ArticlePDF Available

Abstract and Figures

This paper studies the problem of determining the optimal cut-off for pairs trading rules. We consider two correlated assets whose spread is modelled by a mean-reverting process with stochastic volatility, and the optimal pair trading rule is formulated as an optimal switching problem between three regimes: flat position (no holding stocks), long one short the other and short one long the other. A fixed commission cost is charged with each transaction. We use a viscosity solutions approach to prove the existence and the explicit characterization of cut-off points via the resolution of quasi-algebraic equations. We illustrate our results by numerical simulations.
Content may be subject to copyright.
Optimal switching for pairs trading rule:
a viscosity solutions approach
Minh-Man NGO
John von Neumann (JVN) Institute
Vietnam National University
Ho-Chi-Minh City,
man.ngo at jvn.edu.vn
Huyˆen PHAM
Laboratoire de Probabilit´es et
Mod`eles Al´eatoires, CNRS UMR 7599
Universit´e Paris 7 Diderot,
CREST-ENSAE,
and JVN Institute
pham at math.univ-paris-diderot.fr
March 23, 2016
Abstract
This paper studies the problem of determining the optimal cut-off for pairs trading
rules. We consider two cointegrated assets whose spread is modelled by a general
mean-reverting process, and the optimal pair trading rule is formulated as an optimal
switching problem between three regimes: flat position (no holding stocks), long one
short the other and short one long the other. A fixed commission cost is charged with
each transaction. We use a viscosity solutions approach to prove the existence and the
explicit characterization of cut-off points via the resolution of quasi-algebraic equations.
We illustrate our results by numerical simulations.
Keywords: pairs trading, optimal switching, mean-reverting process, viscosity solutions.
MSC Classification: 60G40, 49L25.
JEL Classification: C61, G11.
1
1 Introduction
Pairs trading consists of taking simultaneously a long position in one of the assets Aand
B, and a short position in the other, in order to eliminate the market beta risk, and be
exposed only to relative market movements determined by the spread. A brief history and
discussion of pairs trading can be found in Ehrman [5], Vidyamurthy [15], Elliott et al. [6],
or Gatev et al. [7]. The main aim of this paper is to rationale mathematically these rules
and find optimal cutoffs, by means of a stochastic control approach.
Pairs trading problem has been studied by stochastic control approach in the recent
years. Mudchanatongsuk et al. [11] consider self-financing portfolio strategy for pairs
trading, model the log-relationship between a pair of stock prices by an Ornstein-Uhlenbeck
process and use this to formulate a portfolio optimization and obtain the optimal solution to
this control problem in closed form via the corresponding Hamilton-Jacobi-Bellman (HJB)
equation. They only allow positions that are short one stock and long the other, in equal
dollar amounts. Tourin and Yan [14] study the same problem, but allow strategies with
arbitrary amounts in each stock. Chiu and Wong [3], [4] investigated optimal strategies for
conintegrated assets using mean-variance criterion and CRRA utility function. We mention
also the recent paper by Liu and Timmermann [10] who studied optimal trading strategies
for cointegrated assets with both fixed and random Poisson horizons. On the other hand,
instead of using self-financing strategies, one can focus on determining the optimal cut-offs,
i.e. the boundaries of the trading regions in which one should trade when the spread lies
in. Such problem is closely related to optimal buy-sell rule in trading mean reverting asset.
Zhang and Zhang [16] studied optimal buy-sell rule, where they model the underlying asset
price by an Ornstein-Uhlenbeck process and consider an optimal trading rule determined
by two regimes: buy and sell. These regimes are defined by two threshold levels, and
a fixed commission cost is charged with each transaction. They use classical verification
approach to find the value function as solution to the associated HJB equations (quasi-
variational inequalities), and the optimal thresholds are obtained by smooth-fit technique.
The same problem is studied in Kong’s PhD thesis [8], but he considers trading rules with
three aspects: buying, selling and shorting. Song and Zhang [13] use the same approach
for determining optimal pairs trading thresholds, where they model the difference of the
stock prices Aand Bby an Ornstein-Uhlenbeck process and consider an optimal pairs
trading rule determined by two regimes: long Ashort Band flat position (no holding
stocks). Leung and Li [9] studied the optimal timing to open or close the position subject
to fixed transaction costs (entry and exit), and the effect of Stop-loss level under the
Ornstein-Uhlenbeck (OU) model. They directly construct the value functions instead of
using variational inequalities approach, by characterizing the value functions as the smallest
concave majorant of reward function. Lei and Xu [18] studied the optimal pairs trading
rule in finite horizon with proportional transaction cost by applying numerical method for
solving the system of variational inequalities.
In this paper, we consider a pairs trading problem as in Song and Zhang [13], but differ
in our model setting and resolution method. We consider two cointegrated assets whose
spread is modelled by a more general mean-reverting process, and the optimal pairs trading
2
rule is based on optimal switching between three regimes: flat position (no holding stocks),
long one short the other and vice-versa. A fixed commission cost is charged with each
transaction. We use a viscosity solutions approach to solve our optimal switching problem.
Actually, by combining viscosity solutions approach, smooth fit properties and uniqueness
result for viscosity solutions proved in Pham [12], we are able to derive directly the structure
of the switching regions, and thus the form of our value functions. This contrasts with the
classical verification approach where the structure of the solution should be guessed ad-hoc,
and one has to check that it satisfies indeed the corresponding HJB equation, which is not
trivial in this context of optimal switching with more than two regimes.
The paper is organized as follows. We formulate in Section 2 the pairs trading as an
optimal switching problem with three regimes. In Section 3, we state the system of varia-
tional inequalities satisfied by the value functions in the viscosity sense and the definition
of pairs trading regimes. In Section 4, we state some useful properties on the switching
regions, derive the form of value functions, and obtain optimal cutoff points by relying
on the smooth-fit properties of value functions. In Section 5, we illustrate our results by
numerical examples.
2 Pair trading problem
Let us consider the spread Xbetween two cointegrated assets, say Aand Bmodelled by a
mean-reverting process with boundaries ` {−∞,0}, and `+=:
dXt=µ(LXt)dt +σ(Xt)dWt,(2.1)
where Wis a standard Brownian motion on (Ω,F,F= (Ft)t0,P), µ > 0 and L0
are positive constants, σis a Lipschitz function on (`, `+), satisfying the nondegeneracy
condition σ > 0. The SDE (2.1) admits then a unique strong solution, given an initial
condition X0=x(`, `+), denoted Xx. We assume that `+=is a natural boundary,
`=−∞ is a natural boundary, and `= 0 is non attainable. The main examples are
the Ornstein-Uhlenbeck (OU in short) process or the inhomogenous geometric Brownian
motion (IGBM), as studied in detail in the next sections.
Suppose that the investor starts with a flat position in both assets. When the spread
widens far from the equilibrium point, she naturally opens her trade by buying the under-
priced asset, and selling the overpriced one. Next, if the spread narrows, she closes her
trades, thus generating a profit. Such trading rules are quite popular in practice among
hedge funds managers with cutoff values determined empirically by descriptive statistics.
The main aim of this paper is to rationale mathematically these rules and find optimal
cutoffs, by means of a stochastic control approach. More precisely, we formulate the pairs
trading problem as an optimal switching problem with three regimes. Let {−1,0,1}be the
set of regimes where i= 0 corresponds to a flat position (no stock holding), i= 1 denotes
a long position in the spread corresponding to a purchase of Aand a sale of B, while i=
1 is a short position in X(i.e. sell Aand buy B). At any time, the investor can decide
to open her trade by switching from regime i= 0 to i=1 (open to sell) or i= 1 (open
to buy). Moreover, when the investor is in a long (i= 1) or short position (i=1), she
3
can decide to close her position by switching to regime i= 0. We also assume that it is
not possible for the investor to switch directly from regime i=1 to i= 1, and vice-versa,
without first closing her position. The trading strategies of the investor are modelled by a
switching control α= (τn, ιn)n0where (τn)nis a nondecreasing sequence of stopping times
representing the trading times, with τn a.s. when ngoes to infinity, and ιnvalued
in {−1,0,1},Fτn-measurable, represents the position regime decided at τnuntil the next
trading time. By misuse of notations, we denote by αtthe value of the regime at any time
t:
αt=ι01{0t<τ0}+X
n0
ιn1{τnt<τn+1}, t 0,
which also represents the inventory value in the spread at any time. We denote by gij(x)
the trading gain when switching from a position ito j,i, j {−1,0,1},j6=i, for a spread
value x. The switching gain functions are given by:
g01 (x) = g10 (x) = (x+ε)
g01(x) = g10 (x) = xε,
where ε > 0 is a fixed transaction fee paid at each trading time. Notice that we do not
consider the functions g11 and g11 since it is not possible to switch from regime i=1 to
i= 1 and vice-versa. By misuse of notations, we also set g(x, i, j) = gij (x).
Given an initial spread value X0=x, the expected reward over an infinite horizon
associated to a switching trading strategy α= (τn, ιn)n0is given by the gain functional:
J(x, α) = EhX
n1
eρτng(Xx
τn, ατ
n, ατn)λZ
0
eρt|αt|dti.
The first (discrete sum) term corresponds to the (discounted with discount factor ρ > 0)
cumulated gain of the investor by using pairs trading strategies, while the last integral term
reduces the inventory risk, by penalizing with a factor λ0, the holding of assets during
the trading time interval.
