‘₯−πœ‡π‘‘ 2 /2𝑑
=
1
2πœ‹
𝑒−
π‘₯−.05𝑑 2
2𝑑
𝑒 −.05π‘₯+.00125𝑑
π‘₯−.05𝑑 2 /2𝑑
=
1
2πœ‹
𝑒
−
π‘Žπ‘›π‘‘
πœ‰π‘‘ (π‘₯) = 𝑒 −.05π‘₯+.00125𝑑
1 2
π‘₯ +.05π‘₯−.00125𝑑−.05π‘₯+.00125𝑑
2𝑑
=
1
2πœ‹
𝑒
−
1 2
π‘₯
2𝑑
The Martingale Representation Theorem
Suppose that Mt is a martingale process of the form
dMt = σtdZt + μtdt
So that σt ≠ 0 with probability 1. If Nt is another martingale with respect
to the same probability measure, then there exists a stochastic process t
such that Nt can be expressed in the form dNt = tdMt. This result will be
of use in chapter 7 in order to obtain arbitrage-free pricing of derivatives.
A Discussion On Taylor Series Expansions
To estimate the change in an infinitely differentiable function 𝑦 = 𝑦(t), we have :
dy
1 d2y
y ο€½
οƒ— t  οƒ— 2 οƒ— (t ) 2  ...
dt
2 dt
If Δt is small, higher order terms (involving Δt raised to powers 2 or greater) are negligible compared to terms
just involving Δt. So we have the following approximation when Δt is small:
𝑑𝑦
βˆ†π‘¦ =
βˆ†π‘‘
𝑑𝑑
This can also be expressed in differential notation:
𝑑𝑦 =
𝑑𝑦
𝑑𝑑
𝑑𝑑
𝑇
𝑦 𝑇 −𝑦 0 =
𝑇
𝑑𝑦 𝑑 =
0
0
𝑑𝑦
𝑑𝑑.
𝑑𝑑
Taylor Series And Two Independent Variables
Suppose that 𝑦 = 𝑦(x, t); that is, 𝑦 is a function of the independent variables x
and t. The Taylor series expansion can be generalized to two independent
variables as follows:
ο‚Άy
ο‚Άy
1 ο‚Ά2 y
1 ο‚Ά2 y
1 ο‚Ά2 y
2
2
y ο€½
οƒ— x  οƒ— t  οƒ— 2 οƒ— (x)  οƒ— 2 οƒ— (t )  οƒ—
οƒ— (xt )  ...
ο‚Άx
ο‚Άt
2 ο‚Άx
2 ο‚Άt
2 ο‚Άxο‚Άt
If x = x(t) is a differentiable function of time t, then we have the approximation Δx = x’(t)Δt.
Ignoring higher order terms in Δt, we obtain:
πœ•π‘¦ ′
πœ•π‘¦
𝑑𝑦 =
π‘₯ 𝑑 𝑑𝑑 +
𝑑𝑑
πœ•π‘₯
πœ•π‘‘
𝑇
𝑦 𝑇 −𝑦 0 =
𝑇
𝑑𝑦 𝑑 =
0
0
πœ•π‘¦ ′
πœ•π‘¦
π‘₯ 𝑑 +
𝑑𝑑
πœ•π‘‘
πœ•π‘‘
The Itô Process
Consider an Itô process, which means that the stochastic process Xt
satisfies the equation:
dX t ο€½ a( X t , t )dt  b( X t , t )dZ t
The drift of the process is a, while b2 is the instantaneous variance and
dZt is a standard Brownian motion process. Taking Δt to be a small
change in time and expressing a = a(Xt, t) and b = b(Xt, t) in order to
economize the notation, we can write:
βˆ†π‘‹π‘‘ = π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘
where random variable βˆ†π‘π‘‘ ~ N(0,t)
The Itô Process (Continued)
Now, suppose that 𝑦 = 𝑦(x, t) is an infinitely differentiable function with respect to
the real variables x and t. Now, replace x with Xt so that 𝑦 = 𝑦(𝑋𝑑, 𝑑). Thus, 𝑦 itself
becomes a stochastic process, since it is a function of a stochastic process Xt and time
