An Introduction To Stochastic Volatility Models
An Introduction To Stochastic Volatility Models
GRAU DE MATEMÀTIQUES
AN INTRODUCTION TO
STOCHASTIC VOLATILITY
MODELS
Introduction i
2 Stochastic Integration 7
2.1 Martingales and Brownian motion . . . . . . . . . . . . . . . . . 7
2.2 Integral of mean square integrable processes . . . . . . . . . . . 9
2.3 Extension of the integral . . . . . . . . . . . . . . . . . . . . . . . 12
3 Fundamental Theorems 17
3.1 Itô’s formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Stochastic differential equations . . . . . . . . . . . . . . . . . . . 18
3.3 Girsanov’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Martingale representation theorem . . . . . . . . . . . . . . . . . 21
5 Stochastic Volatility 39
5.1 Empirical motivations . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2 A general approach for pricing . . . . . . . . . . . . . . . . . . . 39
5.3 Heston’s model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6 Conclusions 47
7 Bibliography 49
Bibliography: Articles . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Bibliography: Books . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Bibliography: Online . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
i
Introduction 1
The main goal of this work is to introduce the stochastic volatility models in
mathematical finance and to develop a closed-form solution to option pricing
in Heston’s stochastic volatiltiy model, following the arguments in Heston
1993.
Louis Bachelier, in his thesis "Théorie de la Spéculation", made the first contri-
bution of advanced mathematics to the study of finance in 1900. This thesis
was well received by academics, including his supervisor Henry Poincaré,
and was published in the prestigious journal Annales Scientifiques de l’École
Normale Supérieure. In this pioneering work, the Brownian motion is used for
the modelling of movements in stock prices. In the words of Louis Bachelier,
in Bachelier 1900:
Bachelier argued that, over a short time period, fluctuations in price are inde-
pendent of the current price and the past values of the price, and that these
fluctuations follow a zero mean normal distribution with variance propor-
tional to the time difference. He also assumed that the prices are continuous,
therefore modelled as a Brownian motion (see Bachelier 2011).
Many years later, in the famous article by Black and Scholes, Black and Sc-
holes 1973, prices are modeled as a geometric Brownian motion, whose fluctua-
tions have a lognormal distribution. This model is based on the assumption
that the log returns of a stock price are independent and normally distributed,
with variance proportional to the time difference. The log returns are defined
as log( pi ) log( p j ), where pi and p j denote the prices at times i and j, respec-
tively, with i > j.
pi pj
ri =
pj
Hence, the approxamation
3
4 Historical Background on Stock Price Models
pi pj
= ri ⇡ log(1 + ri ) = log( pi /p j ) = log( pi ) log( p j ) (1.2)
pj
So that the lognormal distribution for the stock price increments proposed by
Black and Scholes obeys to the intuitive idea that the price returns are inde-
pendent and normaly distributed.
Explicitely, as stated in Black and Scholes 1973, the model proposed by Black
and Scholes relies on the following assumptions of an "ideal" market:
(b) The distribution of stock prices at the end of any finite interval is lognor-
mal
Stochastic Integration
Definition 2.1 (Stochastic Process). Let E be a metric space with the borel s-field.
A stochastic process is a family of E-valued random variables ( Xt )t2T defined in a
probability space (W, F , P )
7
8 Stochastic Integration
The following result will be useful when defining the stochastic integral in
the next section:
Proof. Refer to Lamberton and Lapeyre 2007 (page 58, exercise 13).
The following result shows further properties of the Brownian motion which
are not explicit in the definition. For a proof of this result, refer to Corcuera
n.d.
E [ Xt |Fs ] Xs = E [ Xt Xs |Fs ] = E [ Xt Xs ] = 0
Lemma 2.8. Let (Wt )t 0 be a Brownian motion under the probability P. For any
fixed M > 0,
P (max Wt (w ) M ) = 0
t 0
P (min Wt (w ) M) = 0
t 0
TM (w ) = inf{t 0, Wt (w ) = M }
It can be proved that the first passage time distribution has the following
property (see Karatzas and Shreve 2012, p.80, equation 6.2):
We aim to define the integral for mean square integrable processes. By iden-
tifying processes that are versions of one another and defining the natural
RT
norm k X k2S2 = E ( 0 Xt2 dt), it can be shown that S2 is a Hilbert space, with
RT
the scalar product < X, Y >= E ( 0 Xt Yt dt).
n 1
I ( X )t = Â Xti (Wti+1 ^t Wti ^t )
i =0
Let E ⇢ S2 be the set of simple processes. Note that the integral of a simple
process is a random variable. In addition, it is in L2 , and the integral is an
isometry between these two spaces:
Lemma 2.12 (Isometry Property for Simple Processes). The integral defines an
isometry between E and a subspace of L2 (W, F T , P ).
n 1
I ( X )2 = Â Xt2i (Wti+1 Wti )2 + 2 Â Xti Xt j (Wti+1 Wti )(Wt j+1 Wt j )
i =0 0 i < j < n
n 1
E [ I ( X )2 ] = Â E[Xt2i ](ti+1 ti )
i =0
Since X is constant on every interval [ti , ti+1 [, the expression above is equal
to:
n 1 hZ i +1 i hZ T i
ÂE Xt2 dt = E Xt2 dt
i =0 i 0
The following result will allow us to extend the integral to the S2 space:
The proof of this result requires several steps and will be omitted - it can be
found in Karatzas and Shreve 2012 (page 134, problem 2.5). However, to give
an idea on how it is accomplished, note that any continuous process ( Xt )0tT
can be approximated in S2 ([0, T ]) by simple processes Xtn defined as:
T2n 1
Xtn = Â X (i2 n
)1]i/2n ,(i+1)/2n [ (t)
i =0
2.2 Integral of mean square integrable processes 11
However, it is not trivial to obtain a similar result for more general adapted
processes.
