Optimal Trading Strategies Under Arbitrage
Optimal Trading Strategies Under Arbitrage
i
(t, S(t))dt +
K
k=1
i,k
(t, S(t))dW
k
(t)
_
(2.1)
for all i = 1, . . . , d and t [0, T] starting at S(0) R
d
+
and a money market B().
Here : [0, T] R
d
+
R
d
denotes the mean rate of return and : [0, T] R
d
+
R
dK
denotes the volatility. We assume that both functions are measurable.
For the sake of convenience we only consider discounted (forward) prices
and set the interest rate constant to zero; that is, B() 1. The ow of in-
formation is modeled as a right-continuous ltration F = T(t)
0tT
such that
W() = (W
1
(), . . . , W
K
())
T
is a K-dimensional Brownian motion with indepen-
dent components. In Section 2.5, we impose more conditions on the ltration F and
Chapter 2. The Markovian Case 10
the underlying probability space . For the moment, we assume that all stochastic
integrals that appear are measurable with respect to the ltration F. The under-
lying measure and its expectation shall be denoted by P and E, respectively. The
current state of the market S(0) should be clear from the context and so we shall
omit specifying S(0) as an index for measures and expectations in most cases.
We only consider those mean rates of return (, ) and volatilities (, ) that
imply the stock prices S
1
(), , S
d
() exist and are unique and strictly positive.
More precisely, denoting the covariance process of the stocks in the market by
a(, ) (, )
T
(, ), that is,
a
i,j
(t, S(t)) :=
K
k=1
i,k
(t, S(t))
j,k
(t, S(t))
for all i, j = 1, . . . , d and t [0, T], we impose the almost sure integrability condition
d
i=1
_
T
0
([
i
(t, S(t))[ + a
i,i
(t, S(t))) dt < .
Under this condition, the stock prices S
1
(), . . . , S
d
() can be expressed as
S
i
(t) =S
i
(0) exp
__
t
0
_
i
(u, S(u))
1
2
a
i,i
(u, S(u))
_
du+ (2.2)
+
K
k=1
_
t
0
i,k
(u, S(u))dW
k
(u)
_
> 0
for all i = 1, . . . , d and t [0, T]. Furthermore, we assume the existence of a market
price of risk, which generalizes the concept of the Sharpe ratio to several dimensions
by setting the risk factors W
k
() in relation to the mean rates of return
i
(, ).
Denition 1 (Market price of risk). A market price of risk is a progressively mea-
surable process (), which maps the volatility structure (, ) onto the mean rate
of return (, ). That is,
(t, S(t)) = (t, S(t))(t) (2.3)
holds almost surely for all t [0, T].
Chapter 2. The Markovian Case 11
Furthermore, we assume that () is square-integrable, to wit,
_
T
0
|(t)|
2
dt < (2.4)
almost surely. The existence of a market price of risk is a central assumption in
both the Benchmark Approach (see Chapter 10 of Platen and Heath 2006) and in
Stochastic Portfolio Theory (see Section 6 of Fernholz and Karatzas 2009). This
assumption enables us to discuss hedging prices, as we do throughout this thesis.
Similar assumptions have been discussed in the economic literature. For example,
in the terminology of Loewenstein and Willard (2000a), the existence of a square-
integrable market price of risk excludes cheap thrills but not necessarily free
snacks. Theorem 2 of Loewenstein and Willard (2000a) shows that a market with
a square-integrable market price of risk is consistent with an equilibrium where
agents prefer more to less. We discuss the connection between a market price
of risk and its square-integrability with various no-arbitrage notions in Remark 1
below.
Based on the market price of risk, we can now dene the stochastic discount
factor as
Z
(t) := exp
_
_
t
0
T
(u)dW(u)
1
2
_
t
0
|(u)|
2
du
_
(2.5)
with dynamics
dZ
(t) =
T
(t)Z
(t)dW(t) (2.6)
for all t [0, T]. In classical no-arbitrage theory, Z
() be a true martingale, for example by Wong and Heyde (2004), Hulley and
Platen (2009), Mijatovic and Urusov (2009), and the literature therein.
A market price of risk () does not have to be uniquely determined. Unique-
ness is intrinsically connected to completeness, as we shall see in Chapter 3, and we
need not assume it. In general, innitely many market prices of risk may exist. To
illustrate, think of a model with d = 1, K = 2, (, ) 0 and (, ) (1, 1). Then,
for any y R, the constant process () (y, y)
T
is a square-integrable market
price of risk. Another example of this non-uniqueness follows the next proposition.
We observe that the existence of a square-integrable market price of risk
implies the existence of a Markovian square-integrable market price of risk. To see
this, we dene (, ) :=
T
(, )((, )
T
(, ))
(t) := E
_
Z
(T)
Z
(t)
M
T(t)
_
and M
(t) := E
_
Z
(T)
Z
(t)
M
T(t)
_
for t [0, T], where we take the right-continuous modication
3
for each process,
we have M
() M
() and Z
() are
T
S
(T)-measurable, then Z
(T) Z
() and M
(t) M
n
:= T inf
_
t [0, T] :
_
t
0
c
2
(s)ds n
_
,
where n N, we set c
n
() := c()1
n
and observe that
Z
(T)
Z
(t)
=
Z
c
(T)
Z
c
(t)
exp
_
_
T
t
T
(u, S(u))(dW(u) + c(u)du)
1
2
_
T
t
|(u, S(u))|
2
du
_
= lim
n
Z
c
n
(T)
Z
c
n
(t)
exp
_
_
T
t
T
(u, S(u))(dW(u) + c
n
(u)du)
1
2
_
T
t
|(u, S(u))|
2
du
_
with Z
c
() and Z
c
n
() dened as in (2.5). The limit holds almost surely since both
v() and (, ) are square-integrable, which again yields the square-integrability of
3
See Theorem 1.3.13 of Karatzas and Shreve (1991).
Chapter 2. The Markovian Case 14
c(). Since
_
T
0
c
2
n
(t)dt n, Novikovs Condition (see Proposition 3.5.12 of Karatzas
and Shreve 1991) yields that Z
c
n
() is a martingale. Now, Fatous lemma, Girsanovs
theorem and Bayes rule (see Chapter 3.5 of Karatzas and Shreve 1991) yield
M
(t) liminf
n
E
Q
n
_
exp
_
_
T
t
T
(u, S(u))dW
n
(u)
1
2
_
T
t
|(u, S(u))|
2
du
_
M
T(t)
_
,
(2.7)
where dQ
n
() := Z
c
n
(T)dP() is a probability measure, E
Q
n
its expectation operator,
and W
n
() := W() +
_
0
c
n
(u)du a K-dimensional Q
n
-Brownian motion. Since
(, S())c
n
() 0 we can replace W() by W
n
() in (2.1). This yields that the
process S() has the same dynamics under Q
n
as under P. Furthermore, both
(, S()) and M have, as functionals of S(), the same distribution under Q
n
as
under P. Therefore, we can replace the expectation operator E
Q
n
by E in (2.7) and
obtain the rst part of the statement. The last inequality of the statement follows
from setting M = 1
Z
(T)>Z
(T)
and observing that M must equal zero almost
surely.
We remark that the inequality M
() M
()
2
(), which is a
strict local martingale (see Exercise 3.3.36 of Karatzas and Shreve 1991), and thus
M
(0) = E[Z
(T)] = M
(0).
Chapter 2. The Markovian Case 15
Under the assumption that an equivalent local martingale measure exists,
Theorem 12 of Jacka (1992), Theorem 3.2 of Ansel and Stricker (1993) or Theo-
rem 16 Delbaen and Schachermayer (1995c) show that a contingent claim can be
hedged if and only if the supremum over all expectations of the terminal value of
the contingent claim under all equivalent local martingale measures is a maximum.
In our setup, we also observe that the supremum over all M
(0) in the last propo-
sition is a maximum, attained by any Markovian market price of risk. Indeed, we
shall prove in Theorem 2 that, under weak analytic assumptions, claims of the form
M = p(S(T)) can be hedged. The general theory lets us conjecture that all claims
measurable with respect to T
S
(T) can be hedged. Theorem 6 of Chapter 3 conrms
this conjecture.
As pointed out by Ioannis Karatzas in a personal communication (2010),
Proposition 1 might be related to the Markovian selection results, as in Krylov
(1973), Section 4.5 of Ethier and Kurtz (1986), and Chapter 12 of Stroock and
Varadhan (2006). There, the existence of a Markovian solution for a martingale
problem is studied. It is observed that a supremum over a set of expectations
indexed by a family of distributions is attained and the maximizing distribution is
a Markovian solution of the martingale problem. This potential connection needs
to be worked out in a future research project.
From this point forward, we shall always assume the market price of risk to
be Markovian. As we shall see, this choice will lead directly to the optimal trading
strategy.
Remark 1 (Market price of risk and NA, NUPBR, NIA). Proposition 3.6 of Delbaen
and Schachermayer (1994) shows (compare also Proposition 3.2 of Karatzas and
Kardaras 2007) that NFLVR holds, if and only if NA and no unbounded prot
with bounded risk (NUPBR) hold. NUPBR is also known as arbitrage of the
rst kind (compare Ingersoll 1987; Kardaras and Platen 2009) and as the BK
Chapter 2. The Markovian Case 16
property (compare Kabanov 1997; Kardaras 2010, Proposition 1.2). If NUPBR
holds, then, in particular, scalable arbitrage opportunities do not exist.
The existence of a square-integrable market price of risk guarantees the ex-
istence of a positive stochastic discount factor, which again ensures that NUPBR
holds as it is proven in Theorem 3.12 of Karatzas and Kardaras (2007). Moreover,
since it is shown in Lemma 3.1 of Delbaen and Schachermayer (1995b) that NA
holds, if and only if no immediate arbitrage (NIA) holds and the possibility to
make some prot using a credit line is excluded. However, since immediate arbi-
trage is again scalable we can also conclude that NUPBR implies NIA. Therefore, if
NUPBR holds, then NFLVR fails, if and only if the second component of NA fails,
to wit, if and only if it is possible to make some prot using a credit line. Indeed,
the application of this chapters results to the optimal hedging problem of a bond
serves to quantify exactly how much some prot is in a given model.
On the other hand, a careful analysis of Section 10 in Karatzas et al. (1991a)
or Theorems 3.5 and 3.6 in Delbaen and Schachermayer (1995b), using the fact that
the ranges of (, ) and a(, ) are identical, reveals that a necessary condition for
NIA is the existence of a market price of risk that satises an integrability condition
strictly weaker than the condition in (2.4). Furthermore, Theorem 1 of Levental and
Skorohod (1995) and Proposition 1.1 of Lyaso (2010) motivate the integrability
condition in (2.4) to prevent general scalable arbitrage opportunities.
A toy example for a market without a market price of risk P Lebesgue-
almost everywhere (and thus with scalable arbitrage) can be described by a drift
(, ) and a volatility structure (, ) such that the set / [0, T] R
d
+
dened
as / := (t, s) : (t, s) s.t. (t, s)(t, s) = (t, s) has positive measure, by which
we mean p
,
:= P(Lebesgue(t : (t, S(t)) /) > 0) > 0. We can decompose
(, ) uniquely into the sum of two vectors
1
(, ) in the range of (, ) and
2
(, )
orthogonal to its columns. Then, we have
2
(t, s) ,= 0 for all (t, s) / and
Chapter 2. The Markovian Case 17
2
T
(, )(, ) 0 always. Investing according to
2
(, ) would thus switch o the
risk factors and lead to nonnegative mean rate of return
2
T
(, )(, ) = |
2
(, )|
2
.
Investing according to such a strategy (see Section 2.3 for a precise denition) would
lead to a wealth process (as in (2.11) below) that is greater than one with probability
p
,
. This arbitrage opportunity could be leveraged arbitrarily by replacing the
strategy
2
(, ) with
2
(, ) multiplied by a constant, leading to an immediate
and unbounded prot. This line of thought, enriched with deep measure-theoretic
results, is the underlying idea for the proof of the existence of a market price of risk
under the NIA condition in Theorem 3.5 of Delbaen and Schachermayer (1995b).
2.3 Strategies, wealth processes and arbitrage op-
portunities
In this section, we introduce trading strategies, describe investors wealth processes
and dene relative arbitrage. We denote the proportion of the investors wealth
invested in the i
th
stock by
i
. The proportion of the wealth that is not invested
in stocks gets invested in the money market, which yields zero interest rate. The
next denition states this more precisely.
Denition 2 (Markovian trading strategy and associated wealth process). We call
a function : [0, T] R
d
+
R
d
a (Markovian trading) strategy and the process
V
() with dynamics
dV
(t) =
d
i=1
i
(t, S(t))V
(t)
dS
i
(t)
S
i
(t)
(2.8)
for all t [0, T] and with initial condition V
i=1
_
T
0
(
i
(t, S(t))V
(t))
2
a
i,i
(t, S(t))dt < , (2.9)
and the nonnegativity condition V
i=1
i
(t, )dS
i
(t)
=
d
i=1
h
i
(t, )
dS
i
(t)
S
i
(t)
(2.10)
Chapter 2. The Markovian Case 19
=h
T
(t, )(t, S(t)) ((t, S(t))dt + dW(t))
for all t [0, T].
The conditions in (2.4) and (2.9) in conjunction with Holders inequality
yield that
d
i=1
_
T
0
[
i
(t, S(t))V
(t)
i
(t, S(t))[ dt <
almost surely, which guarantees the existence of a strong solution for V
(). If the
condition in (2.9) holds with (, )V
() replaced by (, ) then V
() stays strictly
positive. In this case, analog to (2.2), the solution of the stochastic dierential
equation in (2.8) is given as
V
(t) =exp
__
t
0
T
(u, S(u))(u, S(u))du +
_
t
0
T
(u, S(u))(u, S(u))dW(u)
1
2
_
t
0
T
(u, S(u))a(u, S(u))(u, S(u))du
_
(2.11)
for all t [0, T]. For example, the strategy
0
(, ) 0 invests only in the money
market and its associated wealth process satises V
0
() 1. Usually, trading
strategies do not lead to wealth processes that only depend on the current state of
the market, as the next remark discusses:
Remark 2 (Markovianness of wealth process and dependence on whole path). Ob-
viously, the wealth process of an investor jointly with the stock price process is
Markovian if the investor uses a Markovian trading strategy. Yet, at time t [0, T]
the wealth process does not only depend on the current stock prices S(t) but in most
cases also on past stock prices S(u), u t. Important exceptions from this rule
are the market portfolio
m
i
(t, s) := s
i
/
d
j=1
s
j
and investments in single stocks or
the money market only; that is,
j
i
(t, s) :=
j
(i) for some j 1, . . . , d +1 and for
all (t, s) [0, T] R
d
+
, where
j
represents Kroneckers delta function. However, as
we shall see in Theorem 1 of Section 2.3, the dependence of the associated wealth
Chapter 2. The Markovian Case 20
processes on the past does not represent a problem in our setup for nding optimal
strategies.