t. The Taylor series expansion above can be used to estimate Δ𝑦, the change in 𝑦,
resulting from a change Δt in time:
πœ•π‘¦
πœ•π‘¦
1 πœ•2𝑦
βˆ†π‘¦ 𝑋𝑑 , 𝑑 =
βˆ†π‘‘ +
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘ + βˆ™
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘
2
πœ•π‘‘
πœ•π‘₯
2 πœ•π‘₯
1 πœ•2𝑦
+ βˆ™
2 πœ•π‘₯ πœ•π‘‘
=
2𝑦
1
πœ•
2
+ βˆ™ 2 βˆ†π‘‘
2 πœ•π‘‘
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘ βˆ†π‘‘ + β‹―
πœ•π‘¦
1 πœ•2𝑦
+
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘ + βˆ™ 2 π‘Ž2 βˆ†π‘‘ 2 + 2π‘Žπ‘βˆ†π‘‘βˆ†π‘π‘‘
πœ•π‘₯
2 πœ•π‘₯
1 πœ•2𝑦
1 πœ•2𝑦
2
+ βˆ™ 2 βˆ†π‘‘ + βˆ™
[π‘Ž βˆ†π‘‘ 2 + π‘βˆ†π‘π‘‘ βˆ†π‘‘] + β‹―
2 πœ•π‘‘
2 πœ•π‘₯ πœ•π‘‘
πœ•π‘¦
βˆ†π‘‘
πœ•π‘‘
2
+ 𝑏 2 βˆ†π‘π‘‘
2
The Itô Process (Continued)
In the expansion above, we also economized the notation for the various partial derivatives evaluated at
(x,t)=(Xt,t) by denoting:
πœ•π‘¦ πœ•π‘¦
πœ•π‘¦
πœ•π‘¦
πœ•2𝑦 πœ•2𝑦
πœ•2𝑦 πœ•2𝑦
πœ•2𝑦
πœ•2𝑦
=
𝑋 ,𝑑 ,
=
𝑋 , 𝑑 , 2 = 2 𝑋𝑑 , 𝑑 , 2 = 2 𝑋𝑑 , 𝑑 ,
=
𝑋 ,𝑑 .
πœ•π‘₯ πœ•π‘₯ 𝑑
πœ•π‘‘
πœ•π‘‘ 𝑑
πœ•π‘₯
πœ•π‘₯
πœ•π‘‘
πœ•π‘‘
πœ•π‘₯πœ•π‘‘
πœ•π‘₯πœ•π‘‘ 𝑑
When estimating Δy using the Taylor series expansion, all terms that are negligible compared to Δt as Δt
approaches 0 can be dropped. Since (Zt+Δt - Zt) ~ Z βˆ†π‘‘ with Z ∼ N(0,1), the terms involving (Δt)2, ΔtΔZt,
and higher order terms can all be dropped. Our expansion simplifies to:
πœ•π‘¦
πœ•π‘¦
𝑏2 πœ• 2 𝑦
βˆ†π‘¦ =
βˆ†π‘‘ +
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘ + βˆ™
βˆ†π‘π‘‘
πœ•π‘‘
πœ•π‘‹π‘‘
2 πœ•π‘‹π‘‘2
2
Since (ΔZt)2 ~ Z2Δt, this term is not negligible. We now state a remarkable fact. We can actually replace
(ΔZt)2 with Δt, where the random variable has seemingly disappeared. To show this, we will make use of
the fact that Brownian motion increments are independent for disjoint intervals.
Demonstration That Δt Can Replace (ΔZt)2 In The
Differential y
We begin by subdividing the interval [t, t + Δt] into n smaller subintervals [t
+(i-1)Δt/n, t + iΔt/n] for i = 1, 2, …n. This means that the width of each
subinterval equals Δt/n. In the Taylor series expansion above of Δy, the
2
subintervals lead to (ΔZt)2 being replaced by 𝑛𝑖=1 βˆ†π‘– 𝑍𝑑 with
βˆ†π‘– 𝑍𝑑 = 𝑍
π‘–βˆ†π‘‘
𝑑+ 𝑛
− 𝑍
𝑖−1 βˆ†π‘‘
𝑛
𝑑+
Since
βˆ†π‘– 𝑍𝑑 = 𝑍
π‘–βˆ†π‘‘
𝑑+
𝑛
− 𝑍
𝑑+
𝑖−1 βˆ†π‘‘
𝑛
βˆ†π‘‘
∼ 𝑁 0,
𝑛
the variance of the random variable ΔiZt is:
𝐸 βˆ†π‘– 𝑍𝑑
2
= π‘‰π‘Žπ‘Ÿ βˆ†π‘– 𝑍𝑑
βˆ†π‘‘
=
𝑛
Demonstration That Δt Can Replace (ΔZt)2 In The
Differential y (Continued)
Notice that this result gives us the expected value of the random variable