To extend the integral, we apply the general result that given two metric
spaces X and Y and a dense subset A of X, any uniformly continuous map
f : A ! Y can be extended in a unique way to X by defining fˆ( x ) = lim f ( an )
n
where an is any sequence in A converging to x, and this extension is well
defined and continuous. In addition, if Y is a vector space, the extension
preserves the supremum norm. In our case, we can extend the integral
I : E ! L2 (W, F T , P ) to Î : S2 ! L2 (W, F T , P ) and it remains an isome-
try.
RT
Given a process ( Xt )0tT , we define the process I ( X ) t
by I ( X )t = 0 Xs 1[0,t] (s) dWs .
This process is a continuous martingale:
Proof. Consider a first case in which X is a simple process, and let s < t. Then,
we can assume tn = t and s 2 [tk , tk+1 [.
Z t n
0
Xs dW = Â Xti (Wti+1 Wti )
i =0
k 1 n
= Â Xti (Wti+1 Wti ) + Xtk (Wtk+1 Wtk ) + Â Xti (Wti+1 Wti ) (2.2)
i =0 i = k +1
Now, the first part of the last equation is an Fs measurable random variable,
so its conditional expectation to Fs is the same variable.
For the second part, using the martingale property of the Brownian motion,
E[ Xtk (Wtk+1 Wtk )|Fs ] = Xtk E[Wtk+1 Wtk |Fs ] = Xtk (Ws Wtk )
Finally, for j > k, and by the tower property of the conditional expectation,
k 1
E[ I ( X )t |Fs ] = Â Xti (Wti+1 Wti ) + Xtk (Ws Wtk ) = I ( X )s
i =0
We have proven that the integral of a simple process is a martingale. To
extend this result to the S2 space, it is sufficient to show that the martingale
property is preserved by limits in L2 . This is a consequence of the fact that
the conditional expectation is continuous in L2 . To prove this, recall that for
X, Y in L2 :
" # " #
⇣ ⌘2 ⇣ ⌘2
E E [ X |Ft ] E [Y |Ft ] =E E[X Y |Ft ]
h i h i
E E [( X Y )2 |Ft ] = E ( X Y )2
Where the inequality is a consequence of Jensen’s inequality.
12 Stochastic Integration
The proof is based on the proof in Lamberton and Lapeyre 2007 (page 38,
Proposition 3.4.4).
hZ T i
n+ p 2
= 4E Xs Xsn ds !0
0 n!•
3. lim tn = T, a.e.
n!•
Rt
Lemma 2.16. Given a process X 2 S , 0 Xs2 ds is Ft -measurable.
2.3 Extension of the integral 13
Z s Z t
2
{tn > t} = {w |8s t, | Xu (w )| du < n} = {w | | Xu (w )|2 du < n}
0 0
Rs
Because 0 | Xu (w )|2 du is an increasing function of s. So tn is a stopping time.
In order to prove that the integral is well defined, it is necessary to show that
the limit exists, almost surely. We state this result in the following theorem:
Theorem 2.17. For any process X 2 S , the integral process of X as defined in 2.3
exists almost surely.
The proof is based on the one by Lamberton and Lapeyre 2007. We will need
the following proposition:
Z T Z T n n Z T
0
1{s>t } Hs dWs =
0
 1 A i 1{ s > t i } Hs dWs =  1 Ai 1{s>ti } Hs dWs
i =1 i =1 0
14 Stochastic Integration
n Z T Z T
= Â 1 Ai ti
Hs dWs =
t
Hs dWs
i =1
It follows that:
Z t Z T
Hs dWs = 1{st } Hs dWs
0 0
Now, consider an arbitrary stopping time t and define a decreasing sequence
tn by:
2n 1
( k + 1) T
tn = Â 2n
1{ kT t (k+1)T }
2n 2n
k =0
Rt
Clearly, tn converges to t almost surely. Since the map t 7! 0 Hs dWs is
Rt Rt
almost surely continuous, 0 n Hs dWs converges to 0 Hs dWs almost surely.
Rt RT
By the previous discussion, we know that 0 n Hs dWs = 0 1{stn } Hs dWs for
all n 1. Now,
Z T Z T
! Z T
!
2
E 1{stn } Hs dWs 1{st } Hs dWs =E 1{t <stn } Hs2 ds
0 0 0
I ( X n )t = I ( X m )t
for all m n (2.4)
RT S RT
Since {w, 0 Xs (w )2 ds < •} = {w, 0 Xs (w )2 ds < n} and the set has
n 2N
probability 1, the sequence I ( X n )t converges almost surely.
By construction, and by 2.4, the extended integral has almost surely continu-
ous sample paths. Note that the extended integral does not necessarily have
the martingale property. However, we will show that it is a local martingale.
Proof. The proof is based on the one by Capinski, Kopp, and Traple 2012
(p.140, Proposition 4.25).
The proof is based on the one in Musiela and Rutkowski 2006 (p. 591, Propo-
sition A.7.1). We will need the following result (see Musiela and Rutkowski
2006, p.580, Lemma A.1.2 for a proof):
Proof. (Proposition)
Since M is a local martingale, there exists an increasing sequence of stopping
times (tn )n2N such that P (lim tn = T ) = 1 and such that for each n, the
n
stopped process Mttn is a martingale.
Let 0 s t T. Because M is non-negative, and by the conditional form
of Fatou’s lemma,
Fundamental Theorems
The standard results that follow are basic for the development of the mathe-
matical finance theory.
Definition 3.1 (Itô Process). An Itô process is a process that satisfies the equation:
Z t Z t
X t = X0 + as ds + b s dWs
0 0
Where:
1. X0 is F0 -measurable.
Equivalently, we say that the process satisfies the equation (in differential notation):
dXt = at dt + b t dWt
Proof. Refer to Capinski, Kopp, and Traple 2012 (p. 98, Theorem 3.27).