The change of numeraire, that is, the change of the denomination in which
the wealth process is quoted, is one of the most useful techniques in mathematical
nance; compare Geman et al. (1995) for a derivation and discussion of the change
of numeraire technique. It also plays a fundamental role in this chapter. For every
numeraire, a special market price of risk exists:
Denition 3 (-specic market price of risk). Let (, ) denote a strategy and
(, ) a market price of risk. Dene the corresponding -specic market price of
risk
(t, s) : [0, T] R
d
+
R
K
as
(t, s) := (t, s)
T
(t, s)(t, s). (2.12)
The following computations show that the -specic market price of risk
exactly translates the volatilities into the mean rates of return relative to the wealth
process of (, ). Let (, ) be any other strategy and V
(t)
V
(t)
=exp
_
_
t
0
((u, S(u)) (u, S(u)))
T
(u, S(u))du
+
_
t
0
((u, S(u)) (u, S(u)))
T
(u, S(u))dW(u)
1
2
_
t
0
_
T
(u, S(u))a(u, S(u))(u, S(u))
T
(u, S(u))a(u, S(u))(u, S(u))
_
du
_
and thus after a short calculation,
d
_
V
(t)
V
(t)
_
=
V
(t)
V
(t)
_
(t, S(t)) (t, S(t))
_
T
_
_
(t, S(t)) a(t, S(t))(t, S(t))
_
dt
+ (t, S(t))dW(t)
_
Chapter 2. The Markovian Case 21
=
V
(t)
V
(t)
_
(t, S(t)) (t, S(t))
_
T
(t, S(t))dW
(t), (2.13)
where
W
(t) := W(t) +
_
t
0
(t)V
v,
(t) = v exp
_
_
t
0
T
(u, S(u))dW(u)
1
2
_
t
0
|
(u, S(u))|
2
du
_
(2.15)
for all t [0, T].
Remark 3 (Change of numeraire). The expression in (2.15) should be contrasted to
one in (2.5). The market price of risk (, ) is replaced by the -specic market price
of risk
(, ) when we multiply Z
()V
v,
() is not a true
martingale, then Z
(T)V
(T) V
(T)) = 1
and P(V
(T) > V
()V
() is a true mar-
tingale if and only if there exists an equivalent local martingale measure Q, under
which the stock price processes are local martingales. The question of whether Q
is a martingale measure or only a local martingale measure is not connected to
whether Z
: [0, T] R
d
+
[0, 1]
as
U
(t, s) :=E
Tt,s
_
Z
(T)V
(T)
Z
(T t)V
(T t)
_
(2.16)
=E
Tt,s
_
exp
_
_
T
Tt
T
(u, S(u))dW(u)
1
2
_
T
Tt
|
(u, S(u))|
2
du
__
.
(2.17)
The last equality follows directly from (2.15). As we show in Theorem 1, U
can
be interpreted as a hedging price. It obviously depends on the strategy (, ).
Chapter 2. The Markovian Case 24
Proposition 1 yields that U
t
U
(t, s) =
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(T t, s)D
2
i,j
U
(t, s)
+
d
i=1
d
j=1
s
i
a
i,j
(T t, s)
j
(T t, s)D
i
U
is sucient for U
to
solve the PDE.
The next theorem is the rst key result of this chapter. It shows that U
() and let U
(T).
Thus, whenever Z
()V
(T t, s) +
i
(t, s) (2.19)
for all (t, s) [0, T] R
d
+
. Furthermore, (, ) is optimal: There exists no strategy
(, ) such that
V
v,
(T) V
(T) = V
v,
(T) (2.20)
almost surely for some v < v.
Proof. Let us start by dening the martingale N
() as
N
(t) :=E[Z
(T)V
(T)[T(t)] = Z
(t)V
(t)U
(T t, S(t)) (2.21)
for all t [0, T] and denoting by L the innitesimal generator of S(), that is,
L =
d
i=1
s
i
i
(t, s)D
i
+
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
. (2.22)
Since
dU
(T t, S(t)) =
_
LU
t
U
_
(T t, S(t)) dt
+
K
k=1
d
i=1
S
i
(t)
i,k
(t, S(t))D
i
U
(T t, S(t))W
k
(t)
holds for all t [0, T], the product rule of stochastic calculus and (2.15) yield
dN
(t)
Z
(t)V
(t)
=dU
(T t, S(t)) + U
(T t, S(t))
d(Z
(t)V
(t))
Z
(t)V
(t)
k=1
k
(t, S(t))
d
i=1
S
i
(t)
i,k
(t, S(t))D
i
U
(T t, S(t))dt.
We obtain the equality
dN
(t)
N
(t)
=
K
k=1
_
d
i=1
S
i
(t)
i,k
(t, S(t))
D
i
U
(T t, S(t))
U
(T t, S(t))
k
(t, S(t))
_
dW
k
(t)
Chapter 2. The Markovian Case 26
+ C
(t, S(t))dt,
where
C
(t, s) :=
_
LU
t
U
_
(T t, s)
U
(T t, s)
d
i=1
s
i
D
i
U
(T t, s)
U
(T t, s)
K
k=1
k
(t, s)
i,k
(t, s)
=
1
U
(T t, s)
_
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
U
(T t, s) (2.23)
+
d
i=1
d
j=1
s
i
a
i,j
(t, s)
j
(t, s)D
i
U
(T t, s)
t
U
(T t, s)
_
=0,
since U
(t)
N
(t)
=
K
k=1
_
d
i=1
i,k
(t, S(t))
_
S
i
(t)D
i
log (U
(T t, S(t))) +
i
(t, S(t))
_
k
(t, S(t))
_
dW
k
(t)
=
K
k=1
k
(t, S(t))dW
k
(t)
=
d(Z
(t)V
v,
(t))
Z
(t)V
v,
(t)
,
where the last equality follows from (2.15). Then, N
(0) = v = Z
(0)V
v,
(0) and
both processes N
() and Z
()V
v,
() have the same dynamics such that
Z
(T)V
(T) = N
(T) = Z
(T)V
v,
(T);
see Theorem 1.4.61 of Jacod and Shiryaev (2003). Since zero is an absorbing state
for any nonnegative supermartingale and since Z
(T)V
()V
v,
() is bounded from below by zero, further
Chapter 2. The Markovian Case 27
implying that it is a supermartingale. Assume we have some strategy (, ) such
that (2.20) is satised. Then, we obtain
v E[Z
(T)V
v,
(T)] E[Z
(T)V
(T)] = E[Z
(T)V
v,
(T)] = v, (2.24)
which concludes the proof.
We obtain from (2.21) and the last proof that
V
v,
(t) =
N
(t)
Z
(t)
= V
(t)U
(T t, S(t)), (2.25)
which we can rewrite as
U
(T t, S(t)) =
V
v,
(t)
V
(t)
.
Thus, U
()V
() V
(); to wit,
the optimal wealth process is unique.
The next remarks discuss various assumptions of the last theorem:
Remark 4 (Completeness of the market). One remarkable feature of the last result
is that we have not required the market to be complete. In contrast to Fernholz and
Karatzas (2010), we do not rely on the martingale representation theorem but in-
stead directly derive a representation for the conditional expectation process of the
nal wealth V
(T),
Chapter 2. The Markovian Case 28
an optimal strategy (, ) exists to achieve V
solves the PDE in (2.18). Sucient conditions are existence and dierentiability
conditions on the function H
C
1,2
([0, T] R
d
+
). Then, the proof of Theorem 1 yields that U
automatically solves the PDE in (2.18), at least at all points (t, s) [0, T]R
d
+
that
can be attained by S() at time t. This can be seen from the fact that the process
N
(, ) 0, where C
is not
dierentiable, then we can nd an optimal strategy up to time t
1
, starting from t
1
to
t
2
and so on. This will neither change the optimal strategy (, ) nor the minimal
initial capital v in any way. This small modication allows us to include strate-
gies (, ) with structural breaks, by which we mean strategies whose arbitrage
Chapter 2. The Markovian Case 29
properties are changed at nitely many time steps. An example is a full investment
up to time t (0, T) in one strategy that can be arbitraged and afterwards a full
investment in another strategy that cannot be arbitraged.
Furthermore, as Example 5 of Section 2.6 illustrates, the dierentiability of
U
in the stock price dimension is only a sucient but not a necessary condition
for the existence of an optimal strategy.
The PDE in (2.18) always has the constant function as a solution. The next
result classies U
). The function U
t
U(t, s) =
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(T t, s)D
2
i,j
U(t, s)
+
d
i=1
d
j=1
s
i
a
i,j
(T t, s)
j
(T t, s)D
i
U(t, s),
where we have exchanged (, ) by (, ), has U
(, ) 1 as its minimal nonnega-
tive solution.
Proof. Consider any suciently smooth function
U
: [0, T] R
d
+
R
+
that solves
the PDE in (2.18) and the initial condition
U
(t) := Z
(t)V
(t)
(T t, S(t))
for all t [0, T], which is, as in the proof of Theorem 1, a positive supermartingale.
Thus, we have
(T t, S(t)) =
(t)
Z
(t)V
(t)
Chapter 2. The Markovian Case 30
E
t,S(t)
_
(T)
_
Z
(t)V
(t)
=
E
t,S(t)
_
Z
(T)V
(T)
(t)V
(t)
= U
(T t, S(t))
for all t [0, T]. The second statement of the proposition comes from the fact
that Z
()V
() is a martingale, which implies U
(, ) 1, and from the same
considerations as above.
The hedging price for the stock of Example 4 in Section 2.6, for instance,
is one of many solutions of polynomial growth of the corresponding Black-Scholes
type PDE. For example, consider h
1
(t, s) := s and h
1
(t, s) times the hedging price
of (2.52), that is, h
2
(t, s) := 2s(1/(s
t
h(t, s) =
1
2
s
4
D
2
h(t, s)
with the identical boundary condition h(t, 0) = 0 and h(0, s) = s for all (t, s)
[0, T] R
+
.
The reason for non-uniqueness in this case is the fact that the second-order
coecient has super-quadratic growth preventing standard theory from being ap-
plied; see, for example, Section 5.7.B of Karatzas and Shreve (1991). Furthermore,
the boundary condition at innity is not specied precisely enough. Both solutions
grow polynomially, but clearly h
2
is always smaller than h
1
. In this specic ex-
ample, the corresponding process 1/S() is a three-dimensional Bessel process and
therefore stays away from the boundary. If the drift, however, is removed, it is a
Brownian motion, which can hit zero. Thus, boundary conditions need to be pre-
cisely specied for a PDE in 1/s at zero, which corresponds to the precise boundary
Chapter 2. The Markovian Case 31
condition at innity for the PDE above. Indeed, as the next section shows, the ex-
istence of an arbitrage opportunity is equivalent to the positive probability of some
process imploding to zero under some measure Q, which corresponds exactly to the
observation that 1/S() in the above example can hit zero.
For the special case (, ) 0 in one dimension, and under some assumptions
on the volatility parameter (, ), Ekstrom et al. (2009) suggest a numerical algo-
rithm that utilizes this characterization and nds the minimal nonnegative solution
of a Black-Scholes type PDE that does not have a unique solution.
2.4.2 Hedging of contingent claims
So far, we have started from a given Markovian trading strategy (, ) and then
optimized it. However, one might imagine situations in which one wants to hedge
a contingent claim but does not know a possibly suboptimal strategy (, ) a priori.
How can we nd, in such a situation, an optimal strategy? In the following we re-
solve this problem for Markovian claims. We shall also provide weak sucient con-
ditions for the corresponding hedging price to be dierentiable in Subsection 2.4.3.
We now explicitly allow the associated wealth processes to hit zero.
To simplify computations later on, we introduce some notation. As before,
the expectation operator corresponding to the event S(t) = s is written as E
t,s
.
Using the Markovian structure of our model, we denote, outside of the expectation
operator, by (S
t,s
(u))
u[t,T]
a stock price process with the dynamics of (2.1) and
S(t) = s, in particular, S
0,S(0)
() S(). We observe that Z
(u)/Z
(t) depends
for u (t, T] on T(t) only through S(t) and we write similarly (
Z
,t,s
(u))
u[t,T]
for
(Z
(u)/Z
(t))
u[t,T]
, with
Z
,t,s
(t) = 1 on the event S(t) = s. When we want to
stress the dependence of a process on the state we shall write, for example,
S(t, ).
We emphasize the standing assumptions made in Section 2.2, namely, that
Chapter 2. The Markovian Case 32
the stock price process S() with dynamics specied in (2.1) starting in S(0) R
d
+
is R
d
-valued, unique and stays in the positive orthant. Furthermore, a square-
integrable Markovian market price of risk exists almost surely.
For any measurable function p : R
d
+
[0, ), representing the payo of the
contingent claim, we dene a candidate h
p
: [0, T] R
d
+
[0, ) for the hedging
price of the corresponding European option, similar to the denition of U
in (2.16)
as
h
p
(t, s) := E
Tt,s
_
(T)p(S(T))
_
. (2.26)
The only dierence between h
p
and U
p
i
(t, s) := s
i
D
i
log (h
p
(T t, s)) (2.27)
for all i = 1, . . . , d and (t, s) [0, T] R
d
+
, and with v
p
:= h
p
(T, S(0)), we get
V
v
p
,
p
(T t) = h
p
(t, S(T t))
for all t [0, T]. Furthermore, the strategy
p
(, ) is optimal in the sense of Theo-
rem 1 and h
p
solves the PDE
t
h
p
(T t, s) =
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
h
p
(T t, s) (2.28)
at all points of support (t, s) for S().
Proof. Let us start by dening the martingale N
p
() as
N
p
(t) :=E[Z
(T)p(S(T))[T(t)] = Z
(t)h
p
(T t, S(t))
Chapter 2. The Markovian Case 34
for all t [0, T]. Although h
p
is not assumed to be in C
1,2
((0, T]R
d
) but only to be
locally smooth, we can apply a localized version of Itos formula; see, for example,
Section IV.3 of Revuz and Yor (1999). Then, the product rule of stochastic calculus
can be used to obtain the dynamics of N
p
(). Since N
p
() is a martingale, the
corresponding dt-term must disappear. This observation, in connection with (2.3)
and the positivity of Z
i=1
p
i
(t, S(t))
dS
i
(t)
S
i
(t)
for all t [0, T]. Then, both h
p
(T , S()) and V
v
p
,
p
() are stochastic exponentials
and solve the same stochastic dierential equation. Theorem 1.4.61 of Jacod and
Shiryaev (2003) yields h
p
(T , S()) V
v
p
,
p
().