For i= 0,1,1, let videnote the value functions with initial positions iwhen maximizing
over switching trading strategies the gain functional, that is
vi(x) = sup
α∈Ai
J(x, α), x (`,), i = 0,1,1,
where Aidenotes the set of switching controls α= (τn, ιn)n0with initial position α0=
i, i.e. τ0= 0, ι0=i. The impossibility of switching directly from regime i=±1 to 1 is
formalized by restricting the strategy of position i=±1: if α A1or α A1then ι1=
0 for ensuring that the investor has to close first her position before opening a new one.
3 PDE characterization
Throughout the paper, we denote by Lthe infinitesimal generator of the diffusion process
X, i.e.
Lϕ(x) = µ(Lx)ϕ0(x) + 1
2σ2(x)ϕ00(x).
4
The ordinary differential equation of second order
ρφ Lφ= 0,(3.1)
has two linearly independent positive solutions. These solutions are uniquely determined
(up to a multiplication), if we require one of them to be strictly increasing, and the other
to be strictly decreasing. We shall denote by ψ+the increasing solution, and by ψthe
decreasing solution. They are called fundamental solutions of (3.1), and any other solution
can be expressed as their linear combination. Since `+=is a natural boundary, and `
{−∞,0}is either a natural or non attainable boundary, we have:
ψ+() = ψ(`) = , ψ() = 0.(3.2)
We shall also assume that
lim
x`
x
ψ(x)= 0,lim
x→∞
x
ψ+(x)= 0.(3.3)
We refer to the classical handbook [2] for the existence, uniqueness, and properties of such
functions ψ+and ψ.
Canonical examples
Our two basic examples in finance for Xsatisfying the above assumptions are
Ornstein-Uhlenbeck (OU) process (`=−∞) :
dXt=µXtdt +σdWt,(3.4)
with µ,σpositive constants. In this case, `+=,`=−∞ are natural boundaries,
the two fundamental solutions to (3.1) are given by
ψ+(x) = Z
0
tρ
µ1exp t2
2+2µ
σxtdt, ψ(x) = Z
0
tρ
µ1exp t2
22µ
σxtdt,
and it is easily checked that condition (3.3) is satisfied.
Inhomogeneous Geometric Brownian Motion (IGBM) (`= 0) :
dXt=µ(LXt)dt +σXtdWt, X0>0,(3.5)
where µ,Land σare positive constants. In this case, `+=is a natural boundary,
`= 0 is a non attainable boundary, and the two fundamental solutions to (3.1) are
given by
ψ+(x) = xaU(a, b, c
x), ψ(x) = xaM(a, b, c
x).(3.6)
where
a=pσ4+ 4(µ+ 2ρ)σ2+ 4µ2(2µ+σ2)
2σ2>0,
b=2µ
σ2+ 2a+ 2, c =2µL
σ2,(3.7)
5
and Mand Uare the confluent hypergeometric functions of the first and second kind.
Moreover, by the asymptotic property of the confluent hypergeometric functions (see
[1]), the fundamental solutions ψ+and ψsatisfy condition (3.3), and
ψ+(0+) = 1
ca.(3.8)
In this section, we state some general PDE characterization of the value functions by
means of the dynamic programming approach. We first state a linear growth property and
Lipschitz continuity of the value functions.
Lemma 3.1 There exists some positive constant r(depending on σ) such that for a dis-
count factor ρ>r, the value functions are finite on R. In this case, we have
0v0(x)C(1 + |x|),x(`,),
λ
ρvi(x)C(1 + |x|),x(`,), i = 1,1,
and
|vi(x)vi(y)| C|xy|,x, y (`,), i = 0,1,1,
for some positive constant C.
Proof. The arguments are rather standard and the proof is rejected into the appendix. 2
In the sequel, we fix a discount factor ρ > r so that the value functions viare well-
defined and finite, and satisfy the linear growth and Lipschitz estimates of Lemma 3.1. The
dynamic programming equations satisfied by the value functions are thus given by a system
of variational inequalities:
min ρv0 Lv0, v0max v1+g01 , v1+g01 = 0,on (`,),(3.9)
min ρv1 Lv1+λ , v1v0g10 = 0,on (`,),(3.10)
min ρv1 Lv1+λ , v1v0g10 = 0,on (`,).(3.11)
Indeed, the equation for v0means that in regime 0, the investor has the choice to stay in
the flat position, or to open by a long or short position in the spread, while the equation
for vi,i=±1, means that in the regime i=±1, she has first the obligation to close her
position hence to switch to regime 0 before opening a new position. By the same argument
as in [12], we know that the value functions vi, i = 0,1,1 are viscosity solutions to the
system (3.9)-(3.10)-(3.11), and satisfied the smooth-fit C1condition.
Let us introduce the switching regions:
Open-to-trade region from the flat position i= 0:
S0=nx(`,) : v0(x) = max v1+g01 , v1+g01(x)o
=S01 S01,
6
where S01 is the open-to-buy region, and S01is the open-to-sell region:
S01 =nx(`,) : v0(x)=(v1+g01 )(x)o,
S01=nx(`,) : v0(x)=(v1+g01)(x)o.
Sell-to-close region from the long position i= 1:
S1=nx(`,) : v1(x) = (v0+g10 )(x)o.
Buy-to-close region from the short position i=1:
S1=nx(`,) : v1(x) = (v0+g10 )(x)o,
and the continuation regions, defined as the complement sets of the switching regions:
C0= (`,)\ S0=nx(`,) : v0(x)>max v1+g01 , v1+g01(x)o,
C1= (`,)\ S1=nx(`,) : v1(x)>(v0+g10 )(x)o,
C1= (`,)\ S1=nx(`,) : v1(x)>(v0+g10 )(x)o.
4 Solution
In this section, we focus on the existence and structure of switching regions, and then we
use the results on smooth fit property, uniqueness result for viscosity solutions of the value
functions to derive the form of value functions in which the optimal cut-off points can be
obtained by solving smooth-fit condition equations.
Lemma 4.1
S01 ,µL `0
ρ+µ(`,),S01µL +`0
ρ+µ,,
S1µL `1
ρ+µ,(`,),S1 ,µL +`1
ρ+µ(`,),
where
0< `0:= λ+ρε, `1:= λρε (`0, `0).
Proof. Let ¯x S01 , so that v0(¯x)=(v1+g01 )( ¯x). By writing that v0is a viscosity
supersolution to: ρv0 Lv00, we then get
ρ(v1+g01 )(¯x) L(v1+g01 )(¯x)0.(4.1)
Now, since g01 +g10 =2ε < 0, this implies that S01 S1=, so that ¯x C1. Since v1
satisfies the equation ρv1 Lv1+λ= 0 on C1, we then have from (4.1)
ρg01 x) Lg01 (¯x)λ0.
7
Recalling the expressions of g01 and L, we thus obtain: ρx+ε)µ¯xλ+ 0, which
proves the inclusion result for S01. Similar arguments show that if ¯x S01then
ρg01x) Lg01(¯x)λ0,
which proves the inclusion result for S01after direct calculation.
Similarly, if ¯x S1then ¯x S01or ¯x C0: if ¯x S01, we obviously have the inclusion
result for S1. On the other hand, if ¯x C0, using the viscosity supersolution property of
v1, we have:
ρg10 x) Lg10 (¯x) + λ0,
which yields the inclusion result for S1. By the same method, we shows the inclusion result
for S1.2
We next examine some sufficient conditions under which the switching regions are not
empty.
Lemma 4.2 (1) The switching regions S1and S01are always not empty.
(2)
(i) If `=−∞, then S1is not empty
(ii) If `= 0, and ε < λ
ρ, then S16=.
(3) If `=−∞, then S01 is not empty.
Proof. (1) We argue by contradiction, and first assume that S1=. This means that once
we are in the long position, it would be never optimal to close our position. In other words,
the value function v1would be equal to ˆ
V1given by
ˆ
V1(x) = EhλZ
0
eρtdti=λ
ρ.
Since v1v0+g10 , this would imply v0(x) λ
ρ+εx, for all x(`,), which obviously
contradicts the nonnegativity of the value function v0.
Suppose now that S01=. Then, from the inclusion results for S0in Lemma 4.1, this
implies that the continuation region C0would contain at least the interval (µL`0
ρ+µ,)
(`,). In other words, we should have: ρv0 Lv0= 0 on ( µL`0
ρ+µ,)(`,), and so
v0should be in the form:
v0(x) = C+ψ+(x) + Cψ(x),x > µL `0
ρ+µ`,
for some constants C+and C. In view of the linear growth condition on v0and condition
(3.3) when xgoes to , we must have C+= 0. On the other hand, since v0v1+g01,
and recalling the lower bound on v1in Lemma 3.1, this would imply:
Cψ(x) λ
ρ+xε, x > µL `0
ρ+µ`.
8
By sending xto , and from (3.2), we get the contradiction.
(2) Suppose that S1=. Then, a similar argument as in the case S1=, would imply
that v0(x) λ
ρ+ε+x, for all x(`,). This immediately leads to a contradiction
when `=−∞ by sending xto −∞. When `= 0, and under the condition that ε < λ
ρ,
we also get a contradiction to the non negativity of v0.
(3) Consider the case when `=−∞, and let us argue by contradiction by assuming
that S01 =. Then, from the inclusion results for S0in Lemma 4.1, this implies that the
continuation region C0would contain at least the interval (−∞,µL+`0
ρ+µ). In other words, we
should have: ρv0 Lv0= 0 on (−∞,µL+`0
ρ+µ), and so v0should be in the form:
v0(x) = C+ψ+(x) + Cψ(x),x < µL +`0
ρ+µ,
for some constants C+and C. In view of the linear growth condition on v0and condition
(3.3) when xgoes to −∞, we must have C= 0. On the other hand, since v0v1+g01 ,
recalling the lower bound on v1in Lemma 3.1, this would imply:
C+ψ+(x) λ
ρ(x+ε),x < µL +`0
ρ+µ.
