Vous êtes sur la page 1sur 46

Pricing Volatility Swaps Under Hestons Stochastic

Volatility Model with Regime Switching


Robert J. Elliott

Tak Kuen Siu

Leunglung Chan

January 16, 2006
Abstract
We develop a model for pricing volatility derivatives, such as variance swaps and
volatility swaps under a continuous-time Markov-modulated version of the stochastic

The Corresponding Author: RBC Financial Group Professor of Finance, Haskayne School of
Business, University of Calgary, Calgary, Alberta, Canada, T2N 1N4; Email: relliott@ucalgary.ca;
Fax: 403-770-8104; Tel: 403-220-5540

Lecturer, Department of Actuarial Mathematics and Statistics, Heriot-Watt University, Edin-


burgh, United Kingdom; Email: T.K.Siu@ma.hw.ac.uk

Ph.D. candidate, Department of Mathematics and Statistics, University of Calgary, Calgary,


Alberta, Canada; Email: lchan@math.ucalgary.ca
1
volatility (SV) model developed by Heston (1993). In particular, we suppose the
parameters of our version of Hestons SV model depend on the states of a continuous-
time observable Markov chain process, which can be interpreted as the states of an
observable macroeconomic factor. The market we consider is incomplete in general,
and hence, there is more than one equivalent martingale pricing measure. We adopt
the regime switching Esscher transform used by Elliott et al. (2005) to determine
a martingale pricing measure for the valuation of variance and volatility swaps in
this incomplete market. We consider both the probabilistic approach and the partial
dierential equation, (PDE), approach for the valuation of volatility derivatives.
Key words: Regime Switching Esscher Transform; Markov-modulated Hestons
SV model; Observable Markov Chain Process; Volatility Swaps; Variance Swaps;
Regime Switching OU-process
2
1. Introduction and Summary
Volatility is one of the major features used to describe and measure the uctuations
of asset prices. It is popular as a measure of risk and uncertainty. It plays a signicant
role in three pillars of modern nancial analysis: risk management, option valuation
and asset allocation. There are dierent measures of volatility, including realised
volatility, implied volatility and model-based volatility. Realised volatility is the
standard deviation of the historical nancial returns; implied volatility is the volatility
inferred from the market option price data based on an assumed option pricing model,
such as the Black-Scholes model; model-based volatility includes both parametric and
non-parametric specications of the volatility dynamics, such as the ARCH models
introduced by Engle (1982), the GARCH models introduced by Bollerslev (1986) and
Taylor (1986), independently, and their variants. These models are introduced to
provide a better specication, measurement and forecasting of volatilities of various
nancial assets. Recent major nancial issues, such as the collapse of LTCM, the
Asian nancial crisis and the problems of Barings and Orange Country, reveal that
the global nancial markets have become more volatile. Due to large and frequent
shifts in the volatilities of various assets in the recent past, there has been a growing
and practical need to develop some models with related nancial instruments to
hedge volatility risk. On the other hand, market speculators may be interested in
3
guessing the direction that the volatility may take in the future. These practical needs
facilitate the growth of the market for derivative products related to volatility. The
initial suggestions for such products included options and futures on the volatility
index. Brenner and Galai (1989, 1993) proposed the idea of developing a volatility
index. The Chicago Board Options Exchange (CBOE) introduced a volatility index
in 1993, called VIX, which was based on implied volatilities from options on the S&P
500 index. The VIX Index reects the market expectations of near-term volatility
inferred from the S&P 500 stock index option prices. For a detailed discussion about
various volatility derivative products, see Brenner et al. (2001).
Variance swaps and volatility swaps are popular volatility derivative products
and they have been actively traded in over-the-counter markets since the collapse
of LTCM in late 1998. In particular, the variance swaps and the volatility swaps
on stock indices, currencies and commodities are quoted and traded actively. These
products are popular among market practitioners as a hedge for volatility risk. The
variance swap is a forward contract in which the long position pays a xed amount
K
var
/$1 nominal value at the maturity date and receives the oating amount (
2
)
R
/$1
nominal value, where K
var
is the strike price and (
2
)
R
is the realized variance.
The volatility swap is the same as the variance swap, except that the realized
variance (
2
)
R
is replaced by the realized volatility ()
R
.
4
Recently, there has been a considerable interest in pricing and hedging variance
swaps and volatility swaps. Gr unbuchler and Longsta (1996) developed pricing
models for options on variance based on Hestons stochastic volatility model. Carr
and Madan (1998) showed how derivatives on volatility can be valued. Demeter
et al. (1999) discussed the properties and valuation of volatility swaps. Heston
and Nandi (2000) discussed the valuation of options and swaps on variance and
volatility in the context of discrete-time GARCH models. Brockhaus and Long (2000)
discussed the pricing issues of volatility swaps based on Hestons stochastic volatility
model. Javaheri et al. (2002) developed a partial dierential equation approach for
pricing volatility swaps under a continuous-time version of the GARCH(1, 1) model.
Brenner et al. (2001) considered the use of the Stein-Stein model to value volatility
swaps. Matytsin (2000) considered the valuation of volatility swaps and options on
variance for a stochastic volatility model with jump-diusion dynamics. Howison et
al. (2004) also considered the valuation of volatility swaps for a stochastic volatility
model driven by a jump-diusion. They used an asymptotic approximation for the
solution of the partial dierential equation for pricing. Carr et al. (2005) valued
options on variance which were based on quadratic variation. Swishchuk (2004) used
an alternative probabilistic approach to value variance and volatility swaps under
the Heston (1993) stochastic volatility model. Elliott and Swishchuk (2004) valued
5
variance swaps in the context of a fractional Black-Scholes market. Swishchuk (2005)
developed a model for valuing variance swaps under a stochastic volatility model with
delay.
In this paper, we develop a model for pricing volatility derivatives, such as vari-
ance swaps and volatility swaps under a continuous-time Markov-modulated version
of Hestons stochastic volatility (SV) model, (Heston (1993)). In particular, the pa-
rameters of a continuous-time version of Hestons SV model depend on the states of
a continuous-time observable Markov chain process, which can be interpreted as the
states of an observable macroeconomic factor, such as an observable economic indica-
tor for business cycles and the sovereign ratings for the region. The market described
by the Markov-modulated model is incomplete in general, and hence, there is more
than one equivalent martingale pricing measure. We adopt the regime switching Es-
scher transform introduced by Elliott et al. (2005) to determine a martingale pricing
measure for the valuation of variance and volatility swaps. We consider both the
probabilistic approach and the PDE approach for the valuation of volatility deriva-
tives. We assume that the Markov chain process is observable in both the proba-
bilistic approach and the PDE approach. We shall document economic consequences
for the prices of the variance swaps and volatility swaps of the incorporation of the
regime-switching in the volatility dynamics by conducting a Monte Carlo experiment.
6
The next section describes the model. Section three demonstrates the use of the
probabilistic approach for pricing volatility and variance swaps under the continuous-
time Markov-modulated version of Hestons stochastic volatility model. Section four
considers the use of the PDE approach for the valuation. In Section ve, we shall con-
duct the Monte Carlo experiment for comparing the prices implied by the stochastic
volatility models with and without switching regimes. We shall document the eco-
nomic consequences for the prices of switching regimes. The nal section suggests
some possible topics for further investigation.
2. The Model
In this section, we adopt the probabilistic approach for the valuation of volatility
derivatives under a continuous-time Markov-modulated stochastic volatility model.
Our model can be considered the regime-switching augmentation of the
model by Swishchuk (2004) for pricing and hedging volatility swaps. The
Markov-modulated version of Hestons stochastic volatility model can describe the
consequences for the asset price and volatility dynamics of the transitions of the
states of an observable macroeconomic factor, which aects the asset prices and
volatility dynamics, such as an observable economic indicator of business cycles, or
the sovereign ratings of the region by some international rating agencies, such as
7
the Standard & Poors, Moodys and Fitch, etc. In particular, the parameters in the
asset price dynamics and the stochastic volatility dynamics depend on the state of
an economic indicator, which is described by the observable Markov chain.
First, consider a continuous-time nancial model with two primary securities,
namely a risk-free bond B and a risky stock S. Fix a complete probability space
(, F, P) with P being the real-world probability measure. Let T be the time index
set [0, ). On (, F, P) we consider a continuous-time nite-state observable Markov
chain X := {X
t
}
tT
with state space S which might be the set {s
1
, s
2
, . . . , s
N
}, where
s
i
R
N
, for i = 1, 2, . . . , N. The states of the Markov chain process X describe the
states of an observable economic indicator. Without loss of generality, the state space
of the chain can be identied with the set {e
1
, e
2
, . . . , e
N
} of unit vectors in R
N
.
Write (t) for the generator, or Q-matrix, [
ij
(t)]
i,j=1,2,...,N
of X. Following Elliott
et al. (1994), the following semi-martingale representation theorem for the process
X can be obtained:
X
t
= X
0
+
_
t
0
(s)X
s
ds +M
t
. (2.1)
Here {M
t
}
tT
is an R
N
-valued martingale increment process with respect to the
natural ltration generated by X.
Let W
1
:= {W
1
t
}
tT
and W
2
:= {W
2
t
}
tT
denote two correlated standard
8
Brownian Motions on (, F, P) with respect to the P-augmentation of the ltra-
tion F
W
:= {F
W
t
}
tT
, where W
t
:= (W
1
, W
2
). We assume that
Cov(dW
1
t
, dW
2
t
) = dt , (2.2)
where (1, 1) is the instantaneous correlation coecient between W
1
and W
2
.
We further suppose that the processes X is independent with (W
1
, W
2
). Pan
(2002) documented from her empirical studies on stock indices that the
correlation coecient between the diusive shocks to the volatility level
and the level of the underlying price is signicantly negative. The model
considered here can incorporate this negative correlation coecient.
Let {r(t, X
t
)}
tT
be the instantaneous market interest rate of the bond B, which
depends on the state of the economic indicator described by X; that is,
r(t, X
t
) =< r, X
t
> , t T , (2.3)
where r := (r
1
, r
2
, . . . , r
N
) with r
i
> 0 for each i = 1, 2, . . . , N and < , > denotes
the inner product in R
N
.
To simplify the notation, write r
t
for r(t, X
t
). Then, the dynamics of the price
9
process {B
t
}
tT
for the bond B are described by:
dB
t
= r
t
B
t
dt ,
B
0
= 1 . (2.4)
Suppose the expected appreciation rate {
t
}
tT
of the risky stock S depends on the
state of the economic indicator X and is described by:

t
:= (t, X
t
) =< , X
t
> , (2.5)
where := (
1
,
2
, . . . ,
N
) with
i
R, for each i = 1, 2, . . . , N.
Let {
t
}
tT
denote the long-term volatility level. We assume that
t
depends on
the state of the economic indicator X and is given by:

t
:= (t, X
t
) =< , X
t
> , (2.6)
where := (
1
,
2
, . . . ,
N
) with
i
> 0, for each i = 1, 2, . . . , N.
Let and denote the speed of mean reversion and the volatility of volatility,
respectively. (In general, we could consider a more general case where both the
speed of mean reversion and the volatility of volatility depend on the states of
the economic indicator X.) However, in order to make our model more analytically
tractable, we suppose that both are constant. Suppose that the dynamics of the price
10
process {S
t
}
tT
and the short-term volatility process := {
t
}
tT
of the risky stock
are governed by the following equations:
dS
t
=
t
S
t
dt +
t
S
t
dW
1
t
,
d
2
t
= (
2
t

2
t
)dt +
t
dW
2
t
, (2.7)
where Cov(dW
1
t
, dW
2
t
) = dt.
Note that the variance process
2
t
follows a Cox, Ingersoll and Ross (1985) process.
Let :=

1 ; Write W := {W
t
}
tT
for a standard Brownian motion, which
is independent of W
1
and X. Then, we can write the dynamics of {S
t
}
tT
and := {
t
}
tT
as:
dS
t
=
t
S
t
dt +
t
S
t
dW
1
t
,
d
2
t
= (
2
t

2
t
)dt +
t
dW
1
t
+
t
dW
t
. (2.8)
Let Y
t
denote the logarithmic return ln(S
t
/S
0
) over the interval [0, t]. Then,
Y
t
=
_
t
0
_

1
2

2
u
_
du +
_
t
0

u
dW
1
u
. (2.9)
In our model, there are three sources of randomness: X, W
1
and W
2
. Let F
X
:=
{F
X
t
}
tT
, F
W
1
:= {F
W
1
t
}
tT
and F
W
2
:= {F
W
2
t
}
tT
be the P-augmentation of the
natural ltrations generated by {X
t
}
tT
, {W
1
t
}
tT
and {W
2
t
}
tT
, respectively. Let
11
F
S
:= {F
S
t
}
tT
denote the P-augmentation of the natural ltration generated by
{S
t
}
tT
.
Our model is a regime switching version of Hestons stochastic volatility model
and is in general incomplete. There are, therefore, innitely many equivalent martin-
gale pricing measures. In the sequel, we shall adopt the regime-switching Esscher
transform developed in Elliott et al. (2005) to determine an equivalent martin-
gale pricing measure for the volatility and variance swaps. The regime-switching
Esscher transform provides market practitioners with a convenient and
exible way to determine an equivalent martingale measure in the incom-
plete market. The choice of the equivalent martingale measure can be
justied by the regime-switching minimal entrophy equivalent martingale
(MEMM) measure (See Elliott et al. (2005)). One drawback of using the
Esscher transform is that the pricing rule by the Esscher transform is not
linear, which is considered by nancial economists as a desirable property
for a pricing rule. There are other possible approaches for determining an
equivalent martingale measure in an incomplete market, for instance, the
minimum variance hedging in Due and Richardson (1991) and Schweizer
(1992). In the minimum-variance hedging, the instrinsic value process
corresponding to a given contingent claim is used as the optimal tracking
12
process. The instrinsic value process is dened by the minimal martingale
measure introduced in Follmer and Schweizer (1991) and Schweizer (1991)
and corresponds to the risk-neutral approach for valuing the claim. The
minimum-variance hedging can provide a pertinent solution to address
the pricing and hedging of a contingent claim while the Esscher transform
mainly deals with the valuation of the claim. The hedging strategies are
optimal in sense of minimizing the expected quadratic costs. The Esscher
transform can provide a convenient and intuitive way to price a claim.
Let G
t
denote the right continuous completion of the -algebra F
X
t
F
W
1
t
F
W
2
t
,
for each t T . Let
t
:= (t, X
t
,
t
) denote a regime switching Esscher process,
which is written as follows:

t
= (t, X
t
,
t
) =< (t,
t
), X
t
> , (2.10)
where (t,
t
) := ((t,
t
, e
1
), (t,
t
, e
2
), . . . , (t,
t
, e
N
)) and (t,
t
, e
i
) is F
W
2
t
measurable, for each i = 1, 2, . . . , N. So, (t, X
t
,
t
) is an N-dimensional F
X
t
F
W
2
t
-
measurable random vector.
Following Elliott et al. (2005), for such a process we dene the regime switching
Esscher transform Q