Theorem 3.3 (Itô’s formula). Let g 2 C1,2 ([0, t] ⇥ R, R ) and ( Xt ) be an Itô pro-
cess, with:
dXt = at dt + b t dWt
Then Yt = g(t, Xt ) is also an Itô process, and satisfies the formula:
1
dYt = gt + gx at + gxx b2t dt + gx b t dWt (3.1)
2
Where all partial derivatives of g are evaluated at (t, Xt ).
For a proof of Itô’s formula, refer to Capinski, Kopp, and Traple 2012 (p. 136,
section 4.7). The following example will be particularly relevant:
17
18 Fundamental Theorems
Example 3.4. Let St be an Itô process such that dSt = µSt dt + sSt dWt , and let
g(t, x ) = log(x). Then, if Yt = g(t, St ),
1 2
dYt = (µ s )dt + sdWt
2
Or, equivalently:
1 2
St = exp (µ s )t + sWt
2
We will also require a two-dimensional version of Itô’s formula, adapted from
Musiela and Rutkowski 2006.
Theorem 3.5 (Itô’s formula, two dimensions). Let (Wt1 ), (Wt2 ) be Brownian
motions, g 2 C1,2 ([0, t] ⇥ R2 , R ) and ( Xt ), (Yt ) be Itô processes, with:
dXt = at dt + bt dWt1
dYt = at dt + b t dWt2
Then Zt = g(t, Xt , Yt ) is also an Itô process, and satisfies the formula:
1 1
dZt = gt + gx at + gy at + gxx bt2 + gyy b2t + gxy bt b t dt (3.2)
2 2
+ gx bt dWt1 + gy b t dWt2
Where all partial derivatives of g are evaluated at (t, Xt , Yt ).
Proof. Refer to Capinski, Kopp, and Traple 2012 (p.160, Theorem 5.8).
3.3 Girsanov’s theorem 19
The natural extension of this result to multiple dimensions is also valid. That
is, consider a multi-dimensional version of 3.3,
In our case, (Ft ) is the filtration generated by the Brownian motion (Wt ).
The following result guarantees that under certain conditions on the coeffi-
cient functions, this property is satisfied by solutions of stochastic differential
equations. It has been adapted from Capinski, Kopp, and Traple 2012 (p.174,
Theorem 5.14).
E [ f ( Xt )|FWs ] = E [ f ( Xt )|F Xs ]
Proof. Refer to Capinski, Kopp, and Traple 2012 (p.174, Theorem 5.14).
Theorem 3.9 (Girsanov, simple version). Let (Wt ) be a standard Brownian Mo-
tion and g 2 R. Then, W̃t := Wt + gt is a standard Brownian motion under the
probability P̃ defined by:
dP̃ 1 2
= exp gWT g T
dP 2
Proof. The proof has been adapted from Capinski and Kopp 2012.
It is clear that W̃ has continuous sample paths. We will prove directly that
W̃t has independent and normally distributed increments, by computing the
20 Fundamental Theorems
⇥ ⇤ h 1 2 i
P̃ ( A) = EP̃ 1 A = EP exp gWT g T 1A
2
Now, WT can be expressed as Âin=1 Wti Wti 1 . Similarly, T = Âin=1 (ti ti 1)
n
and 1 A = ’ 1 Ai . We obtain:
i =1
" #
n
1 2
P̃ ( A) = EP ’ exp g(Wti Wti 1 )
2
g ( ti ti 1) 1 Ai
i =1
Ai = {Wti Wti 1
+ g ( ti ti 1) ai }
So, because of the independence property,
n h 1 2 i n
P̃ ( A) = ’ P exp
E g(Wti Wti 1 )
2
g ( ti ti 1 ) 1 Ai = ’ P̃( Ai ) (3.7)
i =1 i =1
h 1 2 i
P ( Ai ) = EP exp g(Wti Wti 1 ) g ( ti ti 1 ) 1 Ai
2
Z
1 2 1 x2
= exp gx g ( ti ti 1 ) p exp dx
{ x + g ( ti ti 1 ) ai } 2 2p (ti ti 1) 2( t i t i 1)
Z ⇣ 2⌘
1 x g ( ti ti 1 )
= p exp dx
{ x + g ( ti ti 1 ) ai } 2p (ti ti 1) 2( t i t i 1 )
Z ⇣ ⌘
1 z2
= p exp dz (3.8)
{ z ai } 2p (ti ti 1) 2( t i t i 1)
Equation 3.8 proves that W̃ti W̃ti 1 are distributed as N (0, ti ti 1 ), and
equation 3.7 proves that they are independent.
Rt
Define W̃t := Wt + 0 gs ds, and define the probability P̃ by:
⇣ Z T Z T ⌘
dP̃ 1
= exp gs dWs gs2 ds (3.10)
dP 0 2 0
Lemma 3.12. Let (Gn ) be a filtration in a probability space (W, F , P ) and let X 2
L2 = L2 (W, F , P ). Then,
L2
E [ X |Gn ] ! E [ X |G• ]
n!•
Where G• = s(Gn , n 2 N ).
sup k Xn k L2 k X k L2 < •
n 2N
By theorem 3.11,
L2
Xn !Y
n!•
E [Y |Gn ] = E [ X• |Gn ]
Hence, for every G 2 [n2N Gn ,
E [Y1G ] = E [ X• 1G ] (3.11)
Define the collection
C = { G 2 G• | E [Y1G ] = E [ X• 1G ]} (3.12)
Clearly, by 3.11,
[n2N Gn ⇢ C ⇢ G• (3.13)
We wish to show that C = G• .