By sending xto −∞, and from (3.2), we get the contradiction. 2
Remark 4.1 Lemma 4.2 shows that S1is non empty. Furthermore, notice that in the case
where `= 0, S1can be equal to the whole domain (0,), i.e. it is never optimal to stay
in the long position regime. Actually, from Lemma 4.1, such extreme case may occur only
if µL `10, in which case, we would also get µL `0<0, and thus S01 =. In that
case, we are reduced to a problem with only two regimes i= 0 and i=1. 2
The above Lemma 4.2 left open the question whether S1is empty when `= 0 and ε
λ
ρ, and whether S01 is empty or not when `= 0. The next Lemma (and Remark 4.2)
provides sufficient condition under which these sets are not empty in the case of IGBM
process.
Lemma 4.3 Let Xbe governed by the Inhomogeneous Geometric Brownian motion in
(3.5), and set
K0(y) := ( c
y)a1
U(a, b, c
y)(yε+λ
ρ)(λ
ρ+ε), y > 0,
K1(y) := ( c
y)a1
U(a, b, c
y)(yελ
ρ)+(λ
ρε), y > 0,
where a,band care defined in (3.7). If there exists y(0,µL+`0
ρ+µ)(resp y > 0) such that
K0(y)(resp. K1)>0, then S01 (resp. S1) is not empty.
Proof. See Appendix. 2
9
Remark 4.2 The above Lemma 4.3 gives a sufficient condition in terms of the function
K0and K1, which ensures that S01 and S1are not empty when the spread is an IGBM
process. Let us discuss how it is satisfied. From the asymptotic property of the confluent
hypergeometric functions, we have: limz→∞ zaU(a, b, z ) = 1. Then by sending Lto infinity
(recall that c=2µL
σ2), and from the expression of K0and K1in Lemma 4.3, we have:
lim
L→∞ K0(y) = lim
c→∞ K0(y) = y2ε= lim
L→∞ K1(y).
This implies that for Llarge enough, one can choose 2ε < y < µL+`0
ρ+µso that K0(y)
>0. Notice also that K0is nondecreasing with Las a consequence of the fact that
∂z zaU(a, b, z) = aU (a+1,b,z )(ab+1)
z<0. In practice, one can check by numerical method
the condition K0(y)>0 for 0 < y < µL+`0
ρ+µ. For example, with µ= 0.8, σ= 0.5 , ρ= 0.1,
λ= 0.07, ε= 0.005, and L= 3, we have µL+`0
ρ+µ= 2.7450, and K0(1) = 0.9072 >0.
Similarly, for Llarge enough, one can find y > 2εsuch that K1(y)>0 ensuring that S1
is not empty. 2
We are now able to describe the complete structure of the switching regions.
Proposition 4.1 1) There exist finite cutoff levels ¯x01 ,¯x01,¯x1,¯x1such that
S1= x1,)(`,),S01= [¯x01,),
S1= (`,¯x1],S01 = (`,¯x01],
and satisfying ¯x01µL+`0
ρ+µ,¯x1µL`1
ρ+µ,¯x1µL+`1
ρ+µ,¯x01 µL`0
ρ+µ. Moreover, ¯x01
<¯x1, i.e. S01 S1=and ¯x01>¯x1, i.e. S01 S1=.
2) We have ¯x1¯x01, and ¯x01 ¯x1i.e. the following inclusions hold:
S01 S1,S01 S1.
Proof. 1) (i) We focus on the structure of the sets S01 and S1, and consider first the case
where they are not empty. Let us then set ¯x01 = sup S01 , which is finite since S01 is not
empty, and is included in (`,µL`0
ρ+µ] by Lemma 4.1. Moreover, since S01is included in
[µL+`0
ρ+µ,), it does not intersect with (`,¯x01 ), and so v0(x)>(v1+g01)(x) for x <
¯x01, i.e. (`,¯x01 ) S01 C0. From (3.9), we deduce that v0is a viscosity solution to
min ρv0 Lv0, v0v1g01 = 0,on (`,¯x01).(4.2)
Let us now prove that S01 = (`,¯x01]. To this end, we consider the function w0=v1+g01
on (`,¯x01]. Let us check that w0is a viscosity supersolution to
ρw0 Lw00 on (`,¯x01 ).(4.3)
For this, take some point ¯x(`,¯x01 ), and some smooth test function ϕsuch that ¯xis
a local minimum of w0ϕ. Then, ¯xis a local minimum of v1(ϕg01 ) by definition of
10
w0. By writing the viscosity supersolution property of v1to: ρv1Lv1+λ0, at ¯xwith
the test function ϕg01 , we get:
0ρ(ϕg01 )(¯x) L(ϕg01 )(¯x) + λ
=ρϕx) Lϕ(¯x)+(ρ+µ)(¯x+`0µL
ρ+µ)
ρϕx) Lϕ(¯x),
since ¯x < ¯x01 µL`0
ρ+µ. This proves the viscosity supersolution property (4.3), and
actually, by recalling that w0=v1+g01,w0is a viscosity solution to
min ρw0 Lw0, w0v1g01 = 0,on (`,¯x01).(4.4)
Moreover, since ¯x01 lies in the closed set S01, we have w0(¯x01) = (v1+g01 )(¯x01 )
=v0(¯x01). By uniqueness of viscosity solutions to (4.2), we deduce that v0=w0on
(`,¯x01], i.e. S01 = (`,¯x01]. In the case where S01 is empty, which may arise only
when `= 0 (recall Lemma 4.2), then it can still be written in the above form (`,¯x01 ]
by choosing ¯x01 `(µL`0
ρ+µ).
By similar arguments, we show that when S1is not empty, it should be in the form:
S1= (`,¯x1], for some ¯x1µL+`1
ρ+µ, while when it is empty, which may arise only
when `= 0 (recall Lemma 4.2), it can be written also in this form by choosing ¯x10
(µL+`1
ρ+µ).
(ii) We derive similarly the structure of S01and S1which are already known to be non
empty (recall Lemma 4.2): we set ¯x01= inf S01, which lies in [ µL+`0
ρ+µ,) since S01is
included in [µL+`0
ρ+µ,) by Lemma 4.1. Then, we observe that v0is a viscosity solution to
min ρv0 Lv0, v0v1g01= 0,on x01,).(4.5)
By considering the function ˜w0=v1+g01, we show by the same arguments as in (4.4)
that ˜w0is also a viscosity solution to (4.5) with boundary condition ˜w0x01) = v0(¯x01).
We conclude by uniqueness that ˜w0=v0on [¯x01,), i.e. S01= [¯x01,). The same
arguments show that S1is in the form stated in the Proposition.
Moreover, from Lemma 4.1 we have : ¯x01µL+`0
ρ+µ>µL+`1
ρ+µ ¯x1and ¯x1µL`1
ρ+µ
>µL`0
ρ+µ ¯x01 .
2) We only consider the case where ¯x1<¯x1, since the inclusion result in this proposition
is obviously obtained when ¯x1¯x1from the above forms of the switching regions. Let
us introduce the function U(x)=2v0(x)(v1+v1)(x) on [¯x1,¯x1]. On (¯x1,¯x1), we
see that v1and v1are smooth C2, and satisfy:
ρv1 Lv1+λ= 0, ρv1 Lv1+λ= 0,
which combined with the viscosity supersolution property of v0, gives
ρU LU= 2(ρv0 Lv0)+2λ0 on (¯x1,¯x1).
11
At x= ¯x1we have v1(x) = v0(x) + xεand v0(x)v1(x) + xεso that 2v0(x)
v1(x) + v1(x), which means Ux1)0. By the same way, at x=¯x1we also have
2v0(x)v1(x) + v1(x), which means U(¯x1)0. By the comparison principle, we
deduce that
2v0(x)v1(x) + v1(x) on [¯x1,¯x1].
Let us assume on the contrary that ¯x1>¯x01. We have v0x01) = v1( ¯x01) + ¯x01ε
and v1x01)> v0(¯x01) + ¯x01ε, so that (v1+v1)( ¯x01)>2v0(¯x01), leading to a
contradiction. By the same argument, it is impossible to have ¯x1<¯x01 , which ends
the proof. 2
Remark 4.3 Consider the situation where `= 0. We distinguish the following cases:
(i) λ > ρε. Then, we know from Lemma 4.2 that S16=. Moreover, for Lsmall enough,
namely L`0, we see from Proposition 4.1 that ¯x01 0 and thus S01 =.
(ii) λρε. Then `10, and for Lsmall enough namely, L `1, we see from
Proposition 4.1 that ¯x10, and thus S1=and S01 =.
2
The next result shows a symmetry property on the switching regions and value functions.
Proposition 4.2 (Symmetry property) In the case `=−∞, and if σ(x)is an even
function and L= 0, then ¯x01= ¯x01 ,¯x1= ¯x1and
vi(x) = vi(x), x R, i {0,1,1}.