P on G
t
with respect to a family of parameters {
u
}
u[0,t]
13
by:
dQ

dP

G
t
=
exp
_
_
t
0

u
dY
u
_
E
P
_
exp
_
_
t
0

u
dY
u
_

F
X
t
F
W
2
t
_ , t T . (2.11)
Note that
_
t
0

u
dY
u
|F
X
t
F
W
2
t
N(
_
t
0

u
(
u

1
2

2
u
)du,
_
t
0

2
u

2
u
du), that is a normal
distribution with mean
_
t
0

u
(
u

1
2

2
u
)du and variance
_
t
0

2
u

2
u
du, under P. Then
given F
X
t
F
W
2
t
, the Radon-Nikodym derivative of the regime switching Esscher
transform is given by:
dQ

dP

G
t
= exp
_
_
t
0

u
dW
1
u

1
2
_
t
0

2
u

2
u
du
_
. (2.12)
The central tenet of the fundamental theorem of asset pricing, which is also called
the fundamental theorem of nance in Ross (2005), established the equivalence be-
tween the absence of arbitrage opportunities and the existence of a positive linear
pricing operator (or positive state space prices). It was rst established by Ross
(1973) in a nite state space setting. Cox and Ross (1976) provided the rst state-
ment of risk-neutral pricing. Ross (1978) and Harrsion and Kreps (1979) then ex-
tended the fundamental theorem in a general probability space and characterized the
risk-neutral pricing as the expectation of the discounted assets payo with respect
to an equivalent martingale measure. Harrison and Kreps (1979), and Harrison and
Pliska (1981, 1983) established the equivalence between the absence of arbitrage and
the existence of an equivalent martingale measure under which all discounted price
14
processes are martingale. The fundamental theorem was then extended by several
authors, including Dybvig and Ross (1987), Back and Pliska (1991), Schachermayer
(1992) and Delbaen and Schachermayer (1994). In our setting, let

:= {

t
}
tT
denote a process for the risk-neutral regime switching Esscher parameters. Due to
the presence of the uncertainty generated the processes X and W
2
, the martingale
condition is characterised by considering an enlarged ltration and requiring:
S
0
= E
Q

_
exp
_

_
t
0
r
s
ds
_
S
t

F
X
t
F
W
2
t
_
, for any t T . (2.13)
One can interpret this condition as one when information about the Markov chain
and the stochastic volatility process are known to the markets agent in advance.
Give these arguments which are similar to those in Elliott et al. (2005), it can be
shown that the martingale condition (2.11) implies that

t
:=

(t, X
t
,
t
) should be
given by

t
=
r(t, X
t
) (t, X
t
)

2
t
=
(t, X
t
,
t
)

t
, t T , (2.14)
where
t
:= (t, X
t
,
t
) G
t
is the market price of risk at time t.
Then

t
=<

(t,
t
), X
t
>, where

(t,
t
) = (
r
1

t
,
r
2

t
, . . . ,
r
N

t
). This is an
N-dimensional F
W
2
t
-measurable random vector.
15
Using (2.12), the Radon-Nikodym derivative of Q

with respect to P is given by:


dQ

dP

G
t
= exp
_
_
t
0
_
r
u

u
_
dW
1
u

1
2
_
t
0
_
r
u

u
_
2
du
_
. (2.15)
By Girsanovs theorem,

W
1
t
= W
1
t
+
_
t
0
(
r
s

s
)ds is a standard Brownian motion with
respect to {G
t
}
tT
under Q

. Since W and X are independent of W


1
, W is a
standard Brownian motion under Q

and X remains unchanged under the


change of the probability measure from P to Q

. Let
2
t
:=
2
t
(r
t

t
).
Then, the dynamics of S and under Q

are
dS
t
= r
t
S
t
dt +
t
S
t
d

W
1
t
,
d
2
t
= (
2
t

2
t
)dt +
t
d

W
1
t
+
t
dW
t
. (2.16)
Let

W
2
t
:=

W
1
t
+ W
t
. Then, the dynamics of can be written as:
d
2
t
= (
2
t

2
t
)dt +
t
d

W
2
t
. (2.17)
When there is no regime switching, (i.e. the Markov chain X is degenerate), the
risk-neutral dynamics under Q

reduce to the risk-neutral dynamics in Heston (1993).


3. The Probabilistic Approach
In this section, we assume the dynamics of the long-term volatility level {
t
}
tT
switch
over time according to one of the regimes determined by the state of an observable
Markov chain. We may interpret the states of the observable Markov chain as those of
16
some observable economic indicator. First, we shall consider the valuation of variance
swaps, which is more simple than the valuation of volatility swaps. Then, we shall
discuss the hedging of variance swaps and volatility swaps.
3.1. Valuation
A variance swap is a forward contract on annualized variance, which is the square
of the realized annual volatility. Let
2
R
(S) denote the realized annual stock variance
over the life of the contract. Then,

2
R
(S) :=
1
T
_
T
0

2
u
du . (3.1)
In practice, variance swaps are written on the realized variance evaluated
based on daily closing prices with the integral in (3.1) replaced by a dis-
crete sum. Hence, variance swaps with payos depending on the realized
variance dened in (3.1) are only approximations to those of the actual
contracts. See Javaheri et al. (2002) for discussions on this point.
Let K
v
and N denote the delivery price for variance and the notional amount
of the swap in dollars per annualized volatility point squared. Then, the payo of
the variance swap at expiration time T is given by N(
2
R
(S) K
v
). Intuitively, the
buyer of the variance swap will receive N dollars for each point by which the realized
annual variance
2
R
(S) has exceeded the variance delivery price K
v
. We can adopt the
17
risk-neutral regime switching Esscher transform Q

for the valuation of the variance


swap. In fact, the value of the variance swap can be evaluated as the expectation of
its discounted payo with respect to the measure Q

, which is exactly the same as


the value of a forward contract on future realized variance with strike price K
v
.
As in Elliott, Sick and Stein (2003), we initially consider the evaluation of the
conditional value, or price, of a derivative given the information about the sample
path of the Markov chain process from time 0 to time T, say F
X
T
. In particular, given
F
X
T
, the conditional price of the variance swap P(X) is given by:
P(X) = E
Q

[e

R
T
0
r
u
du
N(
2
R
(S) K
v
)|F
X
T
]
= e

R
T
0
r
u
du
NE
Q

(
2
R
(S)|F
X
T
) e

R
T
0
r
u
du
NK
v
. (3.2)
Hence, the valuation of the variance swap given F
X
T
can be reduced to the problem
of calculating the mean value of the underlying variance E
Q

(
2
R
(S)|F
X
T
).
Note that under Q

, the volatility dynamics can be written as:

2
t
=
2
0
+
_
t
0
(
2
s

2
s
)ds +
_
t
0

s
d

W
2
s
. (3.3)
Given F
X
T
,
2
t
is a known function of time t. Hence,
E
Q

(
2
t
|F
X
T
) =
2
0
+
_
t
0
[
2
s
E
Q

(
2
s
|F
X
T
)]ds . (3.4)
18
This implies that
dE
Q

(
2
t
|F
X
T
)
dt
= [
2
t
E
Q

(
2
t
|F
X
T
)] . (3.5)
Solving (3.5) gives:
E
Q

(
2
t
|F
X
T
) =
2
0
e
t
+
_
t
0

2
s
e
(ts)
ds . (3.6)
By Itos lemma,

4
t
=
4
0
+
_
t
0
[2
2
s
(
2
s

2
s
) +
2

2
s
]ds +
_
t
0
2
3
s
d

W
2
s
. (3.7)
Hence,
dE
Q

(
4
t
|F
X
T
)
dt
= (2
2
t
+
2
)E
Q

(
2
t
|F
X
T
) 2E
Q

(
4
t
|F
X
T
) . (3.8)
Solving (3.8) gives:
E
Q

(
4
t
|F
X
T
) =
4
0
e
2t
+
_
t
0
e
2(ts)
(2
2
s
+
2
)
_

2
0
e
s
+
_
s
0

2
u
e
(su)
du
_
ds . (3.9)
Hence,
V ar
Q

(
2
t
|F
X
T
)
= E
Q

(
4
t
|F
X
T
) E
2
Q

(
2
t
|F
X
T
)
=

2
0

(e
t
e
2t
) +
_
t
0
_
e
2(ts)
_
2
2

2
s
+
2
_
_
s
0

2
u
e
(su)
du
_
ds

2
_
_
t
0

2
s
e
(ts)
ds
_
2
. (3.10)
19
The results for E
Q

(
2
t
|F
X
T
) and V ar
Q

(
2
t
|F
X
T
) are consistent with those in Shreve
(2004).
Write V :=
2
R
(S) to simplify the notation. Then, E
Q

(V |F
X
T
) can be calculated
as follows:
E
Q

(V |F
X
T
) =
1
T
_
T
0
_

2
0
e
t
+
_
t
0

2
s
e
(ts)
ds
_
dt
=

2
0
T
(1 e
T
) +

T
_
T
0
_
_
t
0

2
s
e
(ts)
ds
_
dt , (3.11)
We evaluate V ar
Q

(V |F
X
T
) as follows:
V ar
Q

(V |F
X
T
)
= Cov
Q

(V, V |F
X
T
)
= 1/T
2
_
T
0
_
T
0
Cov
Q

(
2
t
,
2
s
|F
X
T
)dtds (3.12)
We rst derive an expression for Cov
Q

(
2
t
,
2
s
|F
X
T
). Without loss of generality, we
suppose that t > s. Then, we dene
t,s
as follows:

t,s
:=
2
t
E
Q

(
2
t
|F
W
2
s
F
X
T
)
=
2
t

2
s
e
(ts)

_
t
s

2
u
e
(tu)
du . (3.13)
Then, it is immediate that
E
Q

(
t,s
|F
W
2
s
F
X
T
) = 0 , (3.14)
20
E
Q

(
t,s

2
s
|F
X
T
) = 0 , (3.15)
so
Cov
Q

(
t,s
,
2
s
|F
X
T
) = 0 . (3.16)
Hence,
Cov
Q

(
2
t
,
2
s
|F
X
T
)
= Cov
Q

(
t,s
+
2
s
e
(ts)
+
_
t
s

2
u
e
(tu)
,
2
s
|F
X
T
)
= e
(ts)
V ar
Q

(
2
s
|F
X
T
)
= e
(ts)
_

2
0

(e
s
e
2s
) +
_
s
0
_
e
2(su)
_
2
2

2
u
+
2
_
_
u
0

2
z
e
(uz)
dz
_
du
2
_
_
s
0

2
u
e
(su)
du
_
2
_
. (3.17)
Therefore,
V ar
Q

(V |F
X
T
)
= 1/T
2
_
T
0
_
T
0
e
(ts)
_

2
0

(e
s
e
2s
) +
_
s
0
_
e
2(su)
_
2
2

2
u
+
2
_
_
u
0

2
z
e
(uz)
dz
_
du
2
_
_
s
0

2
u
e
(su)
du
_
2
_
dtds
=

2
0

2
T
2
_
(1 e
T
)T +
1
4
(1 e
2T
)
2
_
+ 1/T
2
_
T
0
_
T
0
e
(ts)
_
_
s
0
_
e
2(su)
_
2
2

2
u
+
2
_
_
u
0

2
z
e
(uz)
dz
_
du
2
_
_
s
0

2
u
e
(su)
du
_
2
_
dtds . (3.18)
For each i = 1, 2, . . . , N, let
i
:=
i
(r
i

i
); Write := (
1
,
2
, . . . ,
N
). Given
21
F
X
T
, the conditional price of the variance swap is given by:
P(X) = e

R
T
0
r
u
du
N
_

2
0
T
(1 e
T
) +

T
_
T
0
_
_
t
0
<
2
, X
t
> e
(ts)
ds
_
dt K
v
_
. (3.19)
Consider now the valuation of the volatility swap given F
X
T
. A stock volatility
swap is a forward contract on the annualized volatility. Let K
s
denote the annualized
volatility delivery price and N is the notational amount of the swap in dollar per
annualized volatility point. Then, the payo function of the volatility swap is given
by N(
R
(S) K
s
), where
R
(S) :=
_
1
T
_
T
0

2
u
du. In other words, the payo of the
volatility swap is equal to the payo of the variance swap when
2
R
(S) is replaced by

R
(S) and K
v
is replaced by K
s
. Given F
X
T
, the conditional price of the volatility
swap is given by:
P
s
(X) = E
Q

[e

R
T
0
r
u
du
N(
R
(S) K
s
)|F
X
T
]
= e

R
T
0
r
u
du
NE
Q

(
R
(S)|F
X
T
) e

R
T
0
r
u
du
NK
s
= e

R
T
0
r
u
du
NE
Q

V |F
X
T
) e

R
T
0
r
u
du
NK
s
. (3.20)
For the valuation of the volatility swap, we need to evaluate E
Q

V |F
X
T
). We
adopt the approximation for E
Q

V |F
X
T
) introduced by Brockhaus and Long (2000),
based on the second-order Taylor expansion for the function

V . This approxima-
tion method has also been adopted in Javaheri et al. (2002) and Swishchuk (2004).
22
It gives
E
Q

V |F
X
T
)
_
E
Q

(V |F
X
T
)
V ar
Q

(V |F
X
T
)
8[E
Q

(V |F
X
T
)]
3/2
, (3.21)
where the term
V ar
Q

(V |F
X
T
)
8[E
Q

(V |F
X
T
)]
3/2
is the convexity adjustment.
Hence, given F
X
T
, the conditional price of the volatility swap can be approximated
as:
P
s
(X) e