The optimality of
p
(, ) follows exactly as in Theorem 1.
The last result generalizes Proposition 3 of Platen and Hulley (2008), where
the same result is derived for a one-dimensional, complete market with a time-
transformed squared Bessel process of dimension four modeling the stock price
process.
We remark that, as before, we neither assumed a complete market nor uti-
lized a representation theorem. In particular, at no point did we assume invertibility
or full rank of the volatility matrix (, ). Under these general assumptions, there
is no hope to be able to hedge all contingent claims on the Brownian motion W(T).
However, W(T) appears in this class of models only as a nuisance parameter and
it is of no economic interest to trade in it directly.
Remark 6 (Delta hedging). Writing (2.27) as
p
i
(t, S(t))
V
v
p
,
p
(t)
S
i
(t)
= D
i
h
p
(T t, S(t))
Chapter 2. The Markovian Case 35
and observing that the left-hand side is the number of shares invested in stock i
at time t shows that the optimal strategy is a delta hedge as in classical Financial
Mathematics, when one tries to hedge a contingent claim. Of course, (, ) of (2.19)
can be interpreted in a similar way: U
(T)(L S
1
(T))
+
_
+ s
1
U
1
(T t, s)
= E
t,s
_
(T)(S
1
(T) L)
+
_
+ LU
0
(T t, s), (2.29)
where
0
(, ) 0 denotes the strategy for holding a monetary unit and
1
(, )
(1, 0, . . . , 0)
T
the strategy for holding stock S
1
().
Proof. The statement follows from the linearity of expectation.
Due to Theorem 2, under weak dierentiability assumptions, optimal strate-
gies exist for the money market, the stock S
1
(T), the call and the put. Thus, the
5
We thank Peter Carr for pointing us to this reference.
Chapter 2. The Markovian Case 36
left-hand side of (2.29) corresponds to the sum of the hedging prices of a put and
the stock, and the right-hand side corresponds to the sum of the hedging prices of
a call and L monetary units. The dierence between this and the classical put-call
parity is that the current stock price and the strike L are replaced by their hedging
prices. Section 2.2 of Bayraktar et al. (2010b) have recently observed an another
version of the put-call parity. Instead of replacing the current stock price by its
hedging price, they replace the European call price by the American call price and
restore the put-call parity this way.
2.4.3 Smoothness of hedging price
Next, we shall provide sucient conditions under which the function h
p
of the last
subsection is suciently smooth. Towards this end, we need the following denition.
Denition 7 (Locally Lipschitz and locally bounded). We call a function f : [0, T]
R
d
+
R locally Lipschitz and locally bounded on R
d
+
if for all s R
d
+
the function
t f(t, s) is right-continuous with left limits and for all M > 0 there exists some
C(M) < such that
sup
1
M
|y|,|z|M
y,=z
[f(t, y) f(t, z)[
|y z|
+ sup
1
M
|y|M
[f(t, y)[ C(M)
for all t [0, T].
In particular, if f has continuous partial derivatives, it is locally Lipschitz
and locally bounded. We require several assumptions in order to show the neces-
sary dierentiability of h
p
in Theorem 3 below. It is subject to future research to
determine the precise conditions which yield the existence of a delta hedge, possibly
without requiring h
p
to be the classical solution of a PDE.
(A1) The functions
k
(, ) and
i,k
(, ) are for all i = 1, . . . , d and k = 1, . . . , K
locally Lipschitz and locally bounded.
Chapter 2. The Markovian Case 37
(A2) For all points of support (t, s) for S() there exist some C > 0 and some
neighborhood | of (t, s) such that
d
i=1
d
j=1
a
i,j
(u, y)
i
j
C||
2
(2.30)
for all R
d
and (u, y) |.
(A3) The payo function p is chosen so that for all points of support (t, s) for S()
there exist some C > 0 and some neighborhood | of (T t, s) such that
h
p
(u, y) C for all (u, y) |.
If h
p
is constant for
d d coordinates, say the last ones, Assumption (A2) can
be weakened to requesting the uniform ellipticity only in the remaining d
d 1
coordinates; that is, the sum in (2.30) goes only to d
d 1 and R
d
d1
.
Assumption (A3) holds in particular if p is of linear growth; that is, if p(s)
C
d
i=1
s
i
for some C > 0 and all s R
d
+
, since
Z
,t,s
()S
t,s
i
() is a nonnegative
supermartingale for all i = 1, . . . , d.
We emphasize that the conditions here are weaker than the ones used in
Section 9 of Fernholz and Karatzas (2010) for the case of the market portfolio,
which can be represented as p(s) =
d
i=1
s
i
. In particular, the stochastic integral
component in Z
k=1
i,k
(u, Y
t,y
(u))dW
k
(u), (2.31)
Y
t,y
i
(t) = y
i
for all u [t, T], for all (t, y) [0, T] R
d
, where Y
t,y
() = (Y
t,y
1
(), . . . , Y
t,y
d
())
T
denotes a
d-dimensional vector. The drift : [0, T] R
d
R
d
and volatility
coecient : [0, T] R
d
R
dK
are assumed to be measurable and to satisfy the
global Lipschitz condition
i=1
[
i
(u, y
1
)
i
(u, y
2
)[ +
i=1
K
k=1
[
i,k
(u, y
1
)
i,k
(u, y
2
)[ C|y
1
y
2
|
for all (u, y
1
, y
2
) [0, T] R
d
R
d
for some constant C > 0.
Then, the stochastic dierential equation in (2.31) has a unique solution
Y
t,s
(). It has a modication, which we again call Y
t,s
(), and which satises the
following continuity property: Fix any countable set of time indices T = t
i
[0, T]
iN
. Then, there exists a subset
with P(
d
with |y
1
y
2
| 2
k2
we have
sup
u[t,T]
|Y
t,y
1
(u) Y
t,y
2
(u)| c
1
()2
c
2
()k
for some constants c
1
(), c
2
() R
+
. In particular, the constants c
1
() and c
2
()
can be chosen independently of t T.
Proof. The lemma basically states Theorem V.37 of Protter (2003). The explicit
continuity comes from an analysis of the arguments in the proof of Kolmogorovs
Lemma; compare Theorem IV.73 of Protter (2003). There, we use Chebyshevs
inequality simultaneously for all t T and then follow the proof line by line.
Chapter 2. The Markovian Case 39
We can now prove the continuity of the process X
t,s,1
() in t and s using a
localization technique:
Lemma 2 (Stochastic ow, locally Lipzschitz). We x a point (t, s) [0, T]R
d
+
so
that X
t,s,1
() is strictly positive and an R
d+1
+
-valued process. Then, under Assump-
tion (A1), we have for all sequences (t
k
, s
k
)
kN
[0, T] R
d
+
with lim
k
(t
k
, s
k
) =
(t, s) that
lim
k
sup
u[t,T]
|X
t
k
,s
k
,1
(u) X
t,s,1
(u)| = 0
almost surely, where we set X
t
k
,s
k
,1
(u) := (s
T
k
, 1)
T
for u t
k
. In particular, for
K() suciently large we have that X
t
k
,s
k
,1
(u, ) is strictly positive and R
d+1
+
-valued
for all k > K() and u [t, T].
Proof. Since the class of locally Lipschitz and locally bounded functions is closed
under summation and multiplication, Assumption (A1) yields that the drift and
diusion coecients of X
u,y,z
() are locally Lipschitz for all (u, y, z) [0, T] R
d
+
R
+
. We start by assuming t
k
t for all k N and obtain
sup
u[t,T]
|X
t
k
,s
k
,1
(u) X
t,s,1
(u)| sup
u[t,t
k
]
|(s
T
k
, 1)
T
X
t,s,1
(u)|
+ sup
u[t
k
,T]
|X
t
k
,s
k
,1
(u) X
t
k
,s,1
(u)|
+ sup
u[t
k
,T]
|X
t
k
,s,1
(u) X
t
k
,S
t,s
(t
k
),
Z
,t,s
(t
k
)
(u)|
for all k N. The rst term on the right-hand side of the last inequality goes to
zero as k increases by the continuity of the sample paths of X
t,s,1
(). The arguments
and the localization technique in the proof of Theorem V.38 in Protter (2003) in
conjunction with Lemma 1 yield that
lim
k
sup
u[
t,T]
|X
t,y
k
,z
k
(u) X
t,s,1
(u)| = 0
Chapter 2. The Markovian Case 40
for all
t t, t
1
, t
2
, . . . and any sequence ((y
T
k
, z
k
)
T
)
kN
R
d+1
+
with (y
T
k
, z
k
)
T
(s
T
, 1)
T
as k almost surely. The convergence is uniformly in
t t, t
1
, t
2
, . . ..
We now choose for (y
T
k
, z
k
)
T
the sequences (s
T
k
, 1)
T
and (S
t,s
T
(t
k
, ),
Z
,t,s
(t
k
, ))
T
for all . This proves the statement if t
k
t for all k N. In the case of the
reversed inequality t
k
t, we observe
sup
u[t,T]
|X
t
k
,s
k
,1
(u) X
t,s,1
(u)| sup
u[t,T]
|X
t
k
,s
k
,1
(u) X
t
k
,s,1
(u)|
+ sup
u[t,T]
|X
t
k
,s,1
(u) X
t,s,1
(u)|,
which again yields continuity, similar to above.
In the second step, we use a technique from the theory of PDEs to conclude
the necessary smoothness of h
p
. The following result has been used by Ekstrom,
Janson and Tysk. We present it here on its own to underscore the analytic compo-
nent of our argument:
Lemma 3 (Schauder estimates and smoothness). Fix a point (t, s) [0, T) R
d
+
and a neighborhood | of (t, s). Suppose Assumption (A1) holds along with inequality
in (2.30) for all R
d
and (u, y) | and some C > 0. Let (f
k
)
kN
denote
a sequence of solutions of the PDE in (2.28) on |, uniformly bounded under the
supremum norm on |. If lim
k
f
k
(t, s) = f(t, s) on | for some function f : |
R, then f solves the PDE in (2.28) on some neighborhood
| of (t, s). In particular,
f C
1,2
(
|).
Proof. We refer the reader to the arguments and references provided in Section 2 of
Janson and Tysk (2006) and Theorem 3.2 of Ekstrom and Tysk (2009). The central
idea is to use the interior Schauder estimates by Knerr (1980) in conjunction with
Arzel`a-Ascoli type of arguments to prove the existence of rst- and second-order
derivatives of f.
We can now prove the smoothness of the hedging price h
p
:
Chapter 2. The Markovian Case 41
Theorem 3. Under Assumptions (A1)-(A3) there exists for all points of support
(t, s) for S() some neighborhood | of (T t, s) such that the function h
p
dened
in (2.26) is in C
1,2
(|).
Proof. We dene p : R
d+1
+
R
+
by p(s
1
, . . . , s
d
, z) := zp(s
1
, . . . , s
d
) and p
M
:
R
d+1
+
R
+
by p
M
() := p()1
p()M
for some M > 0 and approximate p
M
by a
sequence of continuous functions p
M,m
(compare for example Appendix C.4 of Evans
1998) such that lim
m
p
M,m
= p
M
pointwise and p
M,m
2M for all m N. The
corresponding expectations are dened as
h
p,M
(u, y) := E
Tu,y
[ p
M
(S
1
(T), . . . , S
d
(T),
Z
(T))]
for all (u, y)
| for some neighborhood
| of (T t, s) and equivalently
h
p,M,m
.
We start by proving continuity of
h
p,M,m
for large m. For any sequence
(t
k
, s
k
)
kN
[0, T] R
d
+
with lim
k
(t
k
, s
k
) = (t, s), Lemma 2, in connection with
Assumption (A1), yields
lim
k
p
M,m
(S
t
k
,s
k
(T),
Z
,t
k
,s
k
(T)) = p
M,m
(S
t,s
(T),
Z
,t,s
(T)).
The continuity of
h
p,M,m
follows then from the bounded convergence theorem.
Now, Lemma 2.6 of Janson and Tysk (2006), in connection with Assump-
tion (A2), guarantees that
h
p,M,m
is a solution of the PDE in (2.28). Lemma 3 then
yields that rstly,
h
p,M
and secondly, in connection with Assumption (A3), h
p
also
solve the PDE in (2.28) on some neighborhood | of (T t, s). In particular, h
p
is
in C
1,2
(|).
The last theorem generalizes the results in Ekstrom and Tysk (2009) to
several dimensions and to non-continuous payo functions p. Chapters 6 and 15
of Friedman (1976) and Janson and Tysk (2006) have related results, but they
impose linear growth conditions on a(, ) so that the PDE in (2.28) has a unique
solution of polynomial growth. We are especially interested in the situation in
Chapter 2. The Markovian Case 42
which multiple solutions may exist. Heath and Schweizer (2000) present results in
the case when the process corresponding to the PDE in (2.28) does not leave the
positive orthant. As Fernholz and Karatzas (2010) observe, this condition does not
necessarily hold if there is no equivalent local martingale measure. In the case of
Z
() being a martingale, our assumptions are only weakly more general than the
ones in Heath and Schweizer (2000) by not requiring a(, ) to be continuous in the
time dimension. Further results are also obtained by Section III.7 of Kunita (1984),
but under strong continuity assumptions on a(, ). However, in all these research
articles, the authors show that the function h
p
indeed solves the PDE in (2.28) not
only locally but globally and satises the corresponding boundary conditions. We
have here abstained from imposing the stronger assumptions these papers rely on
and concentrate on the local properties of h
p
. For our application, it is sucient to
observe that h
p
(T t, S(t)) converges to p(S(T)) as t goes to T; compare the proof
of Theorem 2.
The next section provides an interpretation of our approach to prove the
dierentiability of h
p
; all problems on the spatial boundary, arising for example
from a discontinuity of a(, ) on the boundary of the positive orthant, have been
conditioned away, so that S() can get close to but never actually attains the
boundary.
2.5 Change of measure
We obtained in Theorem 1 a precise description of an optimal strategy (, ) to
replicate the wealth V
(T)V
by
performing a change of measure. To be able to do then the computations, we
provide the dynamics of the stock price processes and a formula for conditional
Chapter 2. The Markovian Case 43
expectations under the new probability measure in Corollaries 2 and 3. We end
this section by proving a result concerning the change of numeraire in Corollary 4,
illustrating in Proposition 3 how a canonical probability space can be constructed to
satisfy the technical assumptions of this section, and in several remarks discussing
connections of this work to some literature.
Theorem 1.4 of Delbaen and Schachermayer (1995b) shows that NA implies
the existence of a local martingale measure which is absolutely continuous with
respect to P. On the other side, a consequence of this section is the existence of a
local martingale measure under NUPBR, such that P is absolutely continuous with
respect to it. Indeed, as discussed in Remark 1, NA and NUPBR together yield
NFLVR, which again yields an equivalent local martingale measure corresponding
exactly to the one discussed in this section. Another point of view, which we
do not take here, is the recent insight by Kardaras (2010) on the equivalence of
NUPBR and the existence of a nitely additive probability measure that is, in
some sense, weakly equivalent to P and under which S() has some notion of weak
local martingale property.