22 Fundamental Theorems
Now, if we prove that C is a s-algebra, we will have, by equation 3.13 and the
fact that [n Gn generates G• , that C = G• .
Now, Bn := { X• Y > n1 } 2 G• , so
E [ X• 1Bn ] = E [Y1Bn ]
This implies that P ( Bn ) = 0 8n 2 N, so that P ([n Bn ) = 0. Hence X• Y a.s,
and a similar argument shows that X• Y a.s.
Lemma 3.13. Let (Wt ) be a Brownian motion in the filtered probability space (W, F , F, P ),
where F is generated by (Wt ). Consider the set J of stepwise functions f : [0, T ] !
R of the form:
n
f = Â l i 1] t i 1 ,ti ]
i =0
With li 2 R and 0 = t0 < . . . < tn = T. For each f 2 J , define
nZ T 1
Z T o
f
ET = exp f (s)dWs f 2 (s)ds
0 2 0
f
Let Y 2 L2 (F T , P ), and assume that Y is orthogonal to E T for all f 2 J . Then,
Y = 0.
f
Proof. Let f 2 J and Y 2 L2 (F T , P ) orthogonal to E T . Define Gn := s(Wt0 , . . . , Wtn ).
By assumption,
!
n n o
E exp  li Wti Wti 1
Y =0
i =1
! !
n n o n n o
E exp  li Wti Wti 1
E [Y+ |Gn ] = E exp  li Wti Wti 1
E [Y |Gn ]
i =1 i =1
(3.14)
3.4 Martingale representation theorem 23
If, on the contrary, I ( H, then there would exist a centered, non-trivial ran-
dom variable Z 2 H orthogonal to I . We will prove that this is not possible.
f f
Indeed, suppose that such Z exists. Take Yt := Et , with Et as defined in
Lemma 3.13. Then,
h Z T i
f
E Z· Et dWt =0
0
And also
h Z T i
f
E Z· 1+ Et dWt =0 (3.19)
0
f
Now, a similar argument as the one in 3.4 shows that Et is the solution to the
following stochastic differential equation:
f f
dEt = f (t)Et dWt
24 Fundamental Theorems
In particular,
Z T
f f
ET = 1+ f (t)Et dWt (3.20)
0
From 3.19 and 3.20 we conclude:
f
E[ZET ] = 0 (3.21)
Since this is true for any f defined as in 3.13, by 3.13 we conclude that Z =
0.
The Black-Scholes model consists of two stocks S and b, in a time frame [0, T ],
where b is a bank account such that b(t) = ert , and the risky stock S satisfies
the differential equation:
a(t, x ) = µx (4.2)
b(t, x ) = sx
These functions clearly satisfy the regularity conditions in 3.6, so there exists
a unique mean-square integrable solution St with continuous sample paths.
By theorem 3.8, this solution also satisfies the Markov property.
An explicit solution for this stochastic differential equation was found in 3.4:
1 2
St = S0 exp{ (µ s )t + sWt } (4.3)
2
The following are some implicit assumptions of this model, as argued in Cor-
cuera n.d.:
St Su St 1 2
= 1 = exp{ (µ s )(t u) + s(Wt Wu )} 1
Su Su 2
St Su St u S0
⇠
Su S0
25
26 The Black-Scholes Model
Lemma 4.1. Let f be a strategy satisfying integrability conditions (1) and (2). Then,
f 2 F iff
Proof. The proof has been adapted from Lamberton and Lapeyre 2007 (p. 65,
Proposition 4.1.2).
rt
dṼf (t) = rṼf (t)dt + e dVf (t)
dṼf (t) = re rt
(ft1 ert + ft2 St )dt + e rt
(ft1 db(t) + ft2 dSt )
= ft2 ( re rt
St dt + e rt
dSt ) = ft2 dS̃t
1 2
Lemma 4.2. Given a Brownian motion (Wt ), the process Mt := S0⇤ exp{ 2s t +
sWt } is a martingale.
4.1 Risk-neutral measure 27
h i h 1 2 i
E Mt /Ms |Fs = E exp{ s (t s) + s (W̃t W̃s )}|Fs
2
h 1 i
= E exp{ s2 (t s) + s (W̃t W̃s )} (because W̃t W̃s ?Fs )
2
Now,
h 1 2 i
E exp{ s (t s) + s(W̃t W̃s )}
2
Z n
1 2 1 x2 o
= exp s (t s) + sx p exp dx
R 2 2p (t s) 2( t s )
Z n
1 (x s (t s))2 o
= p exp dx = 1
R 2p (t s) 2( t s )
⇣ dQ ⌘ ⇣ µ r 1 (r µ )2 ⌘
= exp WT⇤ T (4.5)
dP t s 2 s2
µ r
Where the process Wt⇤ := Wt + s t is a Brownian motion under P.
1 2 1 2
St⇤ = S0⇤ exp{(r s )t + sWt⇤ } = exp{rt}S0⇤ exp{ s t + sWt⇤ }
2 2
1 2
The previous lemma implies that exp{ 2s t + sWt⇤ } is a martingale under
Q, which proves the result.
Proof. Refer to Capinski and Kopp 2012 (p. 51, Theorem 3.12)
28 The Black-Scholes Model
The following example shows that the market model described above, with
the set of self-financing strategies, has arbitrage opportunities. In order to
eliminate these opportunities, it will be necessary to restrict the set of so-
called admissible strategies, introduced below.
Theorem 4.6. Arbitrage opportunities exist within the class of self-financing strate-
gies.
This example is based on the suicide strategy described in Capinski and Kopp
2012 (p.24-27).
1
ft2 = p
sS̃t T t
The risk free component is determined by the self-financing condition and
an (arbitrary) initial value. Now, the strategy is not almost surely square
integrable, but it will be modified later on. Firstly, we will study its properties.