R
T
0
r
u
du
N
__
E
Q

(V |F
X
T
)
V ar
Q

(V |F
X
T
)
8[E
Q

(V |F
X
T
)]
3/2
K
s
_
. (3.22)
23
3.2. Hedging
Hedging variance swaps and volatility swaps is a challenging but practically im-
portant task. Javaheri et al. (2002) contended that hedging these products is dicult
in practice. Hence, they considered the pricing of a variance swap and a volatility
swap by the expectations of the discounted payos under the real-world probability
measure. This is an example of actuarial-based valuation method. In this section, we
shall discuss the hedging of variance swaps and volatility swaps. Dierent methods on
hedging variance swaps, have been proposed in the literature. These methods include
the simple delta hedging, the delta-gamma hedging, hedging using option portfolios,
hedging using a log contract and the vega hedging, etc. For a comprehensive overview
of various hedging strategies, see Demeter et al. (1999), Howison et al. (2004) and
Windcli et al. (2003). Simple delta hedging does not work well since it is an implicit
linear approximation, which cannot incorporate the eects of large price movements
and the realized variance or volatility will increase substantially when the underlying
share prices move either up or down dramatically. This is the case even one considers
a Geometric Brownian Motion for the price dynamics of the underlying share. Simple
delta hedging is even more dicult to apply when one considers a stochastic volatility
model and a regime-switching stochastic volatility model, which is even more com-
plicated. Hedging using option portfolios and hedging using a log contract works
24
well when one considers an asset price model with non-stochastic volatility. The vega
hedging provides market practitioners with a convenient way to hedge variance swaps
and volatility swaps under a stochastic volatility model, in particular, the Heston SV
model (see Howison et al. (2004) and Carr (2005)). Howison et al. (2004) considered
the use of Vega to hedge volatililty derivatives and derived a general formula for the
Vega of a volatility derivative. Here, following Howison et al. (2004), we adopt the
Vega hedging for a variance swap and a volatility swap since it can provide a conve-
nient way to hedge these products under the regime-switching Hestons SV model. In
the case of the volatility swap, we shall derive an approximate formula for the Vega
of the contract based on the approximate price of the contract.
First, we consider the hedging of the variance swap. Let I
t
:=
_
t
0

2
u
du. Write
F

W
2
t
for the P-augmentation of the -algebra generated by the values of

W
2
up to
and including time t. Note that given F
X
T
, F

W
2
t
is equivalent to F
W
2
t
. Then, using
the results in Section 3.1, the price of the variance swap P(t, X) at time t is given
by:
P(t, X) =
1
T
e

R
T
t
r
u
du
N
_
I
t
+E
Q

_
_
T
t

2
u
du

F
X
T
F
W
2
t
_
TK
v
_
=
1
T
e

R
T
t
r
u
du
N
_
I
t
+

2
t

(e
t
e
T
) +
_
T
t
_
_
s
t
<
2
, X
u
>
e
(su)
du
_
ds TK
v
_
. (3.23)
25
Let
2
t
:=
2
t
, which represents the instantaneous volatility of the variance process
at time t. Then, the Vega of the variance swap is given by:
P(t, X)

=
2N
t
T
2
(e
t
e
T
)
=
2N
t
T
(e
t
e
T
) , (3.24)
which can be evaluated given the current value of the volatility level
t
.
We shall consider the hedging of the volatility swap. First, we dene R
1
(t,
2
t
),
R
2
(t,
2
t
) and R
3
(t,
2
t
) as follows:
R
1
(t,
2
t
) = I
t
+

2
t

(e
t
e
T
) +
_
T
t
_
_
s
t
<
2
, X
u
> e
(su)
du
_
ds ,
R
2
(t,
2
t
) =

2
t

2
T
2

3
[e
(Tt)
2

e
2(Tt)
] +
1
T
2
_
T
t
_
T
t
_
e
(hs)
_
s
0
_
e
2(su)
_
2
2

2
u
+
2
_
_
u
0

2
z
e
(uz)
dz
_
du
2
_
_
s
0

2
u
e
(su)
du
_
2
_
dhds ,
and
R
3
(t,
2
t
) =

2
t

2
T
2
_
(1 e
t
)t +
1
4
(1 e
2t
)
2
_
+ 1/T
2
_
t
0
_
t
0
_
e
(us)
_
s
0
_
e
2(su)
_
2
2

2
u
+
2
_
_
u
0

2
z
e
(uz)
dz
_
du
2
_
_
s
0

2
u
e
(su)
du
_
2
_
duds .
26
From the results in Section 3.1, the price of the volatility swap P
s
(t, X) at time t can
be approximated as:
P
s
(t, X) e

R
T
t
r
u
du
N
__
E
Q

(V |F
X
T
F
W
2
t
)
V ar
Q

(V |F
X
T
F
W
2
t
)
8[E
Q

(V |F
X
T
F
W
2
t
)]
3/2
K
s
_
= e

R
T
t
r
u
du
N
_
_
R
1
(t,
2
t
)
R
3
(t,
2
t
) +R
2
(t,
2
t
)
8R
1
(t,
2
t
)
3/2
K
s
_
. (3.25)
Write R
i
for R
i
(t,
2
t
) (i = 1, 2, 3). Then, the Vega of the volatility swap is approxi-
mated as:
P
s
(t, X)

= e

R
T
t
r
u
du
N
_

R
1
1/2
(e
t
e
T
) +
3
t
8
(R
3
+R
2
)R
1
5/2
(e
t
+e
T
) +

t

2
R
1
4T
2

3
_
e
(Tt)
2
+
1

e
2(Tt)
__
, (3.26)
which can be evaluated given the current value of the volatility level
t
.
4. The P.D.E. Approach
In this section, we adopt a partial dierential equation (P.D.E.) approach for
evaluating the expectations of the discounted values of V and V
2
, which are useful
for computing the prices of the variance swaps and volatility swaps. The P.D.E.
approach has been adopted by Javaheri, et al. (2002) for the valuation and hedging of
volatility swaps within the framework of a GARCH(1, 1) stochastic volatility model.
Here, we provide a regime switching modication of the problem and derive regime
switching P.D.E.s and the corresponding systems of coupled P.D.E.s satised by the
27
expectations of the discounted values of V and V
2
. We adopt a regime switching
version of the Feyman-Kac formula to obtain the regime switching P.D.Es. The
derivation of the regime switching version of the Feyman-Kac formula follows from
the martingale approach and Itos dierentiation rule in Bungton and Elliott (2002).
First, let V
t
:=
2
R,t
(S) :=
1
T
_
t
0

2
u
du. Then, given F
W
2
t
F
X
T
, the price of the
variance swap is given by:
P(X, t) = e

R
T
t
r
u
du
NE
Q

(
2
R,T
(S)|F
W
2
t
F
X
T
) e

R
T
t
r
u
du
NK
v
, (4.1)
and the price of the volatility swap is given by:
P
s
(X, t)
= e

R
T
t
r
u
du
NE
Q

(
R,T
(S)|F
W
2
t
F
X
T
) e

R
T
t
r
u
du
NK
s
. (4.2)
Now, suppose
2
t
= , X
t
= X and V
t
= V are given at time t. Then, the price of
the variance swap is given by:
P(X, , V, t) = E
Q

(P(X, t)|
2
t
= , V
t
= V, X
t
= X) , (4.3)
and the price of the volatility swap is given by:
P
s
(X, , V, t) = E
Q

(P
s
(X, t)|
2
t
= , V
t
= V, X
t
= X) . (4.4)
Bungton and Elliott (2002a, b) adopted a similar method to determine the price of
a standard European call option.
28
By the double expectation formula,
P(X, , V, t)
= NE
Q