βˆ†π‘– 𝑍𝑑 2 . Next we calculate the variance of βˆ†π‘– 𝑍𝑑 2 . From the second equation
above, we see that
βˆ†π‘‘
βˆ†π‘– 𝑍𝑑 ∼ 𝑍
𝑛
with Z ∼ N(0,1)
(βˆ†π‘– 𝑍𝑑
)2
∼
𝑍2
βˆ†π‘‘
𝑛
By the last equation above and the fact that Var(Z2c) = 2c2, we obtain:
π‘‰π‘Žπ‘Ÿ (βˆ†π‘– 𝑍𝑑 )2
2(βˆ†π‘‘)2
=
𝑛2
Demonstration That Δt Can Replace (ΔZt)2 In The
Differential y (Continued)
The Brownian motion increments ΔiZt are independent random variables since
the corresponding time intervals are disjoint. Since the variance of a sum of
independent random variables is the sum of each of their variances, then:
𝑛
π‘‰π‘Žπ‘Ÿ
𝑛
βˆ†π‘– 𝑍𝑑
2
𝑛
π‘‰π‘Žπ‘Ÿ ( (βˆ†π‘– 𝑍𝑑 )2 ) =
=
𝑖=1
𝑖=1
Since we showed above that 𝐸 βˆ†π‘– 𝑍𝑑
2
=
𝑖=1
βˆ†π‘‘
,
𝑛
we see that
𝑛
𝐸
βˆ†π‘– 𝑍𝑑
𝑖=1
2
βˆ†π‘‘
= 𝑛 = βˆ†π‘‘
𝑛
2(βˆ†π‘‘)2 2(βˆ†π‘‘)2
=
2
𝑛
𝑛
Demonstration That Δt Can Replace (ΔZt)2 In The
Differential y (Continued)
Thus, the quantity 𝑛𝑖=1 βˆ†π‘– 𝑍𝑑 2 has an expected value of Δt and a variance
2(Δt)2/n that approaches 0 as n approaches infinity. Since (ΔZt)2 must
approach 𝑛𝑖=1 βˆ†π‘– 𝑍𝑑 2 as n→∞ and as Δt gets arbitrarily small, this
implies that (ΔZt)2 has an expected value of Δt and a variance
approaching 2(Δt)2/n→0. But a random variable with variance 0 is
simply equal to its expected value. So, we conclude that (ΔZt)2 = Δt.
Itô's Formula
πœ•π‘¦
πœ•π‘¦
𝑏2 πœ• 2𝑦
βˆ†π‘¦ =
βˆ†π‘‘ +
π‘Žβˆ†π‘‘ + π‘βˆ†π‘π‘‘ + βˆ™
βˆ†π‘π‘‘
2
πœ•π‘‘
πœ•π‘‹π‘‘
2 πœ•π‘‹π‘‘
πœ•π‘¦
πœ•π‘¦ 1 2 πœ• 2 𝑦
πœ•π‘¦
βˆ†π‘¦ =
+π‘Ž
+ 𝑏
βˆ†π‘‘ + 𝑏
βˆ†π‘π‘‘
2
πœ•π‘‘
πœ•π‘₯ 2 πœ•π‘₯
πœ•π‘₯
2
We can express this result in differential form, which is known as Itô's formula :
πœ•π‘¦
πœ•π‘¦ 1 2 πœ• 2 𝑦
πœ•π‘¦
𝑑𝑦 𝑋𝑑 , 𝑑 =
+π‘Ž
+ 𝑏
𝑑𝑑 + 𝑏
𝑑𝑍𝑑 .
2
πœ•π‘‘
πœ•π‘₯ 2 πœ•π‘₯
πœ•π‘₯
where the partial derivatives are evaluated at (x,t) = (Xt,t).
Itô's Lemma
Given a real-valued function: 𝑦 = 𝑦(π‘₯, 𝑑), define the stochastic process
𝑦(𝑋𝑑, 𝑑),where Xt is an Itô process dXt = a(Xt,t)dt + b(Xt,t)dZt with Zt
denoting standard Brownian motion. By Itô's formula, the differential of
the process 𝑦(𝑋𝑑, 𝑑) satisfies the equation:
πœ•π‘¦ πœ•π‘¦ 1 2 πœ• 2 𝑦
πœ•π‘¦
𝑑𝑦 𝑋𝑑 , 𝑑 = π‘Ž
+
+ 𝑏
𝑑𝑑 + 𝑏 𝑑𝑍𝑑
2
πœ•π‘₯ πœ•π‘‘ 2 πœ•π‘₯
πœ•π‘₯
where the partial derivatives above are evaluated at (x,t) = (Xt,t).
Applying Itô's Lemma
Evaluating stochastic integrals is generally trickier than evaluating
ordinary real valued integrals. We recommend the following 3-step
process to evaluate stochastic integrals:
As a first attempt, apply the form of the solution that mimics the
solution for the analogous problem in ordinary calculus.