Proof. Consider the process Yx
t=Xx
t, which follows the dynamics:
dYt=µYtdt +σ(Yt)d¯
Wt,
where ¯
W=Wis still a Brownian motion on the same probability measure and filtration
of W, and we can see that Yx
t=Xx
t. We consider the same optimal problem, but we use
Ytinstead of Xt, we denote
JY(x, α) = EhX
n1
eρτng(Yx
τn, ατ
n, ατn)λZ
0
eρt|αt|dti,
For i= 0,1,1, let vY
idenote the value functions with initial positions iwhen maximizing
over switching trading strategies the gain functional, that is
vY
i(x) = sup
α∈Ai
JY(x, α), x R, i = 0,1,1.
12
For any α Ai, we see that g(Yx
τn,ατ
n,ατn) = g(Xx
τn, ατ
n, ατn), and so JY(x, α) =
J(x, α). Thus, vY
i(x)JY(x, α) = J(x, α), and since αis arbitrary in Ai, we get: vY
i(x)
vi(x). By the same argument, we have vi(x)vY
i(x), and so vY
i=vi,i {0,1,1}.
Moreover, recalling that Yx
t=Xx
t, we have:
vi(x) = vY
i(x) = vi(x), x R, i {0,1,1}.
In particular, we v1(¯x1) = v1(¯x1) = (v0+g10 )(¯x1)=(v0+g10 )(¯x1), so that ¯x1 S1.
Moreover, since ¯x1= inf S1, we notice that for all r > 0, ¯x1r6∈ S1. Thus, v1(¯x1+r)
=v1x1r)>(v0+g10 )(¯x1r) = (v0+g10 )(¯x1+r), for all r > 0, which means that
¯x1= sup S1. Recalling that sup S1=¯x1, this shows that ¯x1= ¯x1. By the same
argument, we have ¯x01= ¯x01 .2
To sum up the above results, we have the following possible cases for the structure of
the switching regions:
(1) `=−∞. In this case, the four switching regions S1,S1,S01 and S01are not
empty in the form
S1= x1,),S01= [¯x01,),
S1= (−∞,¯x1],S01 = (−∞,¯x01 ],
and are plotted in Figure 1. Moreover, when L= 0 and σis an even function, S1=
−S1and S01 =−S01.
(2) `= 0. In this case, the switching regions S1and S01are not empty, in the form
S1= x1,)(0,),S01= [¯x01,),
for some ¯x1R, and ¯x01>0 by Proposition 4.1. However, S1and S01 may
be empty or not. More precisely, for the case of IGBM process, we have the three
following possibilities:
(i) S1and S01 are not empty in the form:
S1= (0,¯x1],S01 = (0,¯x01 ],
for some 0 <¯x01 ¯x1by Proposition 4.1. Such cases arises for example
when Xis the IGBM (3.5) and for Llarge enough, as showed in Lemma 4.3 and
Remark 4.2. The visualization of this case is the same as Figure 1.
(ii) S1is not empty in the form: S1= (0,¯x1] for some ¯x1<0 by Proposition
4.1, and S01 =. Such case arises when λ > ρε, and for L(λ+ρε), see
Remark 4.3(i). This is plotted in Figure 2.
(iii) Both S1and S01 are empty. Such case arises when λρε, and for L
(ρελ), see Remark 4.3(ii). This is plotted in Figure 3. Moreover, notice that
in such case, we must have λρε by Lemma 4.2(2)(ii), and so by Proposition
4.1, ¯x1µL`1
ρ+µ>0, i.e. S1= [ ¯x1,).
13
Figure 1: Regimes switching regions in cases (1) and (2)(i).
Figure 2: Regimes switching regions in case (2)(ii).
Figure 3: Regimes switching regions in case (2)(iii).
The next result provides the explicit solution to the optimal switching problem.
14
Theorem 4.1 Case (1): `=−∞. The value functions are given by
v0(x) =
A1ψ+(x)λ
ρ+g01 (x), x ¯x01 ,
A0ψ+(x) + B0ψ(x),¯x01 <x<¯x01,
B1ψ(x)λ
ρ+g01(x), x ¯x01,
v1(x) = (A1ψ+(x)λ
ρ, x < ¯x1,
v0(x) + g10 (x), x ¯x1,
v1(x) = (v0(x) + g10 (x), x ¯x1,
B1ψ(x)λ
ρ, x > ¯x1,
and the constants A0,B0,A1,B1,¯x01 ,¯x01,¯x1,¯x1are determined by the smooth-fit
conditions:
A1ψ+(¯x01 )λ
ρ+g01 (¯x01 ) = A0ψ+(¯x01 ) + B0ψ(¯x01 )
A1ψ0
+(¯x01 )1 = A0ψ0
+(¯x01 ) + B0ψ0
(¯x01 )
B1ψx01)λ
ρ+g01x01) = A0ψ+(¯x01) + B0ψx01)
B1ψ0
x01) + 1 = A0ψ0
+x01) + B0ψ0
x01)
A1ψ+x1)λ
ρ=A0ψ+x1) + B0ψ(¯x1) + g10 x1)
A1ψ0
+x1) = A0ψ0
+x1) + B0ψ0
x1)+1
B1ψ(¯x1)λ
ρ=A0ψ+(¯x1) + B0ψ(¯x1) + g10 (¯x1)
B1ψ0
(¯x1) = A0ψ0
+(¯x1) + B0ψ0
(¯x1)1.
Case (2)(i): `= 0, and both S1and S01 are not empty. The value functions have the
same form as Case (1) with the state space domain (0,).
Case (2)(ii): `= 0,S1is not empty, and S01 =. The value functions are given by
v0(x) = (A0ψ+(x),0<x<¯x01,
B1ψ(x)λ
ρ+g01(x), x ¯x01,
v1(x) = (v0(x) + g10 (x),0< x ¯x1,
B1ψ(x)λ
ρ, x > ¯x1,
v1(x) = (A1ψ+(x)λ
ρ,0<x<max(¯x1,0),
v0(x) + g10 (x), x max(¯x1,0),
15
and the constants A0,A1,B1,¯x01>0,¯x1,¯x1<0are determined by the smooth-fit
conditions:
B1ψx01)λ
ρ+g01x01) = A0ψ+(¯x01)
B1ψ0
x01) + 1 = A0ψ0
+x01)
A1ψ+x1)λ
ρ=A0ψ+x1) + g10 (¯x1)
A1ψ0
+x1) = A0ψ0
+x1)+1
B1ψ(¯x1)λ
ρ=A0ψ+(¯x1) + g10 (¯x1)
B1ψ0
(¯x1) = A0ψ0
+(¯x1)1.
Case (2)(iii): `= 0, and S1=S01 =. The value functions are given by
v0(x) = (A0ψ+(x),0<x<¯x01,
λ
ρ+g01(x), x ¯x01,
v1(x) = (A1ψ+(x)λ
ρ, x < ¯x1,
v0(x) + g10 (x), x ¯x1,
v1=λ
ρ,
and the constants A0,A1,¯x01>0,¯x1>0, are determined by the smooth-fit conditions:
λ
ρ+g01x01) = A0ψ+(¯x01)
1 = A0ψ0
+x01)
A1ψ+x1)λ
ρ=A0ψ+x1) + g10 (¯x1)
A1ψ0
+x1) = A0ψ0
+x1)+1.
Proof. We consider only case (1) and (2)(i) since the other cases are dealt with by similar
arguments. We have S01 = (`,¯x01 ], which means that v0=v1+g01 on (`,¯x01 ].
Moreover, v1is solution to ρv1Lv1+λ= 0 on (`,¯x1), which combined with the bound
in the Lemma 3.1, shows that v1should be in the form: v1=A1ψ+λ
ρon (`,¯x1). Since
¯x01 <¯x1, we deduce that v0has the form expressed as: A1ψ+λ
ρ+g01 on (`,¯x01 ].
In the same way, v1should have the form expressed as B1ψλ
ρon (¯x1,) and v0
has the form expressed as B1ψλ
ρ+g01on x01,). We know that v0is solution
to ρv0 Lv0= 0 on (¯x01 ,¯x01) so that v0should be in the form: v0=A0ψ++B0ψ
on (¯x01 ,¯x01). We have S1= x1,), which means that v1=v0+g10 on x1,) and
S1= (`,¯x1], which means that v1=v0+g10 on (`,¯x1]. From Proposition
4.1 we know that ¯x1¯x01, and ¯x01 ¯x1and by the smooth-fit property of value
function we obtain the above smooth-fit condition equations in which we can compute the
cut-off points by solving these quasi-algebraic equations. 2
16
Remark 4.4 The cases (2)(i)-(iii) of Theorem 4.1 imply that one needs to establish the
emptiness or non-emptiness of the sets S01 and S1when `= 0 before finding the cut-off
points. This issue is covered when the spread is IGBM and both of the conditions on K0
and K1of Lemma 4.3 are satisfied, in which case one can apply case (2)(i) of the previous
Theorem. However, when `= 0, we do not know which case (2) of Theorem 4.1 is relevant
when the spread is not IGBM or when it is IGBM but either K0and K1are nonpositive.