Proof.
1
dṼf (t) = ft2 dS̃t = ft2 sS̃t dWt⇤ = p dWt⇤
T t
So that:
Z t
1
Ṽf (t) Ṽf (0) = p dWu⇤
0 T u
Now, let
Z t
1
g(t) = p du
0 T u
Then, Ṽf (t) Ṽf (0) and W ⇤ (g(t)) have the same distribution. A proof of this
result can be found at Capinski and Kopp 2012 - here we will focus on its
consequences. Recall that - see Lemmma 1.8, for any fixed M > 0:
Q (max Wt⇤ M) = 0
t 0
Q (min Wt⇤ M) = 0
t 0
P (max Wt⇤ M) = 0
t 0
P (min Wt⇤ M) = 0
t 0
Proof. (Theorem)
c1t = qt1 + 1
c2t = qt2
We will now define which strategies are considered admissible in the model.
Although several alternatives have been proposed in the literature, here we
follow the definition from:
The next result guarantees that no arbitrage opportunities exist within the
admissible strategies:
30 The Black-Scholes Model
Theorem 4.12. Let H be the payoff of a derivative which is replicated by the strategy
f. Assuming that the option price is an Itô process, the No Arbitrage Principle
implies that the price of the option at time t, Vt , is equal to Vf (t) for all t 2 T.
Proof. This proof is based on the one in Capinski and Kopp 2012 (p.21, Theo-
rem 2.16). Assume that this is not the case, and let t0 be any time in which a
difference between Vf (t) and Vt appears with positive probability. Consider
a strategy y with zero initial value and that buys the cheaper of the two and
sells the most expensive short at time t0 and invests the remaining money in
the bank account. The value Vy ( T ) of this strategy is positive with a positive
probability. Indeed, assuming without a loss of generality that Vf (t0 ) > Vt0 ,
4.3 Completeness
We aim to prove that any square integrable European option can be replicated
by an admissible strategy. The proof of the following theorem is based on the
proof in Capinski and Kopp 2012:
Proof. We will see that, in fact, such a f exists that satisfies the additional
property that Vf⇤ (t) is a martingale under Q, and
h i
Vf (t) = b(t)Vf⇤ (t) = b(t)EQ H ⇤ (t)|Ft (4.9)
dVf (t) = ft1 db(t) + ft2 dSt = ft1 rb(t)dt + ft2 (rSt dt + sSt dWt⇤ ) (Self-financing condition)
⇣ h i⌘ ⇣ h i⌘ h i
d b(t)EQ H ⇤ (t)|Ft = b(t)d EQ H ⇤ (t)|Ft + EQ H ⇤ (t)|Ft db(t) (Itô Formula)
h i
= b(t) X (t)dWt⇤ + rb(t)EQ H ⇤ (t)|Ft dt
" # " #
h i
0 = rb(t)ft1 + rft2 St rb(t)EQ H ⇤ (t)|Ft dt + ft2 sSt b(t) X (t) dWt⇤
b(t) X (t)
ft2 = (4.12)
sS(t)
And substituting in the first equality gives:
h i
ft1 = EQ H ⇤ (t)|Ft X (t)/s
We have obtained a unique strategy that may attain H - now we need to verify
all the conditions in the theorem. Firstly, the following calculation shows that
f attains H:
h i h i
Vf (t) = ft ⇤ (b(t), St ) = b(t))EQ H ⇤ (t)|Ft b(t) X (t)/s + b(t) X (t)/s = b(t)EQ H ⇤ (t)|Ft
In particular, Vf ( T ) = H.
h i
The strategy is admissible because Vf⇤ (t) = EQ H ⇤ |Ft is clearly a martingale.
(4.13)
Note that H should be square-integrable with respect to Q, due to the condi-
tions in Theorem 4.13. In that case, the theorem guarantees the existence of a
replicating strategy f, validating the previous argument. Now, if H = h(ST ),
with h : R ! R being a bounded, measurable function, we can express Vt
as a function of St and t - the condition that h is bounded will be needed to
apply the Markov property. Indeed, following the arguments in Lamberton
and Lapeyre 2007 (p. 69, Remark 4.3.3),
" #
⇣ ⌘
r ( T t) r ( T t) s2 /2)( T t)+s(WT⇤ Wt⇤ )
Vt = EQ [e H |Ft ] = EQ e h St e (r Ft
" #
⇣ ⌘
r ( T t) s2 /2)( T t)+s(WT⇤ Wt⇤ )
= EQ e h St e (r St (because of the Markov property)
4.4 Pricing and hedging 33
" #
⇣ ⌘
r ( T t) s2 /2)( T t)+s(WT⇤ Wt⇤ )
= EQ e h xe(r St = x
x = St
" #
⇣ ⌘
r ( T t) (r s2 /2)( T t)+s(WT⇤ Wt⇤ )
= EQ e h xe
x = St
Where the last equality is a consequence of the fact that Wt⇤ is inde- if WT⇤
pendent of Ft , then it is independent of St . Now, we can express Vt as:
Vt = P(t, St ) (4.14)
Where
" #
⇣ ⌘
r ( T t) s2 /2)( T t)+s(WT⇤ Wt⇤ )
P(t, x ) = EQ e h xe(r (4.15)
And, since WT⇤ Wt⇤ is distributed as N (0, T t) under Q, the previous ex-
pectation is expressed by the following integral:
Z • ⇣ p ⌘ y2 /2
r ( T t) (r s2 /2)( T t)+sy T t ep
P(t, x ) = e h xe dy (4.16)
• 2p
In the case of calls and puts, P(t, x ) can be calculated explicitely, giving rise
to the Black-Scholes pricing formulas. Firstly, we will need to prove that call
and put options are square integrable with respect to Q. The case of a put
option with payoff H = h(ST ) = (K ST )+ , this is clear since the payoff is
bounded by K. We admit the following result, which shows that this is also
the case for a call option, with payoff H = h(ST ) = (ST K )+ .