Our approach via a generalized change of measure is in the spirit of the
work by Follmer (1972; 1973), Meyer (1972), Section 2 of Delbaen and Schacher-
mayer (1995a), and Section 7 of Fernholz and Karatzas (2010). They show that for
the strictly positive P-local martingale Z
(T
),
where
(T)] =
Q(
(),
which do not occur under P, but may occur under a new probability measure Q
constructed below. We further assume that the ltration F is the right-continuous
modication of the ltration generated by the paths , or more precisely by the
projections
t
() := (t). Concerning the original probability measure we assume
that P( : (T) = ) = 0 and that for all t [0, T], is an absorbing state for
Z
(t) = implies Z
(T
). After having ensured its existence, one then can take the
route suggested by Theorem 5 of Delbaen and Schachermayer (1995a) and start
from any probability space satisfying the usual conditions, construct a canonical
probability space satisfying the technical assumptions mentioned above, and then
perform all necessary computations on this space. We shall detail these technical
steps in Proposition 3.
For now, the goal is to construct a measure Q under which the computation
of U
i
:= inft [0, T] : Z
(t) i
with inf := and the sequence of -algebras T
i
:= T(
i
T) for all i N.
We observe that the denition of T
i
is independent of the probability measure and
dene the stopping time
:= lim
i
i
with corresponding -algebra T
,
:=
T(
T) generated by
i=1
T
i,
.
Within this framework, Meyer (1972) and Example 6.2.2 of Follmer (1972)
rely on an extension theorem (compare Chapter 5 of Parthasarathy 1967) to show
the existence of a measure Q on (, T(T)) satisfying
Q(A) = E
P
_
Z
i
T)1
A
(2.32)
for all A T
i,
, where we now write E
P
for the expectation under the original
measure. We summarize these insights in the following theorem:
Chapter 2. The Markovian Case 46
Theorem 4 (Generalized change of measure). There exists a measure Q such that
for all stopping times
T with
T T and for all T(
T
_
D
_
= E
Q
_
D1
1/Z
T)>0
_
, (2.33)
where E
Q
denotes the expectation with respect to the new measure Q. That is,
P is absolutely continuous with respect to Q. Under this measure Q, the process
W() =
_
W
1
(), . . .
W
K
()
_
T
with
W
k
(t
) := W
k
(t
) +
_
t
k
(u, S(u))du (2.34)
for all k = 1, . . . , K and t [0, T] is a K-dimensional Brownian motion stopped at
time
.
Proof. The existence of a measure Q satisfying (2.32) follows as in the discussion
above. Now, for any set A T(
T) we have
A =
_
A
_
T
__
_
i=1
_
A
_
i1
<
T
i
__
.
From the fact that
T holds, if and only if 1/Z
T
_ > 0
_
_
_
_
_
=
i=1
Q
_
A
_
i1
<
T
i
__
=
i=1
E
P
_
Z
i
T)1
A
i1
<
_
=
i=1
E
P
_
Z
T
_
1
A
i1
<
_
=E
P
_
Z
T
_
1
A
_
,
where the last identity holds since P
_
T
_
= 0. This yields the representation of
(2.33). From Girsanovs theorem (compare Theorem 8.1.4 of Revuz and Yor 1999)
Chapter 2. The Markovian Case 47
we obtain that on T
i,
the process
W() is under Q a K-dimensional Brownian
motion stopped at
i
T. Since
i=1
T
i,
generates T
,
and forms a -system, we
get the dynamics of (2.34).
Thus, an equivalent local martingale measure exists, if and only if Q(1/Z
(T) >
0) = 1. On the other hand, if no equivalent local martingale measure exists, then
valuing a wealth process must include the barrier aspect 1/Z
()
of the stochastic discount factor hitting zero. We emphasize that we need not know
(, ) to calculate the corresponding hedging price U
under Q according to
dS
i
(t) = S
i
(t)
K
k=1
i,k
(t, S(t))d
W
k
(t)
and
d
_
1
Z
(t)
_
=
1
Z
(t)
d
k=1
k
(t, S(t))d
W
k
(t)
for all i = 1, . . . , d and t [0, T].
Proof. This is a direct consequence of the representation of
W() in (2.34) and
Denition 1 of the market price of risk.
Chapter 2. The Markovian Case 48
The results of the last corollary play an essential role when we do com-
putations, since the rst hitting time of the reciprocal of the stochastic discount
factor can in most cases be easily represented as a rst hitting time of the stock
price. This now usually follows some more tractable dynamics, as we shall see in
Section 2.6. Theorem 4 also holds for expectations conditioned on T(t): the next
corollary generalizes the well-known Bayes rule for classical changes of measures;
compare Lemma 3.5.3 of Karatzas and Shreve (1991). Similar computations appear
already in Proposition 4.2 of Follmer (1972).
Corollary 3 (Bayes rule, Q-martingale property of 1/Z
()). Let
T denote any
stopping time with
T T. For all T(
T)>0
T(t)
_
= E
P
_
Z
T
_
D[T(t)
_
1
Z
_
t
T
_1
1/Z
(t
T)>0
(2.35)
holds Q-almost surely (and thus P-almost surely) for all t [0, T]. Furthermore,
for any process N(), N()1
1/Z
()>0
is a Q-martingale, if and only if N()Z
() is
a P-martingale. In particular, the process 1/Z
() is a Q-martingale.
Proof. We observe that for all T
_
i
T
_
-measurable random variables D 0,
(2.32) can be rewritten as
E
Q
[D] = E
P
_
Z
i
T
_
D
_
. (2.36)
due to the martingale property of the bounded process Z
i
) under P. We x
a time t [0, T]. For any A T(t), (2.33) with D replaced by D1
A
where
A =
A
_
T > t
_
T
_
T t
_
and the same techniques, as in the proof of Theorem 4,
yield
E
Q
_
D1
1/Z
T)>0
1
A
_
=E
P
_
Z
T
_
D1
A
_
Chapter 2. The Markovian Case 49
=E
P
_
_
E
P
_
Z
T
_
D
T(t)
_
1
Z
_
t
T
_1
A
Z
_
t
T
_
_
_
=
i=1
E
P
_
E
P
_
Z
T
_
D
T(t)
_
1
Z
_
t
T
_
1
i1
<t
A
Z
i
t
T
_
_
=
i=1
E
Q
_
_
E
P
_
Z
T
_
D
T(t)
_
1
Z
_
t
T
_1
i1
<t
A
_
_
=E
Q
_
_
E
P
_
Z
T
_
D
T(t)
_
1
Z
_
t
T
_1
1/Z
(t
T)>0
1
A
_
_
,
where the second-to-last equality relies on the identity of (2.36). This yields the
representation in (2.35). The other statements follow from choosing
T = T, D =
N(T) and D = 1/Z
(T).
For the case of strict local martingales the equivalence of the last corollary is
generally not true. Take as an example N() 1 and Z
()N() Z
()>0
1
1/Z
()>0
is clearly not a local Q-martingale. The reason for this lack of symmetry
is that a sequence of stopping times that converges P-almost surely to T need not
necessarily converge Q-almost surely to T.
We have seen that Theorem 4 implies that 1/Z
() stopped at zero is a
martingale under the new measure. As Delbaen and Schachermayer (1995a) and
Pal and Protter (2010) have discussed, the other direction holds trivially true:
Let Q denote some measure; M() a Q-martingale started at some positive value
M(0) > 0; and T
0
the rst hitting of zero by M(). Then, under the new measure
d
P := M(TT
0
)dQ, the process 1/M() is again a local martingale due to Girsanovs
theorem and Itos formula. It is a martingale, if and only if M() does not hit zero
Chapter 2. The Markovian Case 50
under the original measure Q.
In order to simplify computations even more, the following change of numeraire
for strictly positive wealth processes can be useful.
Corollary 4 (Change of numeraire). Let (, ) and (, ) denote two strategies
such that V
and
E
P
_
Z
T
_
V
T
__
= E
Q
_
_
V
T
_
V
T
_1
1/(Z
T)V
(
T))>0
_
_
, (2.37)
where E
Q
. Under this
measure Q
, the process W
() = (W
1
(), . . . W
K
())
T
with
W
k
(t
) := W
k
(t
) +
_
t
k
(u, S(u))du (2.38)
for all k = 1, . . . , K and t [0, T] is a K-dimensional Brownian motion stopped
at time
:= lim
i
inft [0, T] : Z
(t)V
(, ) here
is exactly the -specic market price of risk from Denition 3. The equality in
(2.13) holds until the stopping time
()V
under Q
according to
dS
i
(t) = S
i
(t)
d
j=1
a
i,j
(t, S(t))
j
(t, S(t))dt +S
i
(t)
K
k=1
i,k
(t, S(t))dW
k
(t) (2.39)
and
d
_
1
Z
(t)V
(t)
_
=
(t, S(t))
1
Z
(t)V
(t)
dW
(t)
for all i = 1, . . . , d and t [0, T]. Furthermore, the statements of Corollary 3 hold
with Q replaced by Q
and Z
() replaced by Z
()V
(T t, s) = Q
_
1
Z
(T)V
(T)
> 0
T(t)
_
. (2.40)
Chapter 2. The Markovian Case 51
Proof. The proof goes exactly along the lines of Theorem 4 and Corollaries 2 and 3
with the obvious modications.
We emphasize the similarity of the Q
.
The next proposition demonstrates how one can construct a probability space
that satises the technical conditions of this section:
Proposition 3 (Canonical probability space). Let (, T, P) denote any probability
space, equipped with a ltration F = T() that satises the usual conditions. There
exists a probability space (
,
T,
P), equipped with a ltration
T(), which supports
a probability measure Q such that
P is absolutely continuous with respect to Q, and
such that (2.33) holds for any
T(
T()-stopping time
T.
Furthermore, (
,
T,
P) has the same distributional properties as (, T, P);
that is, it supports a K-dimensional Brownian motion W(), a vector-valued process
(, ), a vector-valued process (, ), a matrix-valued process (, ), a d-dimensional
progressively measurable stock price process S() that satises (2.1), and a process
Z
) of R
n
-valued functions which are absorbed in the cemetery point
and continuous before absorption. We use here n = 1 +K +d +K +d +dK. The
paths in R
n
are the images of Z
i
(t, s) =
_
D
i
log(R(
m
(t, s))) + 1
d
j=1
m
j
(t, s)D
j
log(R(
m
(t, s)))
_
m
i
(t, s)
for all i = 1, . . . , d and (t, s) [0, T] R
d
+
where
m
(, ) denotes the market port-
folio with
m
i
(t, s) := s
i
/
d
j=1
s
j
and R any positive twice dierentiable function
satisfying some weak boundedness conditions. Then, Fernholz (1999) shows that
the pathwise formula
log
_
V
(T)
V
m
(T)
_
= log
_
R(
m
(T, S(T)))
R(
m
(0, S(0)))
_
+
_
T
0
(t, S(t))dt
Chapter 2. The Markovian Case 53
holds where : [0, T] R
d
+
R is some function that can be written down
explicitly. This yields, in connection with Corollary 4, the formula
U
(T, s) =
1
R(
m
(0, S(0)))
E
Q
m
_
R(
m
(T, S(T))) exp
__
T
0
(t, S(t))dt
_
1
1/(Z
(T)V
m
(T))>0
_
,
which can be used to compute optimal trading strategies.
Remark 8 (Perfect balance and optimal growth). Kardaras (2008) discusses in the
case of the market portfolio
i
(t, s) =
m
i
(t, s) := s
i
/
d
j=1
s
j
for all (t, s) [0, T]
R
d
+
and i = 1, . . . , d the perfect balance condition (, ) = a(, )(, ), which
is exactly the mean rate of return appearing in the dynamics of (2.39). If the
perfect balance condition holds under the real-world measure P, then each
component of the market portfolio
m
(, ) is a martingale. If (, ) is not the market
portfolio then this martingale property usually does not hold for the components
of (, ). However, the condition still implies that the strategy (, ) is growth-
optimal in the sense of Problem 4.6 of Fernholz and Karatzas (2009). That means
that (, ) maximizes the mean rate of return
T
(, )(, ) 1/2
T
(, )a(, )(, )
of the logarithm of the associated wealth process over all strategies (, ). More
generally, if
(, ) = a(, )(, ) + (, )c(, ),
for some c : [0, T]R
d
+
R
K
such that the stochastic exponential of
(, ) c(, )
in (2.12) is a martingale, then there is no arbitrage possible with respect to (, ).
This follows directly from the fact that the martingale property implies that P and
Q
()/V
(t) = exp
_
_
t
0
1
S(u) cu
dW(u)
1
2
_
t
0
1
(S(u) cu)
2
du
_
Chapter 2. The Markovian Case 56
for all (t, s) [0, T] R
+
with s > ct. Thus, the reciprocal 1/Z
() of the stochastic
discount factor hits zero exactly when S(t) hits ct. This follows directly from the
Q-dynamics of 1/Z
S(t)
_
0tT
:=
S(t) ct
0tT
is a Brownian motion, we obtain
U
(T t, s) =E
t,s
_
Z
(T)V
(T)
Z
(t)V
(t)
_
=
1
V
(t)
E
Q
_
p(S(T))1
min
tuT
S(u)cu>0
T(t)
S(t)=s
=
1
V
(t)
E
Q
_
exp
_
c
_
S(T)
S(t)
_
c
2
(T t)
2
_
p(S(T))1
min
tuT
S(u)>0
T(t)
_
S(t)=sct
=
1
V
(t)
_
0
exp
_
c(y s + ct)
c
2
(T t)
2
_
p(y + cT)
1
_
2(T t)
_
exp
_
(y s + ct)
2
2(T t)
_
exp
_
(y + s ct)
2
2(T t)
__
dy
(2.43)
=
1
V
(t)
_
_
cTs
Tt
1
2
exp
_
z
2
2
_
p(z
T t + s)dz
exp(2cs 2c
2
t)
_
cT2ct+s
Tt
1
2
exp
_
z
2
2
_
p(z
T t s + 2ct)dz
_
, (2.44)
where we have plugged in the density of a Brownian motion absorbed at zero (com-
pare Problem 2.8.6 of Karatzas and Shreve 1991) and made use of the substitution
z = (y s + cT)/
T t and z = (y + s + cT 2ct)/
T t, respectively.