_
e

R
T
t
r
u
du

2
R,T
(S) e

R
T
t
r
u
du
K
v

t
= , V
t
= V, X
t
= X
_
, (4.5)
and
P
s
(X, , V, t)
= NE
Q

_
e

R
T
t
r
u
du

R,T
(S) e

R
T
t
r
u
du
K
s

2
t
= , V
t
= V, X
t
= X
_
. (4.6)
Hence, for the evaluation of P(X, , V, t) and P
s
(X, , V, t), we need to compute
1. E
Q

_
e

R
T
t
r
u
du

2
t
= , V
t
= V, X
t
= X
_
2. E
Q

_
e

R
T
t
r
u
du

2
R,T
(S)

2
t
= , V
t
= V, X
t
= X
_
3. E
Q

_
e

R
T
t
r
u
du

R,T
(S)

2
t
= , V
t
= V, X
t
= X
_
The rst expectation is equal to the value of a zero-coupon bond at time t given that

2
t
= , V
t
= V, X
t
= X, which pays one unit of account at maturity time T. It is
given by:
E
Q

_
e

R
T
t
r
u
du

2
t
= , V
t
= V, X
t
= X
_
= E
Q

_
e

R
T
t
r
u
du

X
t
= X
_
:= B(t, T, X) . (4.7)
29
Let B := diagr

, where diagr is the matrix with the vector r on its diagonal.


Write
t
(r) := exp[B(T t)]I. Then, by Elliott and Kopp (2004), B(t, T, X) is
given by:
B(t, T, X) =<
t
(r), X >=< exp[B(T t)], X > I . (4.8)
The third expectation can be approximated by the using the formula in Section
2. More specically,
E
Q

_
e

R
T
t
r
u
du

R,T
(S)

2
t
= , V
t
= V, X
t
= X
_

_
E
Q

_
e
2
R
T
t
r
u
du

2
R,T
(S)

2
t
= , V
t
= V, X
t
= X
_

V ar
Q

_
e
2
R
T
t
r
u
du

2
R,T
(S)

2
t
= , V
t
= V, X
t
= X
_
8
_
E
Q

_
e
2
R
T
t
r
u
du

2
R,T
(S)

2
t
= , V
t
= V, X
t
= X
__ . (4.9)
Hence, in order to provide an approximation to the third expectation, we need to
evaluate
1. E
Q

(e
4
R
T
t
r
u
du

4
R,T
(S)|
2
t
= , V
t
= V, X
t
= X)
2. E
Q

(e
2
R
T
t
r
u
du

2
R,T
(S)|
2
t
= , V
t
= V, X
t
= X).
Suppose H
t
:= {W
2
u
, X
u
|u [0, t]}. Since V
t
is a path integral of
2
t
and
2
t
is a Markov process given knowledge of X, (V
t
,
2
t
) is a two-dimensional Markov
30
process given the knowledge of X. Since X is also a Markov process, (X
t
,
2
t
, V
t
) is a
three-dimensional Markov process with respect to the information set H
t
. Hence,
M
1
(X, , V, t) := E
Q

(e

R
T
t
r
u
du

2
R,T
(S)|
2
t
= , V
t
= V, X
t
= X)
= E
Q

(e

R
T
t
r
u
du

2
R,T
(S)|H
t
) , (4.10)
M
2
(X, , V, t) := E
Q

(e
2
R
T
t
r
u
du

2
R,T
(S)|
2
t
= , V
t
= V, X
t
= X)
= E
Q

(e
2
R
T
t
r
u
du

2
R,T
(S)|H
t
) , (4.11)
and
M
3
(X, , V, t) := E
Q

(e
4
R
T
t
r
u
du

4
R,T
(S)|
2
t
= , V
t
= V, X
t
= X)
= E
Q

(e
4
R
T
t
r
u
du

4
R,T
(S)|H
t
) , (4.12)
Now, write

M
1
(X, , V, t) := e

R
t
0
r
u
du
M
1
(X, , V, t)
= E
Q

(e

R
T
0
r
u
du

2
R,T
(S)|H
t
) , (4.13)

M
2
(X, , V, t) := e
2
R
t
0
r
u
du
M
2
(X, , V, t)
= E
Q

(e
2
R
T
0
r
u
du

2
R,T
(S)|H
t
) , (4.14)
and

M
3
(X, , V, t) := e
4
R
t
0
r
u
du
M
3
(X, , V, t)
31
= E
Q

(e
4
R
T
0
r
u
du

4
R,T
(S)|H
t
) . (4.15)
Then, it can be shown that

M
1
,

M
2
and

M
3
are H
t
-martingales under Q

.
In the sequel, we shall derive the P.D.E. for

M
1
,

M
2
and

M
3
. For each i = 1, 2, 3,
let

M
i
(, V, t) denote the N-dimensional vector (

M
i
(e
1
, , V, t), . . . ,

M
i
(e
N
, , V, t)).
Then,

M
i
(X, , V, t) =<

M
i
(, V, t), X
t
> . (4.16)
Then, by applying Itos dierentiation rule to

M
i
(X, , V, t),

M
i
(X, , V, t) =

M
i
(X, , V, 0) +
_
t
0
_


M
i
u
+(
2
u

2
u
)


M
i

+
1
2

2
u

2

M
i

2
+
2
u


M
i
V
_
du +
_
t
0


M
i


u
d

W
2
u
+
_
t
0
<

M
i
, dX
u
> , (4.17)
and
dX
t
= (t)X
t
dt +dM
t
. (4.18)
Due to the fact that

M
1
,

M
2
and

M
3
are H
t
-martingale under Q

, all terms with


bounded variation in the above Itos intergral representation for

M
i
(i = 1, 2, 3) must
be identical to zero. Hence, for i = 1, 2, 3,

M
i
satises the following P.D.E.:


M
i
t
+(
2
t

2
t
)


M
i

+
1
2

2
t

2

M
i

2
+
2
t


M
i
V
+ <

M
i
, X >= 0 . (4.19)
32
Now, let M
i
(, V, t) denote the N-dimensional vector (M
i
(e
1
, , V, t), . . . , M
i
(e
N
, , V, t)),
where i = 1, 2, 3. Then,
M
i
(X, , V, t) =< M
i
(, V, t), X
t
> . (4.20)
M
1
satises the following P.D.E.:
_
exp
_

_
t
0
r
u
du
___
r
t
M
1
+
M
1
t
+(
2
t

2
t
)
M
1

+
1
2

2
t

2
M
1

2
+
2
t
M
1
V
+ < M
1
, X >
_
= 0 , (4.21)
with terminal condition M
1
(X, , V, T) = V
T
.
M
2
satises the following P.D.E.:
_
exp
_
2
_
t
0
r
u
du
___
2r
t
M
2
+
M
2
t
+(
2
t

2
t
)
M
2

+
1
2

2
t

2
M
2

2
+
2
t
M
2
V
+ < M
2
, X >
_
= 0 , (4.22)
with terminal condition M
2
(X, , V, T) = V
T
.
M
3
satises the following P.D.E.:
_
exp
_
4
_
t
0
r
u
du
___
4r
t
M
3
+
M
3
t
+(
2
t