Lemma 4.14. The payoff H = h(ST ) = (St K )+ of a call option is square inte-
grable with respect to Q.
WT⇤ Wt⇤
Let q = T t and z = p , a standard normal random variable under Q.
q
Then, " #
⇣ p ⌘+
rq s2 q/2
P(t, x ) = E Ke xes qz
Define:
log(K/x ) (r s2 /2)q
d= p (4.17)
s q
p
Note that Ke rq xes qz s2 q/2 0 () z d. We obtain:
" #
⇣ p ⌘
rq s2 q/2
F (t, x ) = E Ke xes qz
1{ z d }
34 The Black-Scholes Model
Z d ⇣ p ⌘ y2 /2
rq s qy s2 q/2 e
= Ke xe p dy
• 2p
Z d y2 /2 Z d p y2 /2
rq e s qy s2 q/2 e
= Ke p dy xe p dy
• 2p • 2p
The value of the first integral is Ke rq times the cumulative distribution func-
tion of a standard normal random variable evaluated p at d: N (d). For the
second integral,
p the change of variable t = y + s q shows that its value is
xN (d s q ). Finally,
r ( T t)
p
P(t, x ) = Ke N (d) xN (d s T t) (4.18)
Theorem 4.15 (Call-put parity). Let Ct , Pt be the price of a call option and a put
option, respectively, at time t, both with strike price K and maturity T. Then,
r ( T t)
Ct Pt = St Ke (4.19)
4.4 Pricing and hedging 35
Proof. We follow closely the arguments and notations in Capinski and Kopp
2012 (p.56, Theorem 3.16).
Note that
ST (ST K )+ + (K ST )+ = K (4.20)
This equation can be deduced by separately considering the cases ST K 0
and ST K 0. Hence,
ST CT + PT = K (4.21)
Multiplying both sides by e rT ,
rT rT rT
ST⇤ e CT + e PT = Ke (4.22)
Now, (St⇤ ), (e rt Ct ), and (e rt Pt ) are Q-martingales (this assertion relies on
the fact that call and put options are replicable), so
rT rT rT rT rt rt
e = E [e |Ft ] = E [ST⇤ e CT + e PT |Ft ] = St⇤ e Ct + e Pt
(4.23)
The result follows by multiplying both sides by ert .
The price of a call option follows from 4.18 and the call-put parity. Indeed,
r ( T t)
Ct = Pt + St Ke (by 4.19)
r ( T t)
p r ( T t)
= Ke N (d) St N ( d s T t ) + St Ke (by 4.18)
p r ( T t)
= St (1 N (d s T t)) Ke (1 N (d))
p r ( T t)
= St N ( d + s T t) Ke N ( d) (4.24)
In particular, the price of a call option is a function of t and St :
Ct = C (t, St ) (4.25)
Where
p r ( T t)
C (t, x ) = xN ( d + s T t) Ke N ( d) (4.26)
4.4.3 Hedging
Now that the pricing formulas have been justified, we aim to obtain explicit
formulas for replicating (or hedging) strategies in the same setting of a Eu-
ropean option with an F T -measurable payoff H = h(ST ) which is square
integrable with respect to Q. We follow the arguments in Fouque, Papanico-
laou, and Sircar 2000, which begin by recalling that the value of the option is
equal to:
Vt = P(t, St ) (4.27)
The fact that Vt is a function of t and St was proven in 4.15 assuming that h
was a bounded function. However, we saw in 4.25 that this is also true for
36 The Black-Scholes Model
call options. Hence, for the following, we assume that either h is bounded or
h( x ) = ( x K )+ .
1
(rert ft1 + ft2 µSt )dt + ft2 sSt dWt = ( Pt + µSt Px + s2 St2 Pxx )dt + sSt Px dWt
2
(4.29)
Where all the partial derivatives of P are evaluated at (t, St ) - we will use
this abbreviation later on without further mention. By the uniqueness of the
expression of an Itô process, we conclude that:
ft1 = e rt
( P(t, St ) Px (t, St )St ) (4.31)
The previous argument also shows the relationship between the pricing func-
tion P(t, x ) and a certain PDE (the Black-Scholes PDE). Indeed, substituting
the value of the hedging strategy in 4.29, we obtain the formula:
1
= Pt + Px µSt + s2 St2 Pxx dt + sSt Px dWt
2
Equating the drift terms in both expressions, we obtain the equation:
1
Pt rP + rPx St + s2 St2 Pxx = 0
2
In particular, the drift terms in equation 4.32 will be equal if P satisfies the
following PDE:
1
Pt rP + rxPx + s2 x2 Pxx = 0, 8 t, x 2 R + (4.33)
2
P(t, x ) = h( x ), 8 t, x 2 R +
Where all the partial derivatives of P are evaluated at (t, x ).
38 The Black-Scholes Model
Chapter 5
Stochastic Volatility
St 1 2
= exp{ (µ s )(t u) + s(Wt Wu )}
Su 2
So that
St Su St 1 2
⇡ log( ) = (µ s )(t u) + s(Wt Wu )
Su Su 2
⇣ ⌘
Which is distributed as N µ 12 s2 )(t u), s (t u) . Hence, the variance
rate of the returns is approximately given by the volatility .