Chapter 2. The Markovian Case 57
Let us consider the investment in the money market only, to wit, the strategy
0
(, ) 0 and V
0
(t) 1 p(s) for all (t, s) [0, T]R
+
. This yields the hedging
price of one monetary unit
U
0
(T t, s) =
_
s cT
T t
_
exp(2cs 2c
2
t)
_
s cT + 2ct
T t
_
, (2.45)
where denotes the cumulative standard normal distribution function. In the
special case c = 0 we have
U
0
(T t, s) = 2
_
s
T t
_
1. (2.46)
For the rst derivative we obtain
s
U
0
(T t, s) =2c exp(2cs 2c
2
t)
_
s cT + 2ct
T t
_
+
2
(T t)
exp
_
(cT s)
2
2(T t)
_
,
which simplies in the case of c = 0 to
s
U
0
(T t, s) =
2
(T t)
exp
_
s
2
2(T t)
_
. (2.47)
It can be easily checked that U
0
solves the PDE in (2.18) for all (t, s) [0, T] R
+
with s > ct. This is sucient to apply Theorem 1 (compare Remark 5) to nd the
optimal hedging strategy of one monetary unit:
0
(t, s) = s
s
log
_
U
0
(T t, s)
_
. (2.48)
For c = 0 we obtain thus from (2.46) and (2.47) the representation
0
(t, s) =
2
s
Tt
_
s
Tt
_
2
_
s
Tt
_
1
> 0, (2.49)
where denotes the standard normal density.
Chapter 2. The Markovian Case 58
It comes at no surprise that, in order to beat the money market, we have to
be long the stock. The strategy
0
(, ) has for c = 0 another interpretation. To
derive it, we observe that U
0
(T t, s) is the probability that a Brownian motion
W() starting at s does not hit zero before time T t. Using the density of the
hitting time T
0
:= inft 0 :
W(t) = 0, (compare for example Proposition 2.8.5
of Karatzas and Shreve 1991) yields
U
0
(T t, s) =Q(T
0
> T t) =
1
2
_
Tt
s
y
3
2
exp
_
s
2
2y
_
dy,
which gives
s
U
0
(T t, s) =
1
2
_
Tt
_
1
y
3
2
s
2
y
5
2
_
exp
_
s
2
2y
_
dy
and
0
(t, s) =1
1
2
_
Tt
s
3
y
5
2
exp
_
s
2
2y
_
dy
U
0
(T t, s)
.
This is exactly
0
(t, s) = 1 s
2
E
Q
_
1
T
0
min
0uTt
W(u) > 0
_
.
It is well-known that a Bessel process allows for arbitrage. Compare for example
Example 3.6 of Karatzas and Kardaras (2007) for an ad-hoc strategy that corre-
sponds to a hedging price of (1) for a monetary unit if S(0) = T = 1. We have
improved here the existing strategies and found the optimal one, which corresponds
in this setup to a hedging price of U
0
(1, 1) = 2(1) 1 < (1).
Remark 10 (Multiple solutions for the PDE in (2.18)). We observe that the hedging
price U
0
in (2.45) depends on the drift c. Also, U
0
is suciently dierentiable,
thus by Proposition 2 uniquely characterized as the minimal nonnegative solution
of the PDE in (2.18), which does not depend on the drift c. The uniqueness of
U
0
by Proposition 2 and the dependence of U
0
on c do not contradict each other,
Chapter 2. The Markovian Case 59
since the nonnegativity of U
0
has only to hold for these points (t, s) [0, T] R
+
that can be attained by S() at time t. For a given time t [0, T], these are
only the points s > ct. Thus, as c increases, the nonnegativity condition weakens
since it has to hold for fewer points, thus U
0
can become smaller and smaller.
Indeed, plugging in (2.45) the point s = ct yields U
0
(T t, ct) = 0. In summary,
while the PDE itself does only depend on the (more easily observable) volatility
structure of the stock price dynamics, the mean rate of return determines where
the PDE has to hold and thus, contributes to determining the exact amount of
possible arbitrage.
In the next example, we price and hedge a European call within the same
class of models as in the last example:
Example 2 (Three-dimensional Bessel process with drift - stock and European call).
Since we do not know a priori any (possibly suboptimal) strategy that leads to the
value (S(T) L)
+
at time T for some strike L 0 we cannot rely on Theorem 1
and have to tackle this question slightly dierently using the results of Theorem 2.
Plugging in p(y) = (y L)
+
in (2.44), dening
h
p
(T t, s) := E
t,s
_
Z
(T)
Z
(t)
(S(T) L)
+
_
for all (t, s) [0, T] R
d
+
with s > ct, and using the notation a b := maxa, b
we can simplify the expected risk-adjusted value as follows:
h
p
(T t, s) =
_
(cTL)s
Tt
1
2
exp
_
z
2
2
_
(z
T t + s L)dz exp(2cs 2c
2
t)
_
(cTL)2ct+s
Tt
1
2
exp
_
z
2
2
_
(z
T t s + 2ct L)dz
=
_
T t
2
exp
_
(s (cT L))
2
2(T t)
_
+ (s L)
_
s (cT L)
T t
_
exp(2cs 2c
2
t)
_
_
T t
2
exp
_
((cT L) 2ct + s)
2
2(T t)
_
Chapter 2. The Markovian Case 60
+ (2ct s L)
_
(cT L) + 2ct s
T t
__
. (2.50)
The modied put-call parity of Corollary 1 could now be applied to give us directly
the hedging price of a European put. If L cT, in particular if L = 0, the last
expression simplies to
h
p
(T t, s) =s
_
s cT
T t
_
+ exp(2cs 2c
2
t)
_
2ct s cT
T t
_
(s 2ct) LU
0
(T t, s),
where U
0
denotes the hedging price of one monetary unit given in (2.45). It is
simply the dierence between the hedging price of the stock and L monetary units
since L cT implies S(T) > cT L almost surely and the call is always exercised.
Using L = 0 we get the value of the stock.
For L = c = 0, the last equality yields h
p
(t, s) = s for all (t, s) [0, T] R
+
and the stock cannot be arbitraged. There are at least two other ways to see this
result right away. Simple computations show directly that Z
(T) = S(0)/S(T) if
c = 0 and thus, for the strategy
1
(, ) 1, which invests fully in the market,
we obtain U
1
(, ) 1. Alternatively, using the representation of U
1
implied by
(2.33) we see that the hedging price is just the expectation of a Brownian motion
stopped at zero, thus the expectation of a martingale started at one. Every method
on its own shows the lack of relative arbitrage with respect to the market if c = 0.
On the other hand, if c > 0, then relative arbitrage with respect to the
market is possible. In this case, the representation implied by (2.33) shows that as
soon as the Brownian motion is stopped, which is the rst time S(t) equals ct, the
value of the random variable of which the expectation is taken jumps to zero and
thus the stopped process is not a martingale any more. Obviously,
U
1
(T t, s) = E
t,s
_
Z
(T)S(T)
Z
(t)s
_
=
1
s
h
p
(T t, s)
Chapter 2. The Markovian Case 61
is suciently dierentiable and thus, Remark 5 and Theorem 1 yield the optimal
arbitrage opportunity
1
(t, s) =
2ct
Tt
_
scT
Tt
_
+
_
scT
Tt
_
+ exp(2cs 2c
2
t)
_
2ctscT
Tt
_
(2cs 4c
2
t + 1)
U
1
(T t, s)
.
We can now nd the corresponding strategy for the call price of (2.50). Assuming
for the sake of notation that L cT, Theorem 2 yields
p
(t, s) =
_
s
_
s L
T t
_
+ s exp(2cs 2c
2
t)
_
2ct s L
T t
_
(2cs + 2cK 4c
2
t + 1) 2c
T t
_
L + s 2ct
T t
__
/h
p
(T t, s)
as the optimal strategy. If c = 0 this simplies to
p
(t, s) =
s
_
_
sL
Tt
_
+ 1
_
s+L
Tt
__
h
p
(T t, s)
.
Two notable observations can be made. First, in this model both the money
market and the stock simultaneously have a hedging price cheaper than their current
price, as long as c > 0. Second, in contrast to the classical theory of Financial
Mathematics, the mean rate of return under the real-world measure does matter
in determining the hedging price of calls (or other derivatives) since it inuences
the possibilities of arbitrage.
We can now also nd a quantile hedge for strategies (, ) whose associated
wealth process V
U
,
(T, s) := inf v > 0 : strategy such that P
s
(V
v,
(T) > V
(T)) 1 ;
(2.51)
that is,
U
,
(T, s) represents how much initial capital is needed to obtain the termi-
nal wealth V
2T
_
exp
_
(y S(0))
2
2T
_
exp
_
(y + S(0))
2
2T
__
dy
=
T
S(0)
2
_
exp
_
(z S(0))
2
2T
_
exp
_
(z + S(0))
2
2T
__
+
_
S(0) z
T
_
+ 1
_
z + S(0)
T
_
,
where S() is a Brownian motion starting at S(0) under Q
S(0)
and T
0
the rst hitting
of zero by S(). The truncation and the optimal strategy can now be computed
as we have done for calls in Example 2. We omit the computations since they do
not contain any new insights.
Pal and Protter (2010) compute call prices for the reciprocal Bessel process
model. This process has appeared several times in the bubbles literature, often
Chapter 2. The Markovian Case 63
called the constant elasticity of variance (CEV) process; see, for example, Sec-
tion 2.2.2 of Cox and Hobson (2005) or Example 1.2 of Heston et al. (2007). We
discuss next how the results of the last examples relate to this model and illustrate
that even under the NFLVR condition relative arbitrage is possible.
Example 4 (Reciprocal of the three-dimensional Bessel process). Let the stock price
S(t) =
S
2
(t)dW(t)
for all t [0, T] with W() denoting a Brownian motion on its natural ltration F =
F
W
. The process
S() is exactly the reciprocal of the process S() of Examples 1 and
2 with c = 0, thus strictly positive. We observe that there is no classical arbitrage
since the mean rate of return is zero and thus P is already a local martingale
measure. However, there is arbitrage possible with respect to the stock. To wit, if
one wants to hold the stock at time T, one should not buy the stock at time zero,
but use the strategy
1
(, ) below for a hedging price smaller than
S(0) along with
the suboptimal strategy
1
(, ) 1. That is, the strategy
1
(, ) contains a bubble
according to Denition 5.
We have already observed that
S(T) = 1/S(T), which is exactly the stochas-
tic discount factor in Example 1 for c = 0 multiplied by
S(t). Thus, as in (2.46)
the hedging price for the stock is
U
1
(T t, s) = 2
_
1
s
T t
_
1 < 1 (2.52)
along with the optimal strategy
1
(t, s) =
2
_
1
s
Tt
_
s
T t
_
2
_
1
s
Tt
_
1
_ + 1 < 1
for all (t, s) [0, T) R
+
similar to (2.49). By Corollary 4, the hedging price U
1
could also be calculated as one minus the probability of explosion (to ) of the
Chapter 2. The Markovian Case 64
process
S() before time T under Q
1
, where it has the dynamics
d
S(t) =
S
3
(t)dt
S
2
(t)dW
1
(t).
Alternatively, one could also calculate the probability of implosion (to zero)
of the process S() = 1/
1
(t)
for all t [0, T] under Q
1
, which is again a Brownian motion as in Example 1 but
now starting at 1/S(0). For pricing calls, we observe
_
S(T) L
_
+
= L
S(T)
_
1
L
1
S(T)
_
+
=
L
S(t)
S(t)
S(T)
_
1
L
S(T)
_
+
for L > 0. Thus, the price at time t of a call with strike L in the reciprocal Bessel
model is the price of L
S(t) puts with strike 1/L in the Bessel model and can be
computed from Example 2 and Corollary 1. For S(0) = 1, simple computations
will lead directly to Equation (6) of Pal and Protter (2010). The optimal strategy
could now be derived with Theorem 2.
The next example
7
illustrates that U
_
1
S(t)>
1
2
dW(t) if t < 1,
1
S(t)>1
_
1
S(t)1
dt + dW(t)
_
if t 1.
Thus, up to time t = 1 the stock price is either constant or evolves as Brownian
motion stopped at 1/2. If at time t = 1 the stock price is less than or equal to
7
We developed this example after a helpful conversation with Daniel Fernholz.
Chapter 2. The Markovian Case 65
one, it stays constant and otherwise evolves as a three-dimensional Bessel process
shifted by one. From (2.46) the hedging price for the money market is
U
0
(t, s) =
_
_
1 if s 1,
2
_
s1
t
_
1 if s > 1
for t [0, 1]. Then, the hedging price U
0
is not continuous for s = 1, thus not
dierentiable. However, there always exists an optimal strategy
0
. For (t, s)
[0, 1] [0, 1] no arbitrage is possible, which implies that
0
(2 t, s) = 0 is optimal.
For (t, s) [0, 1] (1, ) we know that the stock price always stays above one
and the optimal strategy is the one given in (2.49) with s replaced by s 1 on
the right-hand side. For t (1, 2], the function U
0
(t, s) = E
2t,s
[U
0
(1, S(1))] is
easily shown to be suciently dierentiable. Therefore, we can apply Theorem 1 to
obtain
0
(2t, s). We have illustrated that, although U
0
(t, s) is not dierentiable
in the stock price dimension, namely for (t, s) [0, 1] 1 in this example, there
can nevertheless exist an optimal strategy
0
(, ).
2.7 Conclusion
It has been proven that, under weak technical assumptions, there is no equivalent
local martingale measure needed to nd an optimal hedging strategy based upon
the familiar delta hedge. To ensure its existence, weak sucient conditions have
been introduced that guarantee the dierentiability of an expectation parameterized
over time and over the original market conguration. The dynamics of stochastic
processes simplify after a non-equivalent change of measure and a generalized Bayes
rule has been derived. From an analytic point of view, results of Fernholz and
Karatzas (2010) concerning non-uniqueness of the Cauchy problem of (2.18) have
been generalized to a class of PDEs that allow for a larger set of drifts. With this
newly developed machinery, some optimal trading strategies have been computed
Chapter 2. The Markovian Case 66
addressing standard examples for which so far only ad-hoc and not necessarily
optimal strategies have been known.
2.8 Condition that hedging price solves a PDE
In this section, which serves as an appendix to this chapter, we provide a neces-
sary condition for U
k
(, ) are for all k = 1, . . . , K locally Lipschitz and locally
bounded.
In particular, this assumption restricts the possible strategies (, ); however, it
allows for the market portfolio
m
(, ), for example.