2
t
)
M
3

+
1
2

2
t

2
M
3

2
+
2
t
M
3
V
+ < M
3
, X >
_
= 0 , (4.23)
with terminal condition M
3
(X, , V, T) = V
2
T
.
33
Note that with X = e
j
(j = 1, 2, . . . , N),
r
t
= < r, X
t
>= r
j
,

t
= < , X
t
>=
j
. (4.24)
Let M
ij
:= M
i
(e
j
, , V, T), where i = 1, 2, 3 and j = 1, 2, . . . , N. Then, M
i
=
(M
i1
, M
i2
, . . . M
iN
). Hence, M
1
satises the following system of N coupled P.D.E.s:
r
j
M
1j
+
M
1j
t
+(
2
j

2
t
)
M
1j

+
1
2

2
t

2
M
1j

2
+
2
t
M
1j
V
+ < M
1
, e
j
>= 0 , (4.25)
with terminal condition M
1
(e
j
, , V, T) = V
T
.
M
2
satises the following system of N coupled P.D.E.s:
2r
j
M
2j
+
M
2j
t
+(
2
j

2
t
)
M
2j

+
1
2

2
t

2
M
2j

2
+
2
t
M
2j
V
+ < M
2
, e
j
>= 0 , (4.26)
with terminal condition M
2
(e
j
, , V, T) = V
T
.
M
3
satises the following system of N coupled P.D.E.s:
4r
j
M
3j
+
M
3j
t
+(
2
j

2
t
)
M
3j

+
1
2

2
t

2
M
3j

2
+
2
t
M
3j
V
+ < M
3
, e
j
>= 0 , (4.27)
with terminal condition M
3
(e
j
, , V, T) = V
2
T
.
34
Once M
1
, M
2
and M
3
are solved from the above systems of N coupled P.D.E.s,
we can use them to approximate the prices of the variance swap and the volatility
swap.
5. Monte Carlo Experiment
In this section, we shall perform a Monte Carlo Experiment for the prices of the
variance swaps and the volatility swaps implied by the regime-switching Hestons
stochastic volatility model. We shall document economic consequences for the prices
of the variance swaps and the volatility swaps of a regime-switching in Hestons
stochastic volatility model by comparing the prices with those obtained from Hestons
SV model without regime-switching. We shall compute the prices of the variance
swaps and the volatility swaps with various delivery prices under both the regime-
switching Hestons SV model and Hestons SV model without switching regimes by
Monte Carlo simulation. For illustration, we suppose that the number of regimes N =
2 throughout this section. The rst and second regimes, namely X
t
= 1 and X
t
= 2,
can be interpreted as the Good and Bad economic states, respectively. We also
assume that Hestons SV model without switching regimes coincides with the rst
regime of the regime-switching Hestons SV model. In this case, we can investigate
economic consequences for the prices of the variance swaps and the volatility swaps
35
when we allow the possibility that the dynamics of Hestons SV model switches over
time to the one corresponding to the Bad economic states. We generate 10,000
simulation runs for computing each price. All computations were done by C++
codes with GSL functions.
We shall assume some specimen values for the parameters of regime-switching
Hestons SV model and the one without switching regimes. When the economy is
good (bad), the interest rate is high (low). Let r
1
and r
2
denote the annual interest
rates for the Good state and the Bad state, respectively. Then, we suppose that
r
1
= 5% and r
2
= 2%. The appreciation rate of the underlying risky asset is high
(low) when the economy is good (bad). In each case, the appreciation rate should
be higher than the corresponding interest rate. Hence, we suppose that the annual
appreciation rate
1
= 7% for the Good state and the annual appreciation rate

2
= 5% for the Bad state. When the economy is good (bad), the underlying risky
asset is less (more) volatile. Hence, we suppose that
1
= 0.12 and
2
= 0.24. The
speed of mean reversion = 0.2 and the volatility of volatility parameter = 0.08.
We also assume that the correlation coecient is negative and is equal to 0.5.
The transition probabilities of the Markov chain are
11
= 0.5,
12
= 0.5,
21
= 0.5,