In a Stochastic Volatility framework, the stock price is modelled as:
39
40 Stochastic Volatility
Where (St ) represents the stock price, st = f (Yt ) for some positive function
f , and Yt is an Ornstein-Uhlenbeck process, defined by:
h i
E Ẑt E [ Ẑt ] Wt E [Wt ] E [ Ẑt Wt ] rt
cor ( Ẑt , Wt ) = = pp = =r (5.5)
sd( Ẑt )sd(Wt ) t t t
Thus, this model allows us not only to incorporate randomness into the
volatility, but also to specify the correlation between the volatility and the
stock price, or skewness, as argued in the previous section.
r ( T t)
Vt = EQ [e H |Ft ]
So, for each risk neutral measure Q, we are able to find a reasonable option
price Vt . We will now find a family of equivalent risk neutral measures, by
means of the multidimensional Girsanov theorem.
µ r
Indeed, let qt := st , and define:
Z t
Wt⇤ = Wt + qs ds
0
Z T Z T Z T
!
dQ g 1
= exp qs dWs gs dZs (qs2 + gs2 )ds (5.6)
dP 0 0 2 0
h q i
dYt = a(m Yt ) b t (rqt + gt 1 r2 dt + b t dẐt⇤ , (5.8)
⇣q ⌘
Ẑt⇤ = rWt⇤ + 1 r2 Zt⇤ (5.9)
" #
⇣ ⌘
r ( T t)
= EQg e h ST FSt ,Yt (Because of the Markov property)
" #
⇣ ⌘
r ( T t)
= EQg e h ST St = s, Yt = y
s=St ,y=Yt
rt rt rt
d( e P(t, St , Yt )) = re P(t, St , Yt )) + e dP(t, St , Yt )
And the two dimensional Itô formula gives:
rt
d e P(t, St , Yt ) = (5.10)
" q
h ⇣ ⌘i
rt
=e Pt rP + rSt Ps + a(m Yt ) b t rqt + gt 1 r2 Py
#
1 2 2 1 2
+ f (Yt ) St Pss + b t Pyy + r f (Yt )St b t Psy dt
2 2
⇣ ⌘ ⇣ ⌘
+e rt f (Yt )St Ps dWt⇤ + e rt ny b t Py dẐt⇤
h ⇣ µ r q ⌘i
Pt rP + rsPs + a(m y) b(t, s, y) r + g(t, s, y) 1 r2 Py (5.11)
f (y)
1 1
+ f (y)2 s2 Pss + b(t, s, y)2 Pyy + rb(t, s, y)s f (y) Psy = 0
2 2
p
dYt = a(m Yt )dt + s Yt dẐt (5.13)
Where, as before, Ẑt is defined as:
q
Ẑt = rWt + 1 r2 Zt (5.14)
Note that 5.12 and 5.13 can be expressed in the form 3.5, with:
1 1
+ ys2 Css + s2 yCyy + rsysCsy = 0
2 2
Where
⇣ q ⌘
p
l(t, s, y) = s r(µ r) + yg(t, s, y) 1 r2 (5.18)
h i h i
r ( T t) r ( T t)
= EQg e S T 1{ S T K} Ft e KEQg 1{ST K} Ft (5.20)
he r ( T t) S i h i
T r ( T t)
= St E Q g 1{ S T K} Ft e KEQg 1{ST K} Ft (5.21)
St
= St P1 KP(t, T ) P2 (5.22)
h i
e r ( T t) ST
Where P(t, T ) = e r(T t) , P1 = EQg St 1{ S T K } Ft ,
h i
and P2 = EQg 1{ST K } Ft .
Thus, the equation for P(t, T ) P2 is exactly the same as 5.17. Consider the
change of variables x = ln(s), so that:
∂ ∂ 1
= ,
∂s ∂x s
∂2 ∂2 1 1 ∂
=
∂s2 ∂x2 s2 s2 ∂x
Which gives the following PDE for P2 :
1 ∂2 P ∂2 P2
+ s2 y 22 + rsy =0
2 ∂y ∂x∂y
With boundary condition P2 ( T, x, y) = 1{ x log(k)} .
Now, substituting C (t, s, y) in 5.17 and using the previous result, we obtain a
PDE for P1 :
∂P1 ⇣ ∂P ⌘ ⇣ ⌘ ∂P
1 1
s rsP1 + rs + P1 + s a(m y) ly
∂t ∂x ∂y
1 ⇣ ∂2 P1 ∂P1 ⌘ 1 ⇣ ∂P1 ⌘
ys +2 + P1 ys + P1
2 ∂x2 ∂x 2 ∂x
1 2 ∂2 P1 ⇣ ∂2 P ∂P ⌘
1
s ys 2 + rsys + 1 =0
2 ∂y ∂x∂y ∂y
Reordering this expression, we obtain the following PDE for P1 :
1 ∂2 P 1 ∂2 P ∂2 P1
+ y 21 + s2 y 21 + rsy =0
2 ∂x 2 ∂y ∂x∂y
With the boundary condition:
44 Stochastic Volatility
P1 ( T, x, y) = 1{ x log(K )}
Pj ( T, x, y) = 1{ x log(K )}
Where
1
u1 =
2
1
u2 =
2
a j = am, j = 1, 2
b1 = a + l rs
b2 = a + l
q
dy j (t) = a j b j y j (t) dt + s y j (t)dẐt
Assume that f j (t) = f j (t, x j (t), y j (t)) is sufficiently regular to apply the Itô
formula. In this case,
⇣∂f ∂ fj ∂ fj
j
d f j (t) = + r + u j y j (t) + aj b j y j (t) (5.29)
∂t ∂x ∂y
1 ∂2 f j 1 2 ∂2 f j ∂2 f j ⌘
+ y j (t) 2 + s y j (t) 2 + rsy j (t) dt
2 ∂x 2 ∂y ∂x∂y
⇣ ∂ fj ⌘ ⇣ ∂ fj ⌘
+ r + u j y j (t) dWt + a j b j y j (t) dẐt
∂x ∂y
Where all the partial derivatives of f are evaluated at (t, x j (t), y j (t)).