Then, Assumptions (A1), (A1), and (A2) guarantee the necessary smooth-
ness of U
() by
Z
()V
C
1,2
([0, T] R
d
+
) such that
d
i=1
i,k
(t, s)s
i
D
i
H
(t, s) =
k
(t, s) (2.53)
for all k = 1, . . . , K and (t, s) [0, T] R
d
+
. That is, if the covariance process a(t, s)
has a multiplicative inverse a
1
(t, s) on [0, T] R
d
+
, then H
(t, s) =
d
j=1
a
1
i,j
(t, s)
j
(t, s)
i
(t, s)
s
i
. (2.54)
Chapter 2. The Markovian Case 67
This condition basically means that (, ) and (, ) are suciently smooth in time
and space and have an anti-derivative. As Remark 5 discusses, this assumption can
easily be slightly generalized. Applying Itos formula to H
yields
H
(T, S(T)) H
(t, S(t))
_
T
t
_
LH
(u, S(u))
t
H
(u, S(u))
_
du
=
_
T
t
T
(u, S(u))dW(u),
where L is the innitesimal generator of S() dened in (2.22). Collecting all
deterministic terms in a function k
: [0, T] R
d
+
R, we obtain
k
(t, s) :=LH
(t, s) +
t
H
(t, s)
1
2
|
(t, s)|
2
=
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
H
(t, s) +
d
i=1
s
i
i
(s, t)D
i
H
(s, t)
+
t
H
(t, s)
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
i
H
(t, s)D
j
H
(t, s)
=
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
H
(t, s) +
d
i=1
d
j=1
s
i
a
i,j
(t, s)
j
(t, s)D
i
H
(t, s)
+
t
H
(t, s) +
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
D
i
H
(t, s)D
j
H
and
(, ) in (2.53) and
(2.12). Using that, (2.17) can now be written as
U
(t, s) =exp(H
(T t, s))E
Tt,s
_
exp(H
(u, S(u))du
__
.
To proceed, we make the additional assumption that the deterministic function
G
: [0, T] R
d
+
R
+
dened as
G
(t, s) :=exp(H
(T t, s))U
(t, s) (2.56)
=E
Tt,s
_
exp(H
(u, S(u))du
__
,
Chapter 2. The Markovian Case 68
which does not involve a stochastic integral any more, solves the time-inhomogeneous
Cauchy problem
t
G
(t, s) = LG
(t, s) + k
(T t, s)G
). If G
=U
D
i
H
+ exp(H
)D
i
G
,
D
2
i,j
U
=U
D
2
i,j
H
+ U
D
i
H
D
j
H
+ exp(H
)D
i
H
D
j
G
+ exp(H
)D
j
H
D
i
G
+ exp(H
)D
2
i,j
G
.
Therefore, collecting the U
) terms we obtain
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
D
2
i,j
U
+
d
i=1
d
j=1
s
i
a
i,j
j
D
i
U
= U
_
k
t
H
_
+exp(H
)
_
d
i=1
s
i
D
i
G
j=1
a
i,j
(
j
+ s
j
D
j
H
) +
1
2
d
i=1
d
j=1
s
i
s
j
a
i,j
(t, s)D
2
i,j
G
_
= exp(H
)(G
)
_
exp(H
)
t
H
_
G
+ exp(H
)LG
,
where the last equality follows from the identities
d
j=1
j,k
(
j
+ s
j
D
j
H
) =
k
Chapter 2. The Markovian Case 69
and
K
k=1
i,k
k
=
i
for all i, k = 1, . . . , d. That proves the statement since time goes in the reverse
direction.
Chapter 3. Completeness and Relative Arbitrage 70
Chapter 3
Completeness and Relative
Arbitrage
3.1 Introduction
This chapter examines conditions under which contingent claims can be replicated
by dynamic trading in the stock market. Let S() be a continuous, d-dimensional
Ito-process (the stock price process) with respect to a ltration T() and D 0
be an T(T)-measurable random variable (the claim). The question then is when
D can be represented as a stochastic integral of some progressively measurable
process (the trading strategy) with respect to S(). Replicable claims have been
completely characterized if S() satises the No Free Lunch with Vanishing Risk
(NFLVR) condition. This notion was introduced by Delbaen and Schachermayer
(1994) and is equivalent to the existence of an equivalent measure under which S()
is a local martingale. If the supremum over the expected values of D under all
equivalent local martingale measures (ELMMs) is attained, then the claim D can
be replicated.
We generalize this characterization for replicable claims to markets that do
Chapter 3. Completeness and Relative Arbitrage 71
not necessarily satisfy NFLVR, but allow for a stochastic discount factor; this cor-
responds to a weak structural restriction on the drift of the process. Stochastic
discount factors are local martingales and take the place of the Radon-Nikodym
derivatives that are used to change the original measure to an ELMM in the case of
NFLVR. If the supremum over the expected values of D multiplied by all stochastic
discount factors is attained, then the claim D can be replicated.
NFLVR is a mathematical concept introduced to characterize markets that
admit an ELMM, and thus exclude arbitrage opportunities. However, from an eco-
nomic perspective, it is reasonable to consider models which do not satisfy NFLVR.
As Loewenstein and Willard (2000a) and Hugonnier (2010) discuss, models with-
out NFLVR may nevertheless lead to an equilibrium where rational agents have
an optimum. Thus, although NFLVR is a convenient mathematical assumption, it
does not always accurately reect our economic intuition of arbitrage. In partic-
ular, the existence of a stochastic discount factor prevents arbitrage opportunities
from being scaled up, thus allowing for the existence of a numeraire portfolio and
of solutions to utility maximization problems; see Karatzas and Kardaras (2007).
In a similar vein, the theory of real world pricing in the Benchmark
Approach (see Platen and Heath 2006) acknowledges that no ELMM is needed
to have the concept of a price for contingent claims. Models that allow for some
kind of arbitrage are furthermore studied in the framework of Stochastic Portfolio
Theory (see Fernholz 2002; Fernholz and Karatzas 2009), which starts from the
premise of realistically describing the evolution of market weights over long time
horizons and provides simple testable conditions, such as diversity or sucient
intrinsic volatility, under which arbitrage does exist. These insights and ideas lead
to the conclusion that models which impose the existence of a stochastic discount
factor, but do not necessarily additionally assume NFLVR, are a natural class of
models to study. This chapter is therefore a step to close the gap in the theory
Chapter 3. Completeness and Relative Arbitrage 72
between the class of models with and without the assumption of NFLVR.
A nonnegative stock price process S() has been dubbed a bubble, if the set of
ELMMs is nonempty and S() is a strict local martingale under an ELMM. Such a
stock price models an asset that is overpriced compared with its intrinsic value, as
measured by its expectation under an ELMM. It behaves locally like a martingale,
but in the long run, behaves like a strict supermartingale. Academic literature has
recently devoted substantial attention to bubbles, given that they are able to model
seemingly overpriced stocks as in the Internet Bubble within the framework of
NFLVR. We suggest Cox and Hobson (2005), Heston et al. (2007), and Jarrow
et al. (2010) as some initial references to this literature. An asymmetry within
the class of admissible trading strategies is the reason that this phenomenon of
overpriced stock prices appears in models that satisfy NFLVR. Such models allow
for the bond to be sold, but usually do not allow for the stock to be sold in order
to prot from it being overpriced. For this subtle point, we refer the reader to
Yan (1998), where allowable strategies are introduced to avoid the asymmetry
introduced by admissibility constraints. If one is willing to accept the presence of
bubbles, then a natural next step is to allow for some kind of arbitrage, essentially
reecting a bubble in the money market. Such arbitrage arises, for example, after
a change of numeraire with an asset that has a bubble.
Having characterized the claims that can be perfectly replicated, it is a natu-
ral next step to identify the markets in which all claims can be perfectly replicated.
Such a market is then called complete. For markets without arbitrage opportu-
nities, the Second Fundamental Theorem of Asset Pricing (2nd FTAP) gives a
sucient and necessary condition, stating that a market is complete if and only
if the ELMM is unique. This insight regarding the equivalence of the existence of
a replicating strategy for any claim and the uniqueness of a pricing measure can
be traced back to the seminal papers by Harrison and Kreps (1979) and Harrison
Chapter 3. Completeness and Relative Arbitrage 73
and Pliska (1981). For a list of more recent results and references, we point the
reader to Section 1.8 of Karatzas and Shreve (1998). Recently, Lyaso (2010) has
studied completeness in markets where capital gains and additional information to
the investors are modeled separately.
In this chapter, we show that the 2nd FTAP can be extended to markets
that do not proscribe arbitrage. Its generalized version then states that a market
is complete if and only if the stochastic discount factor is unique. Clearly, in the
case of NFLVR, this condition reduces to the classical one since then any stochastic
discount factor generates an ELMM. We conclude that the question regarding the
existence of arbitrage and the question regarding the completeness of the market
can be addressed separately from one another; see also Jarrow et al. (1999) and
Section 10.1 of Fernholz and Karatzas (2009). The proof of the 2nd FTAP in
markets that do not proscribe the existence of arbitrage is simple. It relies on a
change-of-numeraire technique and an application of the classical 2nd FTAP.
Often, however, completeness is too strong a requirement. We instead intro-
duce the notion of quasi-completeness, which only takes into consideration claims
measurable with respect to the stock price ltration. We show that markets whose
drift and diusion components are measurable with respect to the stock price l-
tration are quasi-complete; this generalizes Proposition 1 in Chapter 2, where the
Markovian case is studied.
An important element in Stochastic Portfolio Theory is the concept of (strong)
relative arbitrage. One says that there exists a relative arbitrage opportunity with
respect to some trading strategy () if there exists another trading strategy ()
that outperforms (); that is, if trading according to () yields a higher terminal
wealth than trading according to (). The relative arbitrage is called strong if the
terminal wealth is strictly dominated almost surely. It has been unclear up until
this point whether a relative arbitrage opportunity necessarily implies a strong one.
Chapter 3. Completeness and Relative Arbitrage 74
Having now the characterization of perfect replication and completeness at hand,
we can resolve this question: The existence of relative arbitrage does usually not
imply that of strong relative arbitrage; however, it does if the market is quasi-
complete. We can further state very precise conditions for the existence of both
relative arbitrage and strong relative arbitrage opportunities.
We have included several examples of toy markets that illustrate various
subtle points of our results in the sections that follow. Example 6 demonstrates
that, although a given claim might be measurable with respect to the stock price
ltration, the trading strategy to replicate the claim does not necessarily have
to be measurable with respect to this ltration. Example 7 illustrates that the
drift is important for determining whether a market is quasi-complete or not. In
Example 8, we study a stock price with a bubble whose minimal replicating cost is
not below its current price. Changing this model slightly then yields Example 9,
which treats a model without an ELMM but in which the minimal replicating price
for $1 is again $1. Finally, Example 10 provides the dynamics of a stock price that
implies a strong but diminishing arbitrage opportunity.
We introduce the model and admissible trading strategies in Section 3.2. For
a given claim, we provide necessary and sucient conditions to decide whether it
can be replicated in Section 3.3. In Section 3.4, we state and prove a generalized
version of the 2nd FTAP and discuss the concept of quasi-completeness. We then
apply the tools developed in the previous sections to link relative arbitrage and
strong relative arbitrage in Section 3.5 and we conclude in Section 3.6.
3.2 Setup
This section introduces the probabilistic market model and the concepts of market
prices of risk, stochastic discount factors, trading strategies, and (contingent) claims.
Throughout the chapter, we shall assume that all equalities and inequalities only
Chapter 3. Completeness and Relative Arbitrage 75
hold in an almost-sure sense.
3.2.1 Market model
We x a canonical
1
probability space (, T, P). We assume that = C([0, ), R
K
),
that is, is the space of all continuous functions W() = (W
1
(), . . . , W
K
())
T
taking values in R
K
for some xed K N. Furthermore, we x P so that the
process W() has the law of a K-dimensional Brownian motion. We denote by
F = T(t)
t0
the ltration generated by the paths of W(), and assume it satises
the usual assumptions. We further assume for some xed d N the existence of
a vector of d continuous, adapted processes S() = (S
1
(), . . . , S
d
())
T
with values
in (0, )
d
, which represent the price processes of the risky assets in an economy.
We assume the existence of a K-dimensional, vector-valued process () and of a
dK-dimensional, matrix-valued process (), both progressively measurable with
respect to the underlying ltration F, such that the dynamics of S() can be written
as
dS
i
(t) =S
i
(t)
K
k=1
i,k
(t)
_
k
(t)dt + dW
k
(t)
_
(3.1)
for all t 0 and i = 1, . . . , d. The process () is not assumed to be of rank d or
K. Thus, we do not exclude a-priori stock price models with a non-tradable state
variable, such as stochastic volatility models.
The strict positivity of S() will be guaranteed by imposing S(0) (0, )
d
and the integrability condition
_
T
0
_
|(t)|
2
+
d
i=1
K
k=1
2
i,k
(t)
_
dt < (3.2)
for all T 0.
1
To generalize the results presented here to more general semimartingale models is subject to
future research.
Chapter 3. Completeness and Relative Arbitrage 76
We denote by T
S
() the with the null sets of T augmented, right-continuous
ltration generated by S(). More precisely, we dene T
S
(t) := (S(u), u t) for
all t 0. Since S() is a strong solution of (3.1), we have the inclusion T
S
(t) T(t)
for all t 0.
3.2.2 Market prices of risk and stochastic discount factors
The special structure of the drift is a standard assumption imposed in order to ex-
clude the possibility of an arbitrage opportunity that could otherwise get scaled un-
boundedly; see Section 10 of Karatzas et al. (1991a) and the proof of Theorem 3.5 in
Delbaen and Schachermayer (1995b). We call the process () = (
1
(), . . . ,
K
())
T
in (3.1), which maps the volatility on the drift, a market price of risk.
It is clear that this process is usually not uniquely determined; for example
if the number of rows of () is smaller than the number of columns, that is, if
d < K. In this case, a set of R
K
-valued, F-progressively measurable processes ()
exists such that (3.1) is satised with any () replacing (). To make this precise,
we dene
:=
_
: [0, ) R
K
progressively measurable
_
T
0
|(t)|
2
dt < for all T 0
_
,
t
:=() [ ()() ()() . (3.3)
We call
t
, which as a direct consequence of (3.2) contains (), the set of all
market prices of risk. If any ()
t
replaces () in (3.1), the dynamics of S()
are unchanged. We observe that the stochastic process
m
() dened as
m
() := ()
()(), (3.4)
where denotes the Moore-Penrose pseudo-inverse of a matrix, is again a market
price of risk and therefore is also an element of
t
; see Corollary 1 of Penrose (1955).
Chapter 3. Completeness and Relative Arbitrage 77
For any process () in , we dene Z
(t) := exp
_
_
t
0
T
(u)dW(u)
1
2
_
t
0
|(u)|
2
du
_
(3.5)
for all t 0. For ()
t
, we call Z
i=1
i
(t)dS
i
(t) (3.6)
for all t 0 stays nonnegative.