2
Remark 4.5 1. In Case (1) and Case(2)(i) of Theorem 4.1, the smooth-fit conditions
system is written as:
ψ+(¯x01 ) 0 ψ+(¯x01 )ψ(¯x01 )
0ψx01)ψ+(¯x01)ψx01)
ψ+x1) 0 ψ+x1)ψ(¯x1)
0ψ(¯x1)ψ+(¯x1)ψ(¯x1)
×
A1
B1
A0
B0
=
λρ1g01 (¯x01 )
λρ1g01x01)
λρ1+g10 x1)
λρ1+g10 (¯x1)
(4.6)
and
ψ0
+(¯x01 ) 0 ψ0
+(¯x01 )ψ0
(¯x01 )
0ψ0
x01)ψ0
+x01)ψ0
x01)
ψ0
+x1) 0 ψ0
+x1)ψ0
x1)
0ψ0
(¯x1)ψ0
+(¯x1)ψ0
(¯x1)
×
A1
B1
A0
B0
=
1
1
1
1
.(4.7)
Denote by Mx01 ,¯x01,¯x1,¯x1) and Mx(¯x01 ,¯x01,¯x1,¯x1) the matrices:
Mx01 ,¯x01,¯x1,¯x1) =
ψ+(¯x01 ) 0 0 ψ(¯x01 )
0ψx01)ψ+(¯x01) 0
ψ+x1) 0 0 ψx1)
0ψ(¯x1)ψ+(¯x1) 0
,
Mxx01 ,¯x01,¯x1,¯x1) =
ψ0
+(¯x01 ) 0 0 ψ0
(¯x01 )
0ψ0
x01)ψ0
+x01) 0
ψ0
+x1) 0 0 ψ0
x1)
0ψ0
(¯x1)ψ0
+(¯x1) 0
.
Once Mx01 ,¯x01,¯x1,¯x1) and Mx(¯x01 ,¯x01,¯x1,¯x1) are nonsingular, straightforward com-
putations from (4.6) and (4.7) lead to the following equation satisfied by ¯x01 , ¯x01, ¯x1, ¯x1:
Mxx01 ,¯x01,¯x1,¯x1)1
1
1
1
1
=Mx01 ,¯x01,¯x1,¯x1)1
λρ1g01 (¯x01 )
λρ1g01x01)
λρ1+g10 x1)
λρ1+g10 (¯x1)
.
This system can be separated into two independent systems:
"ψ0
+(¯x01 )ψ0
(¯x01 )
ψ0
+x1)ψ0
x1)#1
×"1
1#=
"ψ+(¯x01 )ψ(¯x01 )
ψ+x1)ψ(¯x1)#1
×"λρ1g01 (¯x01 )
λρ1+g10 x1)#(4.8)
17
and
"ψ0
x01)ψ0
+x01)
ψ0
(¯x1)ψ0
+(¯x1)#1
×"1
1#=
"ψx01)ψ+(¯x01)
ψ(¯x1)ψ+(¯x1)#1
×"λρ1g01x01)
λρ1+g10 (¯x1)#(4.9)
We then obtain thresholds ¯x01 , ¯x01, ¯x1, ¯x1by solving two quasi-algebraic system equa-
tions (4.8) and (4.9). Notice that for the examples of OU or IGBM process, the matrices
Mx01 ,¯x01,¯x1,¯x1) and Mx(¯x01 ,¯x01,¯x1,¯x1) are nonsingular so that their inverses are
well-defined. Indeed, we have: ψ00
+>0 and ψ00
>0. This property is trivial for the case of
OU process, while for the case of IGBM:
d2ψ+(x)
dx2=d
dx a
xa+1 (U(a+ 1, b, c
x)(ab+ 1))
=a(a+ 1)
xa+2 U(a+ 2, b, c
x)(ab+ 1)(ab+ 2) >0,x > 0.
(x)
dx =axa2(bxM(a, b, c
x) + cM(a+ 1, b + 1,c
x))
b,x > 0.
Thus, ψ0
is strictly increasing since M(a, b, c
x) is strictly decreasing, and so ψ00
>0. More-
over, we have:
det Mx01 ,¯x01,¯x1,¯x1)(4.10)
=ψx01 )ψ+(¯x1)ψx1)ψ+(¯x01 )ψ(¯x01)ψ+(¯x1)ψ(¯x1)ψ+(¯x01).
Recalling that ¯x01 <¯x1and ¯x01>¯x1(see Proposition 4.1), and since ψ+is a strictly
increasing and positive function, while ψis a strictly decreasing positive function, we
have: ψx01 )ψ+(¯x1)ψ(¯x1)ψ+( ¯x01 )>0 and ψ(¯x01)ψ+( ¯x1)ψ(¯x1)ψ+( ¯x01)<
0, which implies the non singularity of the matrix Mx01 ,¯x01,¯x1,¯x1). On the other hand,
we have:
detMxx01 ,¯x01,¯x1,¯x1)(4.11)
=ψ0
x01 )ψ0
+x1)ψ0
x1)ψ0
+x01 )ψ0
x01)ψ0
+x1)ψ0
x1)ψ0
+x01).
Since ψ0
+is a strictly increasing positive function and ψ0
is a strictly increasing function,
with ψ0
<0, we get: ψ0
x01 )ψ0
+x1)ψ0
x1)ψ0
+x01 )<0 and ψ0
x01)ψ0
+x1)
ψ0
x1)ψ0
+x01)>0, which implies the non singularity of the matrix Mxx01 ,¯x01,¯x1,¯x1).
2. In Case (2)(ii) of Theorem 4.1, we obtain the thresholds ¯x01>0, ¯x1<0 from the
smooth-fit conditions which lead to the quasi-algebraic system:
"ψ0
x01)ψ0
+x01)
ψ0
(¯x1)ψ0
+(¯x1)#1
×"1
1#=
"ψx01)ψ+(¯x01)
ψ(¯x1)ψ+(¯x1)#1
×"λρ1+g01x01)
λρ1g10 (¯x1)#.(4.12)
18
The non singularity of the matrix above is checked similarly as in case (1) and (2)(i) for
the examples of the OU or IGBM process. Note that ¯x01, ¯x1are independent from ¯x1,
which is obtained from the equation:
λρ1g10 x1)ψ0
+x1) + ψ+(¯x1) = 0.(4.13)
When ¯x10, this means that S1= (0,).
3. In Case (2)(iii) of Theorem 4.1, the threshold ¯x1>0 is obtained from the equation
(4.13), while the threshold ¯x01>0 is derived from the smooth-fit condition leading to the
quasi-algebraic equation:
λρ1g01x01)ψ0
+x01) + ψ+(¯x01)=0.(4.14)
2
5 Numerical examples
In this part, we consider OU process and IGBM as examples.
1. We first consider the example of the Ornstein-Uhlenbeck process:
dXt=µXtdt +σdWt,
with µ,σpositive constants. In this case, the two fundamental solutions to (3.1) are given
by
ψ+(x) = Z
0
tρ
µ1exp t2
2+2µ
σxtdt, ψ(x) = Z
0
tρ
µ1exp t2
22µ
σxtdt,
and satisfy assumption (3.3). We consider a numerical example with the following specifi-
cations: : µ= 0.8 , σ= 0.5 , ρ= 0.1 , λ= 0.07 , ε= 0.005 , L= 0.
Remark 5.1 We can reduce the case of non zero long run mean L6= 0 of the OU process
to the case of L= 0 by considering process Yt=XtLas spread process, because in this
case σis constant. Finally, we can see that, cutoff points translate along L, as illustrated
in figure 6. 2
We recall some notations:
S01 = (−∞,¯x01 ] is the open-to-buy region,
S01= x01,) is the open-to-sell region,
S1= x1,) is Sell-to-close region from the long position i= 1,
S1= (−∞,¯x1] is Buy-to-close region from the short position i=1.
We solve the two systems (4.8) and (4.9) which give
¯x01 = 0.2094,¯x1= 0.0483,¯x1= 0.0483,¯x01= 0.2094,
and confirm the symmetry property in Proposition 4.2.
19
0 1000 2000 3000 4000 5000 6000 7000 8000
−1
−0.5
0
0.5
1
1.5
Threshold and spread simulation
0 1000 2000 3000 4000 5000 6000 7000 8000
−1
−0.5
0
0.5
1
Strategy curve
Spread
¯x01
¯x1
¯x1
¯x01
Figure 4: Simulation of trading strategies
−5 −4 −3 −2 −1 0 1 2 3 4 5
0
2
4
6
8
10
12
Initial spread
Value
v0
v1
v−1
Figure 5: Value functions
In figure 5, we see the symmetry property of value functions as showed in Proposition
4.2. Moreover, we can see that v1is a non decreasing function while v1is non increasing.
The next figure shows the dependence of cut-off point on parameters
20
0 0.5 1
−0.5
0
0.5
µ
The impact of µ
0 0.5 1
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
σ
The impact of σ
0 0.005 0.01 0.015
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
ε
The impact of transaction fee ε
0 0.1 0.2
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
λ
The impact of penalty factor λ
0 0.5 1 1.5
0
0.2
0.4
0.6
0.8
1
1.2
1.4
L
The impact of L
¯x01
¯x1
¯x1
¯x01
Figure 6: The dependence of cut-off point on parameters
In figure 6, µmeasures the speed of mean reversion and we see that the length of
intervals S01 ,S01increases and the length of intervals S1,S1decreases as µgets bigger.
The length of intervals S01 ,S01,S1, and S1decreases as volatility σgets bigger. Lis
the long run mean, to which the process tends to revert, and we see that the cutoff points
translate along L. We now look at the parameters that does not affect on the dynamic
of spread: the length of intervals S01,S01,S1, and S1decreases as the transaction fee ε
gets bigger. Finally, the length of intervals S01,S01decreases and the length of intervals
S1,S1increases as the penalty factor λgets larger, which means that the holding time in
flat position i= 0 is longer and the opportunity to enter the flat position from the other
position is bigger as the penalty factor λis increasing.