22
= 0.5. The notational amount of the variance swap or the volatility swap is 1
million. We suppose that the current economic state X
0
= 1 and that the current
36
volatility level V
0
= 0.12. The delivery prices of the variance swap and the volatility
swap range from 80% to 125% of the current levels of the variance and the standard
deviation of the underlying risky asset, respectively. The time-to-expiry of both the
variance swap and the volatility swap is 1 year. Since the regime-switching Heston
stochastic volatility model is a continuous-time model, we need to discretize it when
we compute the prices of the variance swaps and volatility swaps by Monte Carlo
simulation. We suppose that the number of steps for the discretization is 20. Table
1 displays the prices of the variance swaps for various delivery prices implied by
the regime-switching Heston stochastic volatility model and its non-regime-switching
counterpart.
Table 1: Prices of Variance Swaps with and without Switching Regimes
Delivery Prices Prices with Switching Regimes Prices without Switching Regimes
in % in million in million
80 1.14556 0.847634
85 1.1431 0.845326
90 1.14165 0.843224
95 1.13823 0.840911
100 1.13607 0.838839
105 1.13167 0.836526
110 1.12934 0.834374
115 1.12657 0.832061
120 1.12196 0.829986
125 1.12082 0.82767
37
Table 2 displays the prices of the volatility swaps for various delivery prices im-
plied by the regime-switching Heston stochastic volatility model and its non-regime-
switching counterpart.
Table 2: Prices of Volatility Swaps with and without Switching Regimes
Delivery Prices Prices with Switching Regimes Prices without Switching Regimes
in % in million in million
80 0.632453 0.467974
85 0.624144 0.461601
90 0.616364 0.455246
95 0.607529 0.448786
100 0.599301 0.442436
105 0.589936 0.436079
110 0.581685 0.429701
115 0.573195 0.423343
120 0.563777 0.416992
125 0.55599 0.410641
From Table 1 and Table 2, we see that the prices of the variance swaps and the
volatility swaps implied by the regime-switching Heston stochastic volatility model
are signicantly higher than the corresponding prices of the variance swaps and the
volatility swaps, respectively, implied by the standard Heston stochastic volatility
without switching regimes. This reveals that a higher risk premium is required to
compensate for the risk from the structural change of the volatility dynamics to the
one with higher long-term volatility level due to the possible transitions of the states
38
of the economy to the Bad state. This illustrates the economic signicance of
incorporating the switching regimes in the volatility dynamics for pricing variance
swaps and volatility swaps.
6. Further Research
For further investigation, it is interesting to explore and develop some criteria to de-
termine the number of states of the Markov chain in our framework which will incor-
porate important features of the volatility dynamics for dierent types of underlying
nancial instruments, such as commodities, currencies and xed income securities. It
is also interesting to explore the applications of our model to price various volatility
derivative products, such as options on volatilities and VIX futures, which are a listed
contract on the Chicago Board Options Exchange. It is also of practical interest to
investigate the calibration and estimation techniques of our model to volatility index
options. Empirical studies comparing the performance of models on volatility swaps
are interesting topics to be investigated further.
Acknowledgment
We would like to thank the referee for many valuable comments and suggestions.
39
References
1. Back, K. and Pliska, S. R. (1991) On the fundamental theorem of asset pricing
with an innite state space, Journal of Mathematical Economics, 20, pp. 1-18.
2. Bollerslev, T. (1986) Generalized autoregressive conditional heteroskedasticity,
Journal of Econometrics, 31, pp. 307-327.
3. Brenner, M., and Galai, D. (1989) New nancial instruments for hedging changes
in volatility, Financial Analyst Journal, July/August, pp. 61-65.
4. Brenner, M., and Galai, D. (1993) Hedging volatility in foreign currencies,
Journal of Derivatives, 1, pp. 53-59.
5. Brenner, M., Ou, E., and Zhang, J. (2001) Hedging volatility risk, Working
paper, Stern School of Business, New York University, United States.
6. Brockhaus, O., and Long, D. (2000) Volatility swaps made simple, Risk, 2(1),
pp. 92-95.
7. Bungton, J. and Elliott, R. J. (2002a) Regime switching and European op-
tions, In Stochastic Theory and Control, Proceedings of a Workshop, Lawrence,
K.S., October, Springer Verlag, pp. 73-81.
40
8. Bungton, J. and Elliott, R. J. (2002b) American options with regime switch-
ing, International Journal of Theoretical and Applied Finance, 5, pp. 497-514.
9. Carr, P. (2005) Delta hedging with stochastic volatility, Working paper, Quan-
titative Financial Research, Bloomberg LP, New York.
10. Carr, P., and Madan, D. (1998) Towards a theory of volatility trading, In the
book: Volatility, Risk book publications.
11. Carr, P., Geman H., Madan, D., and Yor, M. (2005) Pricing options on realized
variance, Finance and Stochastics, 9(4), pp. 453-475.
12. Cox, J., Ingersoll, J. and Ross, S. (1985) A theory of the term structure of
interest rates, Econometrica, 53, pp. 385-407.
13. Cox, J. and Ross, S. (1976) The valuation of options for alternative stochastic
processes, Journal of Financial Economics, 3, pp. 145-166.
14. Delbaen, F. and Schachermayer, W. (1994) A general version of fundamental
theorem of asset pricing, Mathematische Annalen, 300, pp. 463-520.
15. Demeter, K., Derman, E., Kamal, M., and Zou, J. (1999) A guide to volatility
and variance swaps, The Journal of Derivatives, 6(4), pp. 9-32.
41
16. Due, D. and Richardson, H. R. (1991) Mean-variance hedging in continuous
time, The Annals of Applied Probability, 1, pp. 1-15.
17. Dybvig, Philip H., and Ross, S. (1987) Arbitrage, In Eatwell J., Milgate M.,
and Newman P. eds. the New Palsgrave: A dictionary of Economics , 1, pp.
100-106.
18. Elliott, R. J., Aggoun, L. and Moore, J. B. (1994) Hidden Markov Models:
Estimation and Control. Springer-Verlag: Berlin-Heidelberg-New York.
19. Elliott, R. J., Sick, G. A. and Stein, M. (2003) Modelling electricity price risk,
Working paper, Haskayne School of Business, University of Calgary.
20. Elliott, R. J., Chan, L. and Siu, T. K. (2005) Option pricing and Esscher
transform under regime switching, Annals of Finance, 1(4), pp. 423-432.
21. Elliott, R. J. and Kopp, E. P. (2004) Mathematics of Financial Markets. Springer-
Verlag: Berlin-Heidelberg-New York.
22. Elliott, R. J., and Swishchuk, A. (2004) Pricing options and variance swaps
in Brownian and fractional Brownian market, Working paper, University of
Calgary.
42
23. Engle, R. (1982) Autoregressive conditional heteroskedasticity with estimates
of the variance of U.K. ination, Econometrica, 50, pp. 987-1008.
24. Follmer, H. and Schweizer, M. (1991) Hedging of contingent claims under
incomplete information, In: M.H.A. Davis and R. J. Elliott (eds.), Applied
Stochastic Analysis, Stochastics Monographs, Vol.5, Gordon and Breach, Lon-
don/New York, pp. 389-414.
25. Gr unbuchler, A., and Longsta, F. (1996) Valuing futures and options on
volatility, Journal of Banking and Finance, 20, pp. 985-1001.
26. Harrison, J. M. and Kreps, D. M. (1979) Martingales and arbitrage in multi-
period securities markets, Journal of Economic Theory, 20, pp. 381-408.
27. Harrison, J. M. and Pliska, S. R. (1981) Martingales and stochastic integrals in
the theory of continuous trading, Stochastic Processes and Their Applications,
11, pp. 215-280.
28. Harrison, J. M. and Pliska, S. R. (1983) A stochastic calculus model of contin-
uous trading: complete markets, Stochastic Processes and Their Applications,
15, pp. 313-316.
29. Heston, S. L. (1993) A closed-form solution for options with stochastic with
43
applications to bond and currency options, Review of Financial Studies, 6(2),
pp. 237-343.
30. Heston, S. L., and Nandi, S. (2000) Derivatives on volatility: some simple
solution based on observables, Working paper, Federal Reserve Bank of Atlanta.
31. Howison, S., Rafailidis A., and Rasmussen, H. (2004) On the pricings and
hedging of volatility derivatives, Applied Mathematical Finance, 11(4), pp. 317-
346.
32. Javaheri, A., Wilmott, P., and Haug, E. G. (2002) GARCH and volatility
swaps, Wilmott Magazine, January, pp. 1-17.
33. Matytsin, A. (2000) Modeling volatility and volatility derivatives, Working pa-
per, Columbia University.
34. Pan, J. (2002) The jump-risk premia implicit in options: evidence from an
integrated time-series study, Journal of Financial Economics, 63, pp. 3-50.
35. Ross, S. A. (1973) Return, risk and arbitrage, Wharton Discussion Paper pub-
lished in Risk and Return in Finance, edited by I. Friend and J. Bicksler, pp.
189-217. Cambridge: Ballinger, 1976.
44
36. Ross, S. A. (1978) A simple approach to the valuation of risky streams, Journal
of Business, 3, pp. 453-476.
37. Ross, S. A. (2005) Neoclassical Finance, Princeton and Oxford: Princeton
University Press.
38. Rubinstein, M. (1976) The valuation of uncertain income streams and the pric-
ing of options, Bell Journal of Economics and Management Science, 7, pp.
407-425.
39. Schachermayer, W. (1992) A Hilbert space proof of the fundamental theorem
of asset pricing in nite discrete time, Insurance: Mathematics and Economics,
11, pp. 249-257.
40. Schweizer, M. (1991) Option hedging for semimartingales, Stochastic Processes
and Their Application, 37, pp. 339-363.
41. Schweizer, M. (1992) Mean-variance hedging for general claims, The Annals of
Applied Probability, 2, pp. 171-179.
42. Shreve, S. E. (2004) Stochastic Calculus for Finance, New York: Springer.
43. Swishchuk, A. (2004) Modeling of variance and volatility swaps for nancial
markets with stochastic volatilities, Wilmott magazine, 2, pp. 64-72.
45
44. Swishchuk, A. (2005) Modeling and pricing of variance swaps for stochastic
volatilities with delay, Wilmott magazine, Forthcoming.
45. Taylor, S. J. (1986) Modelling Financial Time Series. New York: John Wiley
& Sons.
46. Windcli, H., Forsyth, P. A. and Vetzal, K. R. (2003) Pricing methods and
hedging strategies for volatility derivatives, Working paper, University of Wa-
terloo.
46

Vous aimerez peut-être aussi