2. Invoke Itô's Lemma to find the differential of the attempted solution.
3. Integrate both sides of the differential and rearrange the result to
solve for the desired stochastic integral.
1.
Illustration: Applying Itô's Lemma
Suppose that we seek to evaluate
above, we evaluate
𝑇
𝑍 𝑑𝑍𝑑
0 𝑑
𝑇
𝑍 𝑑𝑍𝑑 .
0 𝑑
Using the 3-step technique described
as follows:
Step1:
Attempt the ordinary calculus solution which would suggest that
YT =
𝑇
𝑍 𝑑𝑍𝑑
0 𝑑
1
2
= 𝑍𝑇2
Illustration: Applying Itô's Lemma (Continued)
Step2:
1
Find the differential of our attempted solution π‘Œπ‘‡ = 𝑍𝑇2 using Itô's Lemma. Choose
2
1 2
1 2
𝑦(π‘₯, 𝑑) = x so that Yt = 𝑦(𝑍𝑑, 𝑑) = 𝑍𝑑 . With dXt = dZt = 0·dt+1·dZt, and invoking Itô's
2
2
Lemma, we have:
πœ•π‘¦ πœ•π‘¦ 1 2 πœ• 2 𝑦
πœ•π‘¦
πœ•π‘¦
πœ•π‘¦
1 πœ•2𝑦
π‘‘π‘Œπ‘‘ = 𝑑𝑦 𝑍𝑑 , 𝑑 = 0 βˆ™
+
+ βˆ™ 1 βˆ™ 2 𝑑𝑑 + 1 βˆ™
𝑑𝑍𝑑 =
𝑑𝑍𝑑 +
𝑑𝑑 +
𝑑𝑑
2
πœ•π‘₯ πœ•π‘‘ 2
πœ•π‘₯
πœ•π‘₯
πœ•π‘₯
πœ•π‘‘
2 πœ•π‘₯
πœ•π‘¦
In Itô's Lemma, we must evaluate the partial derivatives at (x,t) = (Zt,t). Since =
πœ•π‘₯
πœ•π‘¦
πœ•2𝑦
π‘₯|π‘₯=𝑍𝑑 = Zt, = 0 and 2 = 1, we have:
πœ•π‘‘
πœ•π‘₯
1
π‘‘π‘Œπ‘‘ = 𝑍𝑑 𝑑𝑍𝑑 + 𝑑𝑑
2
Illustration: Applying Itô's Lemma (Continued)
Step3 :
We will integrate both sides of this equation for 0 ≤ t ≤ T. The left side of the following
equation is based on our discussion concerning the integral of a stochastic differential
at the start of Section 6.1.3. The right side of the following is based on the equation
immediately above:
𝑇
𝑇
𝑇1
π‘‘π‘Œ
=
𝑍
𝑑𝑍
+
𝑑𝑑
𝑑
𝑑
0
0 𝑑
0 2
𝑇
𝑇1
𝑍0 2 = 0 𝑍𝑑 𝑑𝑍𝑑 + 0 𝑑𝑑
2
𝑇
1
2
𝑍𝑇 = 0 𝑍𝑑 𝑑𝑍𝑑 + 𝑇
2
𝑇
π‘Œπ‘‡ − π‘Œ0 =
1
2
𝑍𝑇
2
−
1
2
1
2
Solving for the desired integral
0
1
𝑍𝑑 𝑑𝑍𝑑 = 𝑍𝑇
2
2
1
− 𝑇
2
Application: Geometric Brownian Motion
Geometric Brownian motion is an essential model for characterizing the
stochastic process for a security with value St at time t:
𝑑𝑆𝑑 = πœ‡π‘†π‘‘ 𝑑𝑑 + πœŽπ‘†π‘‘ 𝑑𝑍𝑑
Recall that μ and σ are the geometric mean security return and the standard
deviation of the security return per unit of time. We wish to determine the
value of the security St as a function of time. First, rewrite the differential above
in the form:
𝑑𝑆𝑑
= πœ‡π‘‘π‘‘ + πœŽπ‘‘π‘π‘‘
𝑆𝑑
Application: Geometric Brownian Motion (Continued)
Step 1
To evaluate the integral as though St is a real-valued function:
𝑇
T
𝑑𝑆𝑑
= ln 𝑆𝑑
= ln 𝑆𝑇 − ln 𝑆0
𝑆𝑑
0
0
Application: Geometric Brownian Motion (Continued)
Step 2
To take the differential of lnSt using Itô’s Lemma applying it to the
function y = y(St,t) = lnSt. In this formulation, dy from Itô's formula is
d(lnSt), a is µSt and b is St. We substitute these values into Itô's formula
as follows:
2