For any T > 0, we call any nonnegative T(T)-measurable random variable
D 0 a (contingent) claim. A claim represents a certain monetary payo at time
T. Even without the existence of a traded asset S
i
() with S
i
(T) = D for some
i = 1, . . . , d, there might still exist some trading strategy () and some p > 0 such
that V
p,
(T) = D (respectively, V
p,
(T) D), in which case the claim is said to
be replicated (respectively, superreplicated) by ().
Chapter 3. Completeness and Relative Arbitrage 78
Remark 11 (On the admissibility constraint). In the classical theory of Financial
Mathematics, one also has to introduce a notion of admissibility for trading strate-
gies in order to prevent the investor from following the notorious doubling strate-
gies; see the discussion in Section 6 of Harrison and Kreps (1979). Usually, a more
general condition than the nonnegativity of the corresponding wealth process is
assumed. However, any such condition implies that the risk-adjusted wealth pro-
cess Z
()V
p,
() is a supermartingale; see Strasser (2003). This is no longer true
when one abstains from imposing the no-arbitrage condition. For example, if Z
()
is a strict local martingale and the wealth process is only restricted to stay above
some constant < 0, then Z
()V
p,
() is usually no longer a supermartingale.
This motivates the admissibility constraint made here, which mandates that the
wealth process stay nonnegative. We observe that under NFLVR, due to the super-
martingale property, any wealth process of a (super-)replicating strategy for some
nonnegative claim D 0 is again nonnegative, independently from the admissibil-
ity constraint. This fact will be used in the proofs of Section 3.3, where we apply
results of the no-arbitrage theory to obtain a characterization for claims that can
be replicated in markets that do not proscribe arbitrage. We shall revisit these
observations in Remark 12.
3.3 Existence of (super-)replicating trading strate-
gies
Given a specic claim, it is of interest to specify conditions under which its pay-
o can be obtained by means of dynamic trading in the stocks. Theorem 8.5 of
Karatzas et al. (1991b) provides a sucient condition for the replicability of strictly
positive claims. The authors allow for markets that are incomplete, as well as for
markets that admit arbitrage opportunities. Here, we extend their result to more
Chapter 3. Completeness and Relative Arbitrage 79
general volatility matrices () and to claims that are only nonnegative, and fur-
ther provide a minimal superreplicating price for general nonnegative claims. By
relying on duality methods, the question regarding the existence of superreplicating
strategies has been answered in full generality for markets satisfying NFLVR. We
refer the reader to Jacka (1992), Ansel and Stricker (1993), El Karoui and Quenez
(1995), and Delbaen and Schachermayer (1995c) for more on this topic; see also
Kramkov (1996) and Follmer and Kabanov (1998) for a more general class of mod-
els. In the following, we show that these results also extend to markets without an
ELMM.
Throughout this section, we x an horizon T > 0 and an T (T)-measurable
random variable D 0, which represents the claim. We dene
p := D
0
:= sup
()
E[Z
0
:= 0,
n
:= T inf
_
t [0, T]
m
(t) n
_
for all n N, where
m
() has been dened in (3.4) and denotes the minimum.
For any stopping time
n
and any T(
n
)-measurable random variable
D 0, we
set
p
n
(
D) := sup
()
E
_
Z
(
n
)
D
_
. (3.8)
By analogy with p of (3.7), p
n
(
D) can be considered the minimal price for super-
replicating
D at time
n
, as we shall demonstrate below. Towards this end, we
dene
p
e
n
(
D) := sup
()
e
n
E
_
Z
(
n
)
D
_
, (3.9)
Chapter 3. Completeness and Relative Arbitrage 80
where we denote
e
n
:= () [ (
n
)(
n
) (
n
)(
n
), E[Z
(T)] = 1 , = .
for all n N. The next denition is in the spirit of Delbaen and Schachermayer
(1995c):
Denition 8 (Maximal trading strategy). We call a trading strategy () maximal
if the supremum is a maximum in (3.7) with D = V
p,
(T).
Theorem 8 (b) below will motivate the word maximal, since it shows that
no trading strategy which outperforms a maximal one exists. It is already now clear
that () is maximal if some ()
t
exists such that Z
()V
p,
() is a martingale
up to time T.
We can now resolve the question regarding the existence of a (super-)replicating
strategy for claims in models that do not proscribe arbitrage. As the next lemmas
clarify, our argument utilizes the fact that the existence of a square-integrable mar-
ket price of risk guarantees that the market is basically, up to a stopping time, free
of arbitrage. Thus, we shall be able to nd a sequence of time-consistent trading
strategies, which eventually lead, path-by-path, to a superreplicating strategy.
Lemma 5 (Localized (super-)replication). Assume
D 0 is T(
n
)-measurable for
some n N. Then, the equality p := p
n
(
D) = p
e
n
(
D) holds, and the supremum
in (3.8) is attained if and only if it is attained in (3.9). The supremum p is the
minimal superreplicating price for
D at time
n
. More precisely, if p < then an
admissible trading strategy () exists such that
V
p,
(
n
)
D.
If p < , then there exist an T(
n
)-measurable claim
D
D, a trading
strategy (), and a market price of risk ()
e
n
, such that Z
()V
p,
() is a
martingale up to time
n
and V
p,
(
n
) =
D.
Chapter 3. Completeness and Relative Arbitrage 81
Furthermore, no trading strategy () exists for which V
c,
(
n
)
D, for any
c [0, p
n
(
D)).
Proof. First, assume that there exist () and c [0, p
n
(
D)) such that V
c,
(
n
)
D. We observe that Z
()V
c,()
() is a supermartingale for any ()
t
. Thus, we
obtain
E
_
Z
(
n
)
D
_
E[Z
(
n
)V
c,
(
n
)] c < p
n
(
D)
for all ()
t
, which leads to a contradiction with the denition of p
n
(
D) as a
supremum in (3.8).
Now, observe that
Z() Z
m
(
n
) is a martingale since it is bounded
by n. In particular, it denes a new measure Q on T(T), which is equivalent to
P, by dQ/dP =
Z(T). We introduce a ctional market
S() by
S() S(
n
).
Then,
S
i
() is a Q-local martingale for all i 1, . . . d. Thus, NFLVR holds for
the new market. Since the probability space is the canonical one, any measure
Q
on T(T) under which
S() is a local martingale and which is equivalent to P has a
representation d
Q/dP = Z
D, where
V
p
e
n
(
D),
() is dened as in (3.6) with S() replaced by
S(); see Theorem 9
of Delbaen and Schachermayer (1995c). However, we have S(
n
)
S(
n
),
and therefore V
p
e
n
(
D),
(
n
)
D.
Together with the rst part of the proof, where we have shown that any c 0
that satises V
c,
(
n
) 0 for some trading strategy () also satises c p
n
(
D),
this also yields p
e
n
(
D) p
n
(
D). The inequality in the other direction follows from
the fact that for any ()
e
n
there exists ()
t
with Z
(
n
) = Z
(
n
). To
see this, set (t) = (t)1
tn
+
m
(t)1
t>n
for all t 0.
Corollaries 10 and 14 of Delbaen and Schachermayer (1995c) yield the ex-
istence of a claim
D
D, a trading strategy (), and a market price of risk
Chapter 3. Completeness and Relative Arbitrage 82
()
e
n
, such that Z
()V
p,
() is a martingale up to time
n
and V
p,
(
n
) =
D.
Assume now that the supremum in (3.8) is attained, say by ()
t
, but that
the supremum in (3.9) is not. Then, again by Corollaries 14 and 10 of Delbaen
and Schachermayer (1995c), there exists a claim
D
D with P(
D >
D) > 0 and
a trading strategy (), such that V
p,
(
n
) =
D. However, the supermartingale
property of Z
()V
p,
() leads directly to a contradiction.
The next lemma will be of use in Theorem 5, when we need to prove time
consistency of strategies. It generalizes the equality p
n
(
D) = p
e
n
(
D) of Lemma 5.
The measurability of the essential suprema in the following lemma is guaranteed as
in Lemma 7 below.
Lemma 6 (Suciency of local martingale measures). Fix n N. Assume again
that
D 0 is T(
n
)-measurable for some n N. Then, the equality
ess sup
()
E
_
Z
(
n
)
Z
(
n1
)
T(
n1
)
_
= ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
T(
n1
)
_
(3.10)
holds.
Proof. As in the proof of Lemma 5, each ()
e
n
corresponds to some ()
t
.
So, we need only show that the left-hand side is less than or equal to the right-
hand side in (3.10). Towards this end, we x n N and introduce the process
W
Q
() = (W
Q
1
(), . . . , W
Q
K
())
T
with
W
Q
k
() := W
k
() +
_
0
m
k
(u)du
for all k = 1, . . . , K. Analogously to (3.5), we dene
Z
,Q
() := exp
_
_
0
T
(u)dW
Q
(u)
1
2
_
0
|(u)|
2
du
_
and observe
Z
() Z
m
()Z
m
,Q
() (3.11)
Chapter 3. Completeness and Relative Arbitrage 83
for all () . Fix any ()
t
and denote by
i
iN
a sequence of stopping
times dened as
i
:=
n
inf
_
t
n1
m
,Q
(t) iZ
m
,Q
(
n1
)
_
for all i N. The equality in (3.11) and Fatous lemma yield
E
_
Z
(
n
)
Z
(
n1
)
T(
n1
)
_
= E
_
Z
m
(
n
)
Z
m
(
n1
)
Z
m
,Q
(
n
)
Z
m
,Q
(
n1
)
T(
n1
)
_
= E
_
Z
m
(
n
)
Z
m
(
n1
)
lim
i
Z
m
,Q
(
i
)
Z
m
,Q
(
n1
)
T(
n1
)
_
liminf
i
E
_
Z
m
(
n
)
Z
m
(
n1
)
Z
m
,Q
(
i
)
Z
m
,Q
(
n1
)
T(
n1
)
_
= liminf
i
E
_
Z
(i)
(
n
)
Z
(i)
(
n1
)
T(
n1
)
_
ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
T(
n1
)
_
,
since
(i)
:=
_
m
() +1
i
t
_
()
m
()
__
e
n
for all i N; this proves the statement.
We continue by introducing the sequence of random variables
D
n
:= ess sup
()
E
_
Z
(T)
Z
(
n
)
D
T(
n
)
_
=: ess sup
()
n
0 (3.12)
for all n N. If p = D
0
< , then D
n
< for all n N. We discuss in the next
lemma the measurability of each D
n
; in particular, we show that D
n
represents a
claim:
Lemma 7 (Measurability of D
n
). For any n N, the essential supremum D
n
of
(3.12) is T(
n
)-measurable.
Chapter 3. Completeness and Relative Arbitrage 84
Proof. Fix n N. Theorem A.3 of Karatzas and Shreve (1998) yields that D
n
exists
and is T(
n
)-measurable. This is due to the observation that for any
(1)
(),
(2)
()
t
, there exists
(3)
()
t
, which is dened by
(3)
(
n
)
(1)
(
n
) and
(3)
(t) :=
(1)
(t)1
D
(1)
n
D
(2)
n
+
(2)
(t)1
D
(1)
n
<D
(2)
n
(3.13)
for all t >
n
and therefore satises the fork property
D
(3)
n
= D
(1)
n
D
(2)
n
, (3.14)
where denotes the maximum.
The next lemma proves a dynamic programming principle (DPP). It is es-
sential for the results that follow below.
Lemma 8 (Multiplicative DPP). The sequence of random variables (D
n
)
nN
sat-
ises the equalities
D
n1
= ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
D
n
T(
n1
)
_
if D
n
< for all n N.
Proof. Fix n N such that D
n
< . In conjunction with the fork property of
(3.13) and (3.14), Theorem A.3 of Karatzas and Shreve (1998) yields the existence
of a sequence of random variables (
(i)
)
iN
such that D
(i)
n
D
n
as i . Fix
> 0 and dene i
as
i
:= min
_
i N
(i)
n
D
n
_
.
Then, D
(i
)
n
is T(
n
)-measurable and we obtain
ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
D
n
T(
n1
)
_
ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
D
(i
)
n
T(
n1
)
_
+
D
n1
+ .
Chapter 3. Completeness and Relative Arbitrage 85
Since the choice of was arbitrary, this yields one inequality; the other direction
follows from
D
n1
ess sup
()
E
_
Z
(
n
)
Z
(
n1
)
ess sup
()
E
_
Z
(T)
Z
(
n
)
D
T(
n
)
_
T(
n1
)
_
= ess sup
()
E
_
Z
(
n
)
Z
(
n1
)
D
n
T(
n1
)
_
and Lemma 6.
Next, we set p
n
:= p
n
(D
n
) for all n N, as in (3.8). The following time
consistency follows from the same argument as the DPP of Lemma 8: The sequence
(p
n
)
nN
satises
p
n
= p (3.15)
for all n N, with p as in (3.7). We can now state and prove the main result of
this section:
Theorem 5 ((Super-)replicating strategy). There exists no trading strategy ()
for which V
c,
(T) D, for any c [0, p). If p < , then a trading strategy ()
exists such that V
p,
(T) D. If the supremum in (3.7) is attained, then one can
choose () so that V
p,
(T) = D, that is, the claim D can be exactly replicated by
dynamic hedging. If an ELMM exists, then the supremum in (3.7) can be replaced
by the supremum over all ()
t
for which Z
() is a martingale.
Proof. The rst part of the statement follows as in Lemma 5. Assume in the
following that p < , thus D
n
< for all n N. Now, we inductively construct
for each n N trading strategies
(n)
() that satisfy
V
p,
(n)
(
n1
) = V
p,
(n1)
(
n1
) (3.16)
and V
p,
(n)
(
n
) D
n
. According to Lemma 5 and due to (3.15), there exist a
contingent claim
D
1
D
1
, a trading strategy
(1)
(), and a market price of risk
Chapter 3. Completeness and Relative Arbitrage 86
(1)
()
e
1
, such that V
p,
(1)
(
1
) =
D
1
and Z
(1)
()V
p,
(1)
() is a martingale up
to time
1
. Assume that we have determined
(n1)
(),
(n1)
(), and
D
n1
:=
V
p,
(n1)
(
n1
) D
n1
, such that Z
(n1)
()V
p,
(n1)
() is a martingale up to time
n1
. We observe that p
n
(D
n
+
D
n1
D
n1
) = p
n
= p, since by Lemma 5
p p
e
n
(D
n
+
D
n1
D
n1
)
sup
()
e
n
E
_
Z
(
n1
)
_
ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
D
n
T(
n1
)
_
+
D
n1
D
n1
__
sup
()
E
_
Z
(
n1
)
D
n1
_
= p
due to the DPP of Lemma 8. By Lemma 5 again, there exist a contingent claim
D
n
D
n
+
D
n1
D
n1
D
n
, a trading strategy
(n)
(), and a market price of
risk
(n)
()
e
n
such that V
p,
(n)
(
n
) =
D
n
and Z
(n)
()V
p,
(n)
() is a martingale
up to time
n
.