2. We now consider the example of Inhomogeneous Geometric Brownian Motions which
has stochastic volatility, see more details in Zhao [17] :
dXt=µ(LXt)dt +σXtdWt, X0>0,
where µ,Land σare positive constants. Recall that in this case, the two fundamental
solutions to (3.1) are given by
ψ+(x) = xaU(a, b, c
x), ψ(x) = xaM(a, b, c
x),
21
where
a=pσ4+ 4(µ+ 2ρ)σ2+ 4µ2(2µ+σ2)
2σ2>0,
b=2µ
σ2+ 2a+ 2, c =2µL
σ2,
Mand Uare the confluent hypergeometric functions of the first and second kind. We can
easily check that ψis a monotone decreasing function, while
+(x)
dx =a
xa+1 (U(a+ 1, b, c
x)(ab+ 1)) >0,x > 0,
so that ψ+is a monotone increasing function. Moreover, by the asymptotic property of the
confluent hypergeometric functions (cf.[1]), the fundamental solutions ψ+and ψsatisfy
the condition (3.3).
Case (2)(i): Both S1and S01 are not empty. Let us consider a numerical example with
the following specifications: : µ= 0.8 , σ= 0.5 , ρ= 0.1 , λ= 0.07 , ε= 0.005 , and we set
L= 10. Note that, in this case the condition in Lemma 4.3 is satisfied, and we solve the
two systems (4.8) and (4.9) which give
¯x01 =8.2777,¯x1= 9.3701,¯x1=8.4283,¯x01= 9.5336.
0 5 10 15 20 25
10
20
30
40
50
60
70
Initial spread
Value
v0
v1
v−1
Figure 7: Value functions
In the figure 7, we can see that v1is non decreasing while v1is non increasing. More-
over, v1is always larger than v0, and v1.
The next figure 8 shows the dependence of cut-off points on parameters (Note that the
condition in Lemma 4.3 is satisfied for all parameters in this figure).
22
0 0.5 1
4
5
6
7
8
9
10
µ
The impact of µ
0 0.5 1
8
8.5
9
9.5
10
σ
The impact of σ
0 0.005 0.01 0.015
8
8.5
9
9.5
10
ε
The impact of transaction fee ε
0 0.05 0.1 0.15 0.2
8
8.5
9
9.5
10
λ
The impact of penalty factor λ
5 10 15
4
6
8
10
12
14
16
L
The impact of L
¯x01
¯x1
¯x1
¯x01
Figure 8: The dependence of cut-off point on parameters
We can make the same comments as in the case of the OU process, except for the
dependence with respect to the long run mean L. Actually, we see that when Lincreases,
the moving of cutoff points is no more translational due to the non constant volatility.
Case (2)(ii):S01 is empty. Let us consider a numerical example with the following
specifications: : µ= 0.8, σ= 0.3, ρ= 0.1, λ= 0.35, ε= 0.55, and L= 0.5. We solve the
two systems (4.12) and (4.13) which give
¯x1= 0.1187,¯x1=0.8349,¯x01= 2.7504.
Case (2)(iii): Both S1and S01 are empty. Let us consider a numerical example with
the following specifications: : µ= 0.8, σ= 0.3, ρ= 0.2, λ= 0.05, ε= 0.65, and L= 0.1.
The two equations (4.14) and (4.13) give
¯x1= 0.4293,¯x01= 0.9560.
23
Appendix
A Proof of Lemma 3.1
The lower bound for v0and viare trivial by considering the strategies of doing nothing.
Let us focus on the upper bound. First, by standard arguments using Itˆo’s formula and
Gronwall lemma, we have the following estimate on the diffusion X: there exists some
positive constant r, depending on the Lipschitz constant of σ, such that
E|Xx
t| Cert(1 + |x|),t0,(A.1)
E|Xx
tXy
t| ert|xy|,t0,(A.2)
for some positive constant Cdepending on ρ,Land µ. Next, for two successive trading
times τnand σn=τn+1 corresponding to a buy-and-sell or sell-and-buy strategy, we have:
Eheρτng(Xx
τn, ατ
n, ατn) + eρσng(Xx
σn, ασ
n, ασn)i(A.3)
EheρσnXx
σneρτnXx
τniEhZσn
τn
eρt(µ+ρ)|Xx
t|dti+EhZσn
τn
eρtµLdti,
where the second inequality follows from Itˆo’s formula. When investor is staying in flat
position (i= 0), in the first trading time investor can move to state i= 1 or i=1, and
in the second trading time she has to back to state i= 0. So that, the strategy when we
stay in state i= 0 can be expressed by the combination of infinite couples: buy-and-sell,
sell-and-buy, for example: states 0 10 10 1010... it means:
buy-and-sell, sell-and-buy, sell-and-buy, buy-and-sell,.... We deduce from (A.3) that for any
α A0,
J(x, α)EhZ
0
eρt(µ+ρ)|Xx
t|dti+µL
ρ.
Recalling that, when investor starts with a long or short position (i=±1) she has to close
first her position before opening a new one, so that for α A1or α A1,
J(x, α) |x|+EhZτ1
0
eρt(µ+ρ)|Xx
t|dti+EhZτ1
0
eρtµLdti
+EhZ
τ2
eρt(µ+ρ)|Xx
t|dti+EhZ
τ2
eρtµLdti
|x|+EhZ
0
eρt(µ+ρ)|Xx
t|dti+µL
ρ,
which proves the upper bound for viby using the estimate (A.1). By the same argument,
for two successive trading times τnand σn=τn+1 corresponding to a buy-and-sell or sell-
and-buy strategy, we have:
Eheρτng(Xx
τn, ατ
n, ατn) + eρσng(Xx
σn, ασ
n, ασn)
eρτng(Xy
τn, ατ
n, ατn)eρσng(Xy
σn, ασ
n, ασn)i
EheρσnXx
σneρτnXx
τneρσnXy
σn+eρτnXy
τni
EhZσn
τn
eρt(µ+ρ)|Xx
tXy
t|dti,
24
where the second inequality follows from Itˆo’s formula. We deduce that
|vi(x)vi(y)| sup
α∈Ai|J(x, α)J(y, α)|
|xy|+EhZ
0
eρt(µ+ρ)|Xx
tXy
t|dti,
which proves the Lipschitz property for vi, i = 0,1,1 by using the estimate (A.2). 2
B Proof of Lemma 4.3
Suppose that S01 =. Then, from the inclusion results for S0in Lemma 4.1, this implies
that the continuation region C0would contain at least the interval (0,µL+`0
ρ+µ). In other
words, we should have: ρv0 Lv0= 0 on (0,µL+`0
ρ+µ), and so v0should be in the form:
v0(x) = C+ψ+(x) + Cψ(x),0<x<µL +`0
ρ+µ,
for some constants C+and C. From the bounds on v0in Lemma 3.1, and (3.2), we must
have C= 0.
Next, for 0 < x y, let us consider the first passage time τx
y:= inf{t:Xx
t=y}of the
inhomogeneous Geometric Brownian motion. We know from [17] that
Exeρτx
y=x
yaU(a, b, c
x)
U(a, b, c
y)=ψ+(x)
ψ+(y).(B.4)
We denote by ¯v1(x;y) the gain functional obtained from the strategy consisting in changing
position from initial state xand regime i= 1, to the regime i= 0 at the first time Xx
thits
y(0 < x y), and then following optimal decisions once in regime i= 0:
¯v1(x;y) = E[eρτx
y(v0(y) + yε)Zτx
y
0
λeρtdt],0< x y.
Since v0(y) = C+ψ+(y), for all 0 < y < µL+`0
ρ+µ, and recalling (B.4) we have:
¯v1(x;y) = E[eρτx
y(C+ψ+(y) + yε)Zτx
y
0
λeρtdt]
=ψ+(x)
ψ+(y)(C+ψ+(y) + yε+λ
ρ)λ
ρ
=v0(x) + ψ+(x)
ψ+(y)(yε+λ
ρ)λ
ρ,0< x y < µL +`0
ρ+µ.
Now, by definition of v1, we have v1(x)¯v1(x;y), so that:
v1(x)v0(x) + ψ+(x)
ψ+(y)(yε+λ
ρ)λ
ρ,0< x y < µL +`0
ρ+µ.
By sending xto zero, and recalling (3.6) and (3.8), this yields
v1(0+)v0(0+) + K0(y) + ε, 0< y < µL +`0
ρ+µ.
25
Therefore, under the condition that there exists y(0,µL+`0
ρ+µ) such that K(y)>0, we
would get:
v1(0+)> v0(0+) + ε,
which is in contradiction with the fact that we have: v0v1+g01 , and so: v0(0+)v1(0+)
ε.
Suppose that S1=, in this case v1=λ/ρ. By the same argument as the above
case, we have
v0(x)E[eρτx
y(v1(y) + yε)] = E[eρτ x
y(λ
ρ+yε)]
=λ
ρ+yεψ+(x)
ψ+(y).
by (B.4). By sending xto zero, and recalling (3.6) and (3.8), we thus have
v0(0+) λ
ρ+ε+K1(y)y > 0.(B.5)
Therefore, under the condition that there exists y > 0 such that K1(y)>0, we would get:
v0(0+)>λ
ρ+ε,
which is in contradiction with the fact that we have: v1v0+g10 , and so: λ
ρ=v1(0+)
v0(0+)ε.2
References
[1] Abramowitz, M. and I. Stegun (1972): Handbook of mathematical functions: with formulas,
graphs, and mathematical tables, Courier Dover Publications.
[2] Borodin, A. and P. Salminen (2002): Handbook of Brownian motion: facts and formulae,
Springer.