ο‚Άy ο‚Άy 1 2 2 ο‚Ά y οƒΉ
ο‚Άy
d (lnS t ) ο€½ οƒͺS t
   St
dt


S
dZ t
t
2οƒΊ
ο‚ΆS ο‚Άt 2
ο‚ΆS
ο‚ΆS 


1
1 2 2 1 οƒΉ
1
d (ln S t ) ο€½ οƒͺ S t
 0 ο€­  St
dt  S t
dZ t
2 οƒΊ
St
2
St
S t 
οƒͺ
1

οƒΉ
d ln( S t ) ο€½ οƒͺ  ο€­  2 οƒΊ dt  dZ t
2


Application: Geometric Brownian Motion (Continued)
Step 3
By integrating both sides of this equation from t = 0 to t = T, which results in:
𝑇
𝑇
𝑑 ln 𝑆𝑑 =
0
0
1 2
πœ‡ − 𝜎 𝑑𝑑 + πœŽπ‘‘π‘π‘‘
2
𝑇
𝑇
1 2
ln(𝑆𝑑 ) = πœ‡ − 𝜎 𝑑 + πœŽπ‘π‘‘
2
0
0
𝑆𝑇
1 2
𝑙𝑛𝑆𝑇 − 𝑙𝑛𝑆0 = 𝑙𝑛 = πœ‡ − 𝜎 𝑇 + πœŽπ‘π‘‡
𝑆0
2
Exponentiating both sides of this equation and multiplying by S0 yields:
𝑆𝑇 = 𝑆0 𝑒
1
2
πœ‡− 𝜎 2 𝑇+πœŽπ‘π‘‡
Returns and Price Relatives
The constant α = μ - σ2/2 is known as the mean logarithmic return of the
security per unit time. If we express the security price in the form:
𝑆𝑇 = 𝑆0 𝑒 𝛼𝑇+πœŽπ‘π‘‡
then ln(ST/S0)=αT+σZT is known as the log of price relative or the
logarithmic return. It is also useful to calculate the variance of the
logarithmic return.
𝑆𝑇
𝑆𝑇
π‘‰π‘Žπ‘Ÿ 𝑙𝑛
= 𝐸 (𝑙𝑛 − 𝛼𝑇)2 = 𝐸 𝜎 2 𝑍𝑇2 = 𝜎 2 𝐸 𝑍𝑇2 = 𝜎 2 𝑇
𝑆0
𝑆0
Returns and Price Relatives (Continued)
In comparison, let’s examine the expected value and variance of the arithmetic
return on 𝑆𝑇 over time T. The arithmetic return over time T is defined as r
=𝑆𝑇 /S0 – 1. To derive expected arithmetic return over time T, we first observe
that E[αT + σZT]= αT and Var[αT + σZT] = E[𝜎 2 𝑍𝑇2 ] = 𝜎 2 𝑇. Since αT + σZT ~
N(αT, 𝜎 2 𝑇) , by equation (2.26) in section 2.6.4, we have:
𝐸 π‘Ÿ =𝐸
𝑒 𝛼𝑇+πœŽπ‘π‘‡
−1 =𝑒
1
𝛼𝑇+2𝜎 2 𝑇
The variance of the arithmetic return is:
π‘‰π‘Žπ‘Ÿ π‘Ÿ =
2𝑇
𝜎
𝑒
−1
2𝑇
2𝛼𝑇+𝜎
𝑒
−1
Itô’s Formula: Numerical Illustration
Suppose, for example, that the following Itô process describes the price
path St of a given stock:
dS t ο€½ .01S t dt  .015S t dZ t
The differential for this process describes the infinitesimal change in the
price of the stock. The expected rate of return per unit time is .01 and the
standard deviation of the return per unit time is .015. The solution for
this equation giving the actual price level at a point in time is given by:
ST ο€½ S 0 e
2