⇥ h ⇥ ⇤ i
E f j (t) Fs ]= E E 1{ x j (T ) log(K )} x j (t) = x, y j (t) = y x = x j (t),y=y j (t)
Fs
5.3 Heston’s model 45
h ⇥ ⇤ i
E E 1{ x j ( T ) log(K )} Ft Fs = f j (s)
Thus, the drift term in 5.29 must vanish. We obtain the PDE:
∂ fj ∂ fj ∂ fj
+ r + u j y j (t) + aj b j y j (t) (5.30)
∂t ∂x ∂y
1 ∂2 f j 1 2 ∂2 f j ∂2 f j
+ y j (t) 2 + s y j (t) 2 + rsy j (t) =0
2 ∂x 2 ∂y ∂x∂y
Clearly, f j must satisfy the boundary condition
f j ( T, x, y) = 1{ x log(K )}
⇣ ⌘
= P xj (T ) log(K ) x j (t) = x, y j (t) = y
The probabilities 5.31 are not easily obtained in a closed form. However,
consider the characteristic function g j defined as:
h i
ifx j ( T )
g j (t, x, y; f) = E e x j (t) = x, y j (t) = y
We admit (see Heston 1993, Appendix p.341) that g j can be expressed as:
a⇣ 1 c j edt ⌘
Cj (t; f) = rfit + 2 (b j rsfi + d j )t 2log (5.33)
s 1 cj
bj rsfi + d j 1 ed j t
D j (t; f) = (5.34)
s2 1 ced j t
and
bj rsfi + d
cj = (5.35)
bj rsfi d
q
dj = (rsfi b j )2 s2 (2u j fi f2 ) (5.36)
To obtain the desired result, the characteristic functions can be inverted in the
following way:
Z • he iflog(K ) g i
1 1 j ( t, x, y; f )
Pj (t, x, y) = + Re df (5.37)
2 p 0 if
The following are illustrative plots of the call option price in Heston’s model,
according to the obtained closed-form solution. The underlying code has been
obtained from Roberts n.d.
46 Stochastic Volatility
Figure 5.1: Heston pricing of a call option according to the closed-form solu-
tion (the dashed line) at time t = 0, with K = 100, T = 2, r = 0.04, a = 0.5,
r = 0.5, Y0 = 0.1, m = 0.1, s = 0.1 compared to the Black-Scholes call price
(the regular line) at time t = 0, with K = 100, T = 2, r = 0.04 and s = 0.1,
and the payoff function ( x K )+ (the thick line).
Figure 5.2: Heston pricing of a call option according to the closed-form solu-
tion (the dashed line) at time t = 0, with X0 = 80, K = 100, T = 2, r = 0.04,
a = 1, r = 0.5, m = 0.1, s = 0.1 compared to the Black-Scholes call price (the
regular line) at time t = 0, with K = 100, T = 2, r = 0.04 and S0 = 80.
Chapter 6
Conclusions
47
48 Conclusions
Chapter 7
Bibliography
Bibliography: Articles
Mandelbrot, Benoit (1963). ?The variation of certain speculative prices? In: The
journal of business 36.4, pp. 394–419.
Black, Fischer and Myron Scholes (1973). ?The pricing of options and corpo-
rate liabilities? In: Journal of political economy 81.3, pp. 637–654.
Heston, Steven L (1993). ?A closed-form solution for options with stochastic
volatility with applications to bond and currency options? In: Review of
financial studies 6.2, pp. 327–343.
Courtault, Jean-Michel et al. (2000). ?Louis Bachelier on the centenary of The-
orie de la speculation? In: Mathematical Finance 10.3, pp. 339–353.
Bibliography: Books
Bachelier, Louis (1900). Theorie de la speculation. Gauthier-Villars.
Fouque, Jean-Pierre, George Papanicolaou, and K Ronnie Sircar (2000). Deriva-
tives in financial markets with stochastic volatility. Cambridge University Press.
Rogers, L Chris G and David Williams (2000). Diffusions, Markov processes and
martingales: Volume 2, Ito calculus. Vol. 2. Cambridge university press.
Jacod, Jean and Philip Protter (2004). Probability essentials. Springer Science &
Business Media.
Kallenberg, Olav (2006). Foundations of modern probability. Springer Science &
Business Media.
Musiela, Marek and Marek Rutkowski (2006). Martingale methods in financial
modelling. Vol. 36. Springer Science & Business Media.
Lamberton, Damien and Bernard Lapeyre (2007). Introduction to stochastic cal-
culus applied to finance. CRC press.
Bachelier, Louis (2011). Louis Bachelier’s theory of speculation: the origins of mod-
ern finance. Princeton University Press.
Gatheral, Jim (2011). The volatility surface: a practitioner’s guide. Vol. 357. John
Wiley & Sons.
Capinski, Marek and Ekkehard Kopp (2012). The Black–Scholes Model. Cam-
bridge University Press.
Capinski, Marek, Ekkehard Kopp, and Janusz Traple (2012). Stochastic calculus
for finance. Cambridge University Press.
Karatzas, Ioannis and Steven Shreve (2012). Brownian motion and stochastic cal-
culus. Vol. 113. Springer Science & Business Media.
Cohn, Donald L (2013). Measure theory. Springer.
49
50 Bibliography
Bibliography: Online
Peter Morters, Yuval Peres (2008). Brownian Motion. url: https://www.stat.
berkeley.edu/~peres/bmbook.pdf.
Corcuera, Jose Manuel. Introduction to Quantitative Finance. url: http://www.
ub.edu/plie/personal_PLiE/corcuera_HTML/ifqfe3.pdf.
Roberts, Dale. Implementations of the Heston stochastic volatility model. url: https:
//github.com/daleroberts/heston.