[3] Chiu M.C. and H.Y. Wong (2011): “Mean-variance portfolio selection of cointegrated assets”,
Journal of Economic Dynamics and Control, vol. 35, 1369-1385.
[4] Chiu M.C. and H.Y. Wong (2013): “Optimal investment for insurer with cointegrated assets”,
Insurance: mathematics and economics, vol. 52, 52-64.
[5] Ehrman, D. (2006): The handbook of pairs trading: Strategies using equities, options, and
futures, vol 240, John Wiley and Sons.
[6] Elliott, R., Van der Hoek J. and W. Malcom (2005): “Pairs trading”, Quantitative Finance,
vol 5, 271-276.
[7] Gatev E.W., Goetzmann W. and K. Rouwenhorst (2006): “Performance of a relative-value
arbitrage rule”, Review of Financial Studies, vol. 19, 797-827.
[8] Kong, H. T. (2010): “Stochastic control and optimization of assets trading”, PhD thesis,
University of Georgia.
26
[9] Leung, T. and X. Li (2013): “Optimal Mean Reversion Trading with Transaction Costs and
Stop-Loss Exit”, Social Science Research Network Working Paper Series.
[10] Liu J. and A. Timmermann (2013): “Optimal convergence for trade strategies”, Review of
Financial Studies, vol. 26, 1048-1086.
[11] Mudchanatongsuk, S., Primbs J. and W. Wong (2008): “Optimal pairs trading: A stochastic
control approach”, American Control Conference, IEEE, 2008.
[12] Pham H. (2007): “On the smooth-fit property for one-dimensional optimal switching prob-
lem”, eminaire de Probabilit´es XL, 187-199, Springer.
[13] Song, Q. and Q. Zhang (2013): “An optimal pairs-trading rule”, Automatica, vol 49, 3007-
3014.
[14] Tourin A. and R. Yan (2013): “Dynamic pairs trading using the stochastic control approach”,
Journal of Economic Dynamics and Control, vol 37, 1972-1981.
[15] Vidyamurthy, G. (2004): Pairs Trading: quantitative methods and analysis, John Wiley and
Sons.
[16] Zhang, H. and Q. Zhang (2008): “Trading a mean-reverting asset: Buy low and sell high”,
Automatica, vol 44, 1511-1518.
[17] Zhao B. (2009): “Inhomogeneous geometric Brownian motions”, Available at SSRN 1429449.
[18] Lei, Y. and Xu, J (2015): “Costly arbitrage through pairs trading”, Journal of Economic
Dynamics and Control, vol 56, 1-19.
27
... [23] considered an optimal switching problem using the viscosity solution method. [20] considered the optimal switching problem with finite regimes. [26] dealt with a discontinuous coefficients problem with finite-regime switchings. ...
... Then by (20), (21), (22) and (23), we get ...
... Here the first inequality is derived by using (21), (22), (23) and u ≤ 1, and the second inequality is derived from (20). ...
... Specifically, Leung and Li (2015) studies the optimal timing to open or close the position subject to fixed transaction costs and the effect of stop-loss level under the OU process by constructing the value function directly. Zhang and Zhang (2008), Song and Zhang (2013), and Ngo and Pham (2016) studied the optimal pairs trading rule that is based on optimal switching among two (buy and sell) or three (buy, sell, and flat) regimes with a fixed commission cost for each transaction, and solve the problem by finding viscosity solutions to the associated HJB equations (quasi-variational inequalities). Lei and Xu (2015) studied the optimal pairs trading rule of entering and exiting the asset market in finite horizon with proportional transaction cost for two convergent assets. ...
... Note that, Mudchanatongsuk et al. (2008) assumed no transaction cost and considered the strategy that always shorts one stock and longs the other in equal dollar amount, i.e., p(t)dL p (t) + q(t)dM q (t) = 0 or p(t)dM p (t) + q(t)dL q (t) = 0 at time t. Lei and Xu (2015) and Ngo and Pham (2016) considered a delta-neutral strategy that always long one share of a stock and short one share of the other stock, i.e., dy p (t) = −dy q (t) = 1 or dy p (t) = −dy q (t) = −1 at time t. Here, we also consider a delta-neutral strategy that requires the total of positive and negative delta of two assets is zero, hence it suggests that the number of shares of stock P bought (or sold) at time t are same as the number of shares of stock Q sold (or bought), i.e., ...
... Equation (8) implies that dy q (t) = −dy p (t) at any time t. Comparing to Lei and Xu (2015) and Ngo and Pham (2016), we remove the constraint dy p (t) = −dy q (t) = 1 or −1 and allow y p (t) = −y q (t) to be a control variable. Using Equations (5) and (8), the dynamics of g(t) in Equation (7) can be simplified as ...
Article
Full-text available
Optimal trading strategies for pairs trading have been studied by models that try to find either optimal shares of stocks by assuming no transaction costs or optimal timing of trading fixed numbers of shares of stocks with transaction costs. To find optimal strategies that determine optimally both trade times and number of shares in a pairs trading process, we use a singular stochastic control approach to study an optimal pairs trading problem with proportional transaction costs. Assuming a cointegrated relationship for a pair of stock log-prices, we consider a portfolio optimization problem that involves dynamic trading strategies with proportional transaction costs. We show that the value function of the control problem is the unique viscosity solution of a nonlinear quasi-variational inequality, which is equivalent to a free boundary problem for the singular stochastic control value function. We then develop a discrete time dynamic programming algorithm to compute the transaction regions, and show the convergence of the discretization scheme. We illustrate our approach with numerical examples and discuss the impact of different parameters on transaction regions. We study the out-of-sample performance in an empirical study that consists of six pairs of U.S. stocks selected from different industry sectors, and demonstrate the efficiency of the optimal strategy.
... Indeed, most studies in the literature consider the two-regime case. The exceptions are Pham et al. [21] and Ngo and Pham [17], who study problems with three regimes. The setting of their optimization problem in Pham et al. [21] includes transaction costs, denoted by g ij , which are incurred by switching from regime i to j, and they impose a (strict) triangular condition, g ik < g ij + g jk , j i, k, which prevents any simultaneous multiple switches. ...
... However, Ngo and Pham [17], Suzuki [26], and Suzuki [27] formally prohibit the direct switching from i to k in their transaction rules. Thus, to switch from i to k, the strategy requires two consecutive simultaneous switches by way of j, that is, i →j → k. ...
... Therefore, we need to consider the simultaneous switching theory more rigorously, which is the subject of this paper. Ngo and Pham [17] do not treat simultaneous switchings explicitly. Their way of solving their problem is rather ad hoc and specific to the particular problem, rather than being applicable to more general problems. ...
... In this paper, we investigate the problem of optimal timing to trade a mean-reverting price process with finite deadlines to enter and subsequently exit the market. We consider a general mean-reverting diffusion with an affine drift that encapsulates a number of well-known and important models, including the Ornstein-Uhlenbeck (OU) model (Ornstein and Uhlenbeck, 1930), Cox-Ingersoll-Ross (CIR) model (Cox et al., 1985), inhomogeneous geometric Brownian motion (IGBM) model (Zhao, 2009;Ngoa and Pham, 2016), and Jacobi model (Ackerer et al., 2016). ...
Preprint
We study the optimal timing strategies for trading a mean-reverting price process with afinite deadline to enter and a separate finite deadline to exit the market. The price process is modeled by a diffusion with an affine drift that encapsulates a number of well-known models,including the Ornstein-Uhlenbeck (OU) model, Cox-Ingersoll-Ross (CIR) model, Jacobi model,and inhomogeneous geometric Brownian motion (IGBM) model.We analyze three types of trading strategies: (i) the long-short (long to open, short to close) strategy; (ii) the short-long(short to open, long to close) strategy, and (iii) the chooser strategy whereby the trader has the added flexibility to enter the market by taking either a long or short position, and subsequently close the position. For each strategy, we solve an optimal double stopping problem with sequential deadlines, and determine the optimal timing of trades. Our solution methodology utilizes the local time-space calculus of Peskir (2005) to derive nonlinear integral equations of Volterra-type that uniquely characterize the trading boundaries. Numerical implementation ofthe integral equations provides examples of the optimal trading boundaries.
... Our model is also similar to Ngo and Pham [8]. They consider the problem of determining the optimal cut-off of the pair trading rule for a three-state regime-switching model. ...
Article
Full-text available
In this paper, we study a free boundary problem, which arises from an optimal trading problem of a stock whose price is driven by unobservable market status and noise processes. The free boundary problem is a variational inequality system of three functions with a degenerate operator. We prove that all the four switching free boundaries are no-overlapping, monotonic and C∞-smooth by the approximation method. We also completely determine their relative localities and provide the optimal trading strategies for the stock trading problem.
... Then, Lei and Xu (2015) proposed a model for determining multiple entry and exit-points during a trading period, and illustrated their results with applications to dual-listed Chinese stocks. Ngo and Pham (2014) frames the pairs trading problem as a regime switching model between three regimes: flat positions, one long position on one asset and a short position on the other, and vice versa . Finally, Lintilhac (2014), Cartea and Jaimungal (2015) considered applications to portfolios of co-integrated assets, both generalizing the model in Tourin and Yan (2013). ...
Preprint
We propose a pairs trading model that incorporates a time-varying volatility of the Constant Elasticity of Variance type. Our approach is based on stochastic control techniques; given a fixed time horizon and a portfolio of two co-integrated assets, we define the trading strategies as the portfolio weights maximizing the expected power utility from terminal wealth. We compute the optimal pairs strategies by using a Finite Difference method. Finally, we illustrate our results by conducting tests on historical market data at daily frequency. The parameters are estimated by the Generalized Method of Moments.