.
015
οƒͺ  .01 ο€­
2
οƒͺ 
οƒΉ
οƒΆ
οƒ·T .015Z t οƒΊ
οƒ·

οƒΈ
Itô’s Formula: Numerical Illustration (Continued)
ο‚— The expected value and variance of the log of price relative are given by:
 ST οƒΉ

1 2
.015 2 οƒΆ
οƒ·οƒ· οƒ— 1 ο€½ .0098875
E οƒͺln
οƒΊ ο€½ T ο€½ T ο€­  T ο€½  .01 ο€­
2
2 οƒΈ

 S0 
 ST οƒΉ
2
2
Var οƒͺln
ο€½

T
ο€½
.
015
οƒ— 1 ο€½ .000225
οƒΊ
S0 

ο‚— The expected value and variance of the arithmetic return over a single period are
given by:
𝐸 π‘Ÿ =𝑒
π‘‰π‘Žπ‘Ÿ π‘Ÿ = 𝑒 𝜎
2𝑇
1
𝛼𝑇+2𝜎2 𝑇
− 1 𝑒 2𝛼𝑇+𝜎
2𝑇
−1 = 𝑒
1
.0098875×1+ .0152 ×1
2
2 ×1
= 𝑒 .015
−1 𝑒
− 1 = .01005
2×.0098875×1+.0152 ×1
= .00022957
Application: Forward Contracts
Let St be a stock price that follows a geometric Brownian motion:
𝑑𝑆𝑑 = πœ‡π‘†π‘‘ 𝑑𝑑 + πœŽπ‘†π‘‘ 𝑑𝑍𝑑
Let Ft,T be the time t price of a forward contract that trades on that stock,
settling at time T:
𝐹𝑑,𝑇 = 𝑆𝑑 𝑒 π‘Ÿ(𝑇−𝑑)
Recall Itô’s Formula:
ο‚Άy ο‚Άy 1 2 ο‚Ά 2 y
ο‚Άy
dy ο€½ [a

 b
]dt  b
dZ t
2
ο‚Άx ο‚Άt
2
ο‚Άx
ο‚Άx
Application: Forward Contracts (Continued)
Here, we choose y(x,t) = xer(T-t) so that Ft,T = y(St, t). In this case a =μSt
and b = σSt. Note that
πœ•π‘¦
πœ•π‘₯
=
πœ•π‘¦
π‘Ÿ(𝑇−𝑑)
𝑒
,
πœ•π‘‘
= −π‘Ÿπ‘†π‘‘
𝑒 π‘Ÿ(𝑇−𝑑) ,
and
πœ•2 𝑦
πœ•π‘₯ 2
= 0 evaluated at (x,t) = (St,t).
Applying Itô’s Formula gives:
𝑑𝐹𝑑,𝑇 = πœ‡π‘†π‘‘ 𝑒 π‘Ÿ 𝑇−𝑑 − π‘Ÿπ‘†π‘‘ 𝑒 π‘Ÿ
𝑇−𝑑
𝑑𝑑 + πœŽπ‘†π‘‘ 𝑒 π‘Ÿ
𝑇−𝑑
𝑑𝑍𝑑
Since Ft,T = Ster(T-t), the price of a forward contract on a single equity also
follows a geometric Brownian motion process as follows
dFt ,T ο€½  ο€­ r Ft ,T dt  Ft ,T dZ t
Download