... The application of stochastic control to pairs trading originated in the work of Mudchanatongsuk et al. (2008). Further models of timeconsistent pairs trading strategies were developed recently in Chiu andWong (2011), Tourin andYan (2013), Ngo and Pham (2014), Lei and Xu (2015), and Leung and Li (2015). Besides, Hogan et al. (2004) provided a definition of the concept of statistical arbitrage and empirical measures to test the presence of opportunities. ...
Article
Full-text available
We consider a pairs trading stochastic control problem with transaction costs and constraints on the gross market exposure, and propose a new monotone Finite Difference scheme approximating the viscosity solution of the Hamilton–Jacobi–Bellman equation characterizing the optimal trading strategies. Given a fixed time horizon and a portfolio of two cointegrated assets, the agent trades the spread between the two assets and the trading strategy is defined as the possibly negative portfolio weight maximizing the expected exponential utility derived from terminal wealth. Furthermore, trades incur transaction costs comprised of explicit transactions fees and commissions and the implicit cost due to slippage. These costs are modeled as a linear or square root function of the trading rate and respectively added or subtracted from the observable asset price at the time when a buy or a sell order enters the market. Our main contribution is the derivation of a robust approximation for the nonlinear transaction cost term in the Hamilton–Jacobi–Bellman equation. Finally, we combine our monotone Finite Difference scheme with a Monte Carlo sampling method to analyze the effects of transaction fees and slippage on the trading policies’ performance.
Chapter
This chapter introduces the analytical theory of variational inequalities (VIs) and considers issues such as the existence of solutions or their qualitative properties (comparison principles, regularity of solutions, etc.), with a view towards applicability to realistic problems in management science and finance and importantly with a view towards setting the scene for the development of numerical methods for the approximation of VIs. Both the finite and infinite dimensional cases are covered in a self-contained fashion, while all results are applied in detailed fully worked examples to the three model problems introduced in the previous chapter, which are revisited using these results. All analytical tools are introduced essentially from scratch, at a level suitable for the intended audience and are motivated and illustrated with examples.
Article
Continuous-time pairs trading rules are often developed based on the diffusion limit of the first-order vector autoregressive (VAR(1)) cointegration models. Empirical identification of cointegration effects is generally made according to discrete-time error correction representation of vector autoregressive (VAR(p)) processes, allowing for delayed adjustment of the price deviation. Motivated by this, we investigate the continuous-time dynamic pairs trading problem under a class of path-dependent models. Under certain regular conditions, we prove the existence of the optimal strategy and show that it is related to a system of Riccati partial differential equations. The proof is developed by the means of functional Ito’s calculus. We conduct a numerical study to analyze the sensitivities of the pairs trading strategy with respect to the initial market conditions and the memory length.
Article
Full-text available
This paper is concerned with a pairs trading rule. The idea is to monitor two historically correlated securities. When divergence is underway, i.e., one stock moves up while the other moves down, a pairs trade is entered which consists of a pair to short the outperforming stock and to long the underperforming one. Such a strategy bets the "spread" between the two would eventually converge. In this paper, a difference of the pair is governed by a mean-reverting model. The objective is to trade the pair so as to maximize an overall return. A fixed commission cost is charged with each transaction. In addition, a stop-loss limit is imposed as a state constraint. The associated HJB equations (quasi-variational inequalities) are used to characterize the value functions. It is shown that the solution to the optimal stopping problem can be obtained by solving a number of quasi-algebraic equations. We provide a set of sufficient conditions in terms of a verification theorem. Numerical examples are reported to demonstrate the results.
Article
Full-text available
We study an equity-based pairs trading strategy. We find that 85% of the total variation in firm-level pairwise correlations of stock returns is not explained by a comprehensive list of variables that may affect return correlations. We also find that stocks with high return correlations in the past have similar returns in the future, even after controlling for the possible determinants of the return correlation. A trading strategy based on stocks that deviate from their pairs generates significant abnormal returns. We show that this return is not driven purely by the short-term reversal phenomenon. Consistent with the adaptive market efficiency theory, the return to this simple pairs trading strategy has diminished over time.. 1 Which company's stock returns comove the most with those of Whirlpool Corporation? We identify the 50 stocks that have the highest return correlations with Whirlpool between 2001 and 2005. Maytag ranks only 22 nd among these 50 firms. Number one on the list is Continental Airlines, No. 4 Alaska Airlines, No.9 American Airlines, and No.25 Mesa Air group. After some research, we find the hidden link: Whirlpool owns a financial subsidiary that leases aircrafts. In January 2006, the portfolio of the 50 stocks returned 3.6% while Whirlpool lost 3.7%. However, in February 2006, Whirlpool more than made it up by returning a whopping 11.3% while the 50-stock portfolio stayed flat at 0.7%.
Article
Motivated by the industry practice of pairs trading, we study the optimal timing strategies for trading a mean-reverting price spread. An optimal double stopping problem is formulated to analyze the timing to start and subsequently liquidate the position subject to transaction costs. Modeling the price spread by an Ornstein-Uhlenbeck process, we apply a probabilistic methodology and rigorously derive the optimal price intervals for market entry and exit. As an extension, we incorporate a stop-loss constraint to limit the maximum loss. We show that the entry region is characterized by a bounded price interval that lies strictly above the stop-loss level. As for the exit timing, a higher stop-loss level always implies a lower optimal take-profit level. Both analytical and numerical results are provided to illustrate the dependence of timing strategies on model parameters such as transaction costs and stop-loss level.
Article
We study the optimal trading policy of an arbitrageur who can exploit temporary mispricing in a market with two convergent assets. We build on the model of Liu and Timmermann (2013) and include transaction costs, which impose additional limits to the implementation of such convergence trade strategy. We show that the presence of transaction costs could reveal an endogenous stop-loss concern in a certain economy, which affects the optimal policy of the arbitrageur in significant ways. Using pairs of Chinese stock shares that are dual-listed in Shanghai and in Hong Kong, we show that our co-integration strategy is generally superior to the relative-value pairs trading strategy studied in Gatev et al. (2006). Several extensions of our model are also discussed.
Article
Convergence trades exploit temporary mispricing by simultaneously buying relatively underpriced assets and selling short relatively overpriced assets. This paper studies optimal convergence trades under both recurring and nonrecurring arbitrage opportunities represented by continuing and “stopped” cointegrated price processes and considers both fixed and stochastic (Poisson) horizons. Conventional long-short delta neutral strategies are generally suboptimal and it can be optimal to simultaneously go long (or short) in two mispriced assets. Optimal portfolio holdings critically depend on whether the risky asset position is liquidated when prices converge. Our theoretical results are illustrated on pairs of Chinese bank shares traded on both the Hong Kong and China stock exchanges.
Article
This paper considers the optimal investment problem for an insurer that invests in cointegrated assets subject to the random payments of insurance claims. The insurer’s objective is to maximize the expected utility of the terminal wealth subject to the cointegration dynamics of risky assets and the risk of paying out random liabilities with a compound Poisson process. We solve the continuous-time investment problems for the class of the constant relative risk averse utility function using the framework of the HJB equation. An explicit solution is derived by recognizing an exponential affine form in the derivation process. We then investigate the risk-preference of insurers toward statistical arbitrage from pairs-trading using the analytical results. Although a financial market with cointegrated risky assets implies the existence of statistical arbitrage opportunities, insurers may not be interested in those opportunities due to the social responsibility of a high level of risk aversion. However, if insurers are forced to trade cointegrated assets, the derived optimal solution enhances the investment performance.
Article
We propose a model for analyzing dynamic pairs trading strategies using the stochastic control approach. The model is explored in an optimal portfolio setting, where the portfolio consists of a bank account and two co-integrated stocks and the objective is to maximize for a fixed time horizon, the expected terminal utility of wealth. For the exponential utility function, we reduce the problem to a linear parabolic Partial Differential Equation which can be solved in closed form. In particular, we exhibit the optimal positions in the two stocks.
Article
Pairs trading is a common strategy used by hedge funds. It consists of a long position in one asset and a short position in another. When the spread between the two assets is observed to be out of equi- librium, a long position is taken in the underpriced asset and a short position in the overpriced one. If the spread narrows, both positions are closed, thus generating a prot. If, on the other hand, the spread increases, the position is liquidated at a pre-determined stop-loss level. We study when to optimally liquidate a pairs trading strategy when the dierence between the two assets is modeled by an Ornstein-Uhlenbeck process. We also provide a sensitivity analysis in the model parameters.
Book
I: Theory.- I. Stochastic processes in general.- II. Linear diffusions.- III. Stochastic calculus.- IV. Brownian motion.- V. Local time as a Markov process.- VI. Differential systems associated to Brownian motion.- Appendix 1. Briefly on some diffusions.- II: TABLES OF DISTRIBUTIONS OF FUNCTIONALS OF BROWNIAN MOTION AND RELATED PROCESSES.- 1. Brownian motion.- 2. Brownian motion with drift.- 3. Reflecting Brownian motion.- 4. Bessel process of order ?.- 5. Bessel process of order 1/2.- 6. Bessel process of order zero.- 7. Ornstein-Uhlenbeck process.- 8. Radial Ornstein-Uhlenbeck process.- 9. Geometric Brownian motion.- Appendix 2. Special functions.- Appendix 3. Inverse Laplace transforms.- Appendix 4. Differential equations and their solutions.- Appendix 5. Formulae for n-fold differentiation.