Now, the DPP of Lemma 8 yields
V
p,
(n)
(
n1
) =E
_
Z
(n)
(
n
)
Z
(n)
(
n1
)
V
p,
(n)
(
n
)
T(
n1
)
_
ess sup
()
e
n
E
_
Z
(
n
)
Z
(
n1
)
(D
n
+
D
n1
D
n1
)
T(
n1
)
_
ess sup
()
e
n
_
E
_
Z
(
n
)
Z
(
n1
)
D
n
T(
n1
)
_
+ E
_
Z
(
n
)
Z
(
n1
)
T(
n1
)
_
(
D
n1
D
n1
)
_
=D
n1
+
D
n1
D
n1
=
D
n1
,
and thus V
p,
(n)
(
n1
) V
p,
(n1)
(
n1
). Assume that the event V
p,
(n)
(
n1
) >
V
p,
(n1)
(
n1
) has positive probability. Since Z
(n1)
()V
p,
(n1)
() is a martingale,
this implies that the event V
p,
(n)
(
n1
) < V
p,
(n1)
(
n1
) should also have pos-
Chapter 3. Completeness and Relative Arbitrage 87
itive probability, leading to a contradiction. Thus, this inductive procedure yields
trading strategies
(n)
() that satisfy (3.16) and V
p,
(n)
(
n
) D
n
for all n N.
We dene a new trading strategy () as
(t) =
(1)
(0)1
0
(t) +
n=1
(n)
(t)1
t(
n1
,n]
for all t 0. We observe that
V
p,
(
n
) =
D
n
D
n
holds for all n N. We now x any such that
n()
= T for some n().
Then, we obtain
V
p,
(T)() = V
p,
(
n()
)() D
n()
() = D()
with equality if the supremum in (3.7) is attained, due to the observation that
D
n
= D for all n N in that case. Since for almost all such an n() exists,
() (super-)replicates D.
If an ELMM exists, we are in the context of the classical theory of Finan-
cial Mathematics and then it is sucient to take the supremum in (3.7) over all
ELMMs to obtain the minimal superreplicating price; see Delbaen and Schacher-
mayer (1995c).
The previous theorem proves, in particular, a conjecture in Chapter 2. There,
the Markovian case is discussed and it is demonstrated that the supremum in (3.7)
is always attained, as long as D is measurable with respect to T
S
(), the ltration
generated by the stock price processes S(). For path-independent European-style
claims, an explicit trading strategy for the exact replication is constructed, but the
question of whether path-dependent claims could be hedged is not resolved. For a
more precise statement of these results, see Theorem 6 below.
We wish to draw the readers attention to a few subtle points concerning
the previous theorem. First of all, even if the supremum in (3.7) is not attained,
Chapter 3. Completeness and Relative Arbitrage 88
there might nevertheless exist a trading strategy () that replicates D, that is, a
trading strategy such that V
p,
(T) = D. The claim D = S(2) in Example 8 below
illustrates this point. However, such a trading strategy () is not maximal in the
sense of Denition 8. Second, the replicating price p in (3.7) depends strongly on
the admissibility constraint V
p,
() 0, as the next remark discusses:
Remark 12 (Relevance of the admissibility constraint). We have observed in Re-
mark 11 that the precise choice of the admissibility constraint is not relevant for
determining the costs of replicating a nonnegative claim in markets without arbi-
trage. This is no longer true in markets that do not proscribe arbitrage. Indeed,
if we allow for strategies () whose associated wealth process is only required to
stay above a constant < 0, then the minimal nonnegative price p
to (super-
)replicate a claim D can be computed as
p
:= sup
()
E[Z
(T)(D + )] p.
In particular, it is possible that p
1,1
(t) =
1
S
1
(t)
1
[1,)
(t)
_
1 I
1
1
t>
1
_
,
2,2
(t) =
1
S
2
(t)
1
[1,)
(t)
_
1 I
2
1
t>
2
_
,
for all t 0, where we set
i
:= inf t 0 [ S
i
(t) 1 for i = 1, 2. Thus, up
to time t = 1, the market does not move. Then, one of the stock price processes
has the dynamics of the reciprocal of a three-dimensional Bessel process, while the
other one has the same dynamics only until it hits 1. The sign of W
1
(1) decides
which of the two processes has which dynamics.
We observe that I
1
is not measurable with respect to the stock price ltration
T
S
() up to the stopping time
1
2
> 1. More precisely, for any stopping time
<
1
2
, any event A T
S
( ) is independent of the event W
1
(1) 0; thus
W
1
(1) 0 / T
S
( ).
If ()
t
denotes any market price of risk, then
1
(t) = 1/S
1
(t) for t 1
(t [1,
1
]) if I
1
= 0 (I
1
= 1) and
2
(t) = 1/S
2
(t) for t 1 (t [1,
2
]) if I
2
= 0
(I
2
= 1). We thus obtain from Itos formula
Z
(t) =
S
1
(0)
S
1
(t)
S
2
(0)
S
2
(t)
Z
(1)
2
i=1
_
1 + I
i
_
c
i
(, t
i
, t) 1
_
_
for all t 1 with
c
i
(, t
0
, t
1
) := exp
_
_
t
1
t
0
i
(u)dW
i
(u)
1
2
_
t
1
t
0
2
i
(u)du
_
for all i = 1, 2 and t
0
, t
1
0.
We now x T = 2 and D = 1 and obtain
p = sup
()
E[Z
(2)]
= E
_
S
1
(0)S
2
(0)
S
1
(2)S
2
(2)
_
Chapter 3. Completeness and Relative Arbitrage 90
=
1
2
_
E
_
S
2
(0)
S
2
(2)
W
1
(1) 0
_
+E
_
S
1
(0)
S
1
(2)
W
1
(1) < 0
__
= 2(2) 1,
where denotes the cumulative standard normal distribution function and the last
equality is derived from the expectation of the reciprocal of a three-dimensional pro-
cess, starting at 2; see, for example, (2.46). Furthermore, the supremum is attained,
for example by (). Thus, there exists a trading strategy, () = (
1
(),
2
())
T
,
which exactly replicates D = 1 and which can be explicitly represented as
i
(t) =
2
2 t
(1 I
i
)1
[1,2]
(t)
_
S
i
(t)
2 t
_
for i = 1, 2, where denotes the standard normal density, and the corresponding
(unique) wealth process
V
p,
(t) = 1
[0,1)
(t)p + 1
[1,2]
(t)
2
i=1
(1 I
i
)
_
2
_
S
i
(t)
2 t
_
1
_
for all t [0, 2]; compare (2.49) and (2.46).
We observe that V
p,
( ) depends for all stopping times > 1 on I
1
, thus is
not measurable with respect to the stock price ltration T
S
( ) for all stopping times
(1,
1
2
). Therefore, there exists no trading strategy () that is measurable
with respect to the stock price ltration T
S
() and that replicates D = 1 for initial
costs p.
The last example can easily be adapted to an example for an arbitrage-free
market with a claim that is T
S
(T)-measurable for some T > 1 and that can be
replicated by a maximal T()-measurable trading strategy, but not by a maximal
T
S
()-measurable trading strategy. Towards this end, we introduce a new market
with two stocks
S
i
() := 1/S
i
() for i = 1, 2; both of them are now local martingales,
one of them stopped at
i
. Now, we consider the claim D = S
1
(2)I
2
+ S
2
(2)I
1
. In
order to ensure the measurability of D with respect to T
S
(2), we replace
i
by
i
1.5. Then, we proceed with the argument of the previous example.
Chapter 3. Completeness and Relative Arbitrage 91
3.4 Completeness and Second Fundamental The-
orem of Asset Pricing
In this section, we extend the Second Fundamental Theorem of Asset Pricing to
include markets that do not proscribe arbitrage opportunities. Furthermore, we
discuss two notions of completeness. To start, we formally introduce the concept
of a complete market in the next denition:
Denition 9 (Complete market). A market is called complete if for all T > 0 and
all bounded T(T)-measurable random variables D 0 there exist p > 0 and a
maximal trading strategy () that replicates D; that is, there exists a maximal
trading strategy () such that V
p,
(T) = D. Alternatively, if there exist some
T > 0 and some T(T)-measurable random variable D 0 for which no maximal
replication exists, then the market is called incomplete on [0, T].
In particular, by the martingale representation theorem, a market is complete
if d = K and (t) is invertible for all t > 0; see Section 10.1 of Fernholz and Karatzas
(2009). As previously noted, the notion of completeness is often too strong and we
therefore introduce the notion of quasi-completeness, a slight generalization:
Denition 10 (Quasi-complete market). We call a market quasi-complete if for every
T > 0 and every bounded T
S
(T)-measurable random variable D 0, there exists
a maximal trading strategy () that replicates D.
It follows directly from this denition that any complete market is necessarily
quasi-complete but not vice versa. We call a function g : R
+
C(R
+
, R
d
+
) R
non-anticipative functional if g(t, x()) = g(t, x( t)) for all t R
+
and all x()
C(R
+
, R
d
+
); that is g is non-anticipative if g(t, x()) depends on the path of x only
up to time t. We have the following result, which generalizes Pag`es (1987), Due
(1988)
2
, and Proposition 1 in Chapter 2:
2
We thank Martin Schweizer for pointing us to this reference.
Chapter 3. Completeness and Relative Arbitrage 92
Theorem 6 (Sucient conditions for quasi-completeness). If S() of (3.1) can be
represented as the unique solution of
dS
i
(t) =S
i
(t)
K
k=1
i,k
(t, S())
_
k
(t, S())dt + dW
k
(t)
_
,
where
i,k
and
k
are non-anticipative functionals for all i = 1, , d and k =
1, , K, then the market is quasi-complete. Furthermore, for any T > 0 and
T
S
(T)-measurable random variable D 0,
m
() maximizes the expression in (3.7).
Proof. The proof of Proposition 1 in Chapter 2 carries through with only minor
modications.
We emphasize that we have not assumed that the volatility matrix ()
has full rank in the previous theorem. As demonstrated in the next example,
an incomplete market is generally not quasi-complete if
m
() is not progressively
measurable with respect to T
S
():
Example 7 (Relevance of drift for quasi-completeness). We set K = 1, d = 1,
S(0) = 1, (t) = 0,
m
(t) = 0 for all t [0, 1], and (t) = 1/S(t),
m
(t) =
1
S(t)
t1
[W(1)[
for all t > 1. This market is a slight extension of Example 1 in Chapter 2. We con-
sider D = 1, which is T
S
(2)-measurable. For any ()
t
, where
t
is introduced
in (3.3), we obtain
E
_
Z
(2)
Z
(1)
D
T(1)
_
=
_
(1 c) exp(2c)(1 (1 + c))
_
c=
1
|W(1)|
,
where denotes the standard normal cumulative distribution function; see (2.45).
The last function is decreasing in c 0. We thus obtain
sup
()
E[Z
(2)D] = 2(1) 1;
however, the supremum is not attained. Thus, the model is not quasi-complete.
Chapter 3. Completeness and Relative Arbitrage 93
The stock price process in the previous example is sometimes called a bub-
ble, since under the ELMM its price tends to decrease in expectation due to its
strict local martingality. We refer the reader to Jarrow et al. (2010) for a deni-
tion, further references, and a thorough discussion regarding bubbles in incomplete
markets. The next lemma prepares the proof of the Second Fundamental Theorem
of Asset Pricing:
Lemma 9 (Rank of volatility matrix in complete market). If a market is complete,
then rank((t)) = K Lebesgue-almost everywhere. In particular, d K.
Proof. Fix some T > 0 and assume a complete market. We now show rank((t)) =
K Lebesgue-almost everywhere on [0, T]. We introduce a new, ctional market
with d + 1 stocks
_
S
1
()
V
p,
()
, . . . ,
S
d
()
V
p,
()
,
1
V
p,
()
_
,
where p is dened in (3.7) with D = 1 and () is the corresponding maximal trading
strategy, as for example determined in the proof of Theorem 5, such that V
p,
(T)
1. Then, Theorems 11 and 4 of Delbaen and Schachermayer (1995c) yield together
that NFLVR holds for the new market. If we denote the volatility matrix of the
new market by (), then a simple computation shows that rank(()) rank( ()),
and hence, that the new market is also complete. Although we have not assumed
d + 1 K, the argument of Theorem 1.6.6 in Karatzas and Shreve (1998) works
and proves the result.
We can now formulate and prove the Generalized Second Fundamental The-
orem of Asset Pricing:
Theorem 7 (Generalized Second Fundamental Theorem of Asset Pricing). A mar-
ket is complete if and only if any process ()
t
satises (t) =
m
(t) Lebesgue-
almost everywhere.
Chapter 3. Completeness and Relative Arbitrage 94
Proof. If the market is complete, then we have rank((t)) = K Lebesgue-almost
everywhere by Lemma 9. This is equivalent to the Lebesgue-almost everywhere
uniqueness of () in
t
. For the reverse direction, we observe that the supremum
in (3.7) is always taken over a singleton, and is thus trivially attained.
We remark that any complete market implies Z
() Z
m
() for all ()
t
.
Thus, in the no-arbitrage framework, this directly translates into the uniqueness
of the ELMM. However, it is important to note that the question regarding the
completeness of the market can be addressed separately from the question regarding
the existence of arbitrage; see also Jarrow et al. (1999).
3.5 Relative arbitrage and strong relative arbi-
trage
In this section, we analyze the interplay of relative arbitrage and strong relative
arbitrage opportunities. The concept of relative arbitrage traces back to Merton
(1973), where the term dominant portfolio is used. He writes:
Security (portfolio) A is dominant over security (portfolio) B, if on
some known date on the future, the return on A will exceed the return
on B for some possible states of the world, and will be at least as large
as on B, in all possible states of the world.
We also refer to Jarrow et al. (2007; 2010) for a thorough discussion of Mertons
no-dominance principle in connection with the existence of bubbles. Delbaen and
Schachermayer (1994; 1995c) coined the term maximal element for a terminal
wealth V
p,
(T) that cannot be dominated by another terminal wealth V
p,
(T). In
the following, we use the terminology of Stochastic Portfolio Theory; see Fernholz
Chapter 3. Completeness and Relative Arbitrage 95
and Karatzas (2009). This line of research does not focus on nding the right con-
ditions to exclude arbitrage opportunities, but instead studies these opportunities;
see, for example, Fernholz and Karatzas (2010), where relative arbitrage with re-
spect to the market portfolio is studied. We now provide the precise denition on
which we shall rely:
Denition 11 (Relative and classical arbitrage). We say that there exists relative
arbitrage with respect to a trading strategy () over the time horizon [0, T] if
there exists a trading strategy () such that P(V
p,
(T) V
p,
(T)) = 1 and
P(V
p,
(T) > V
p,
(T)) > 0. We say that () is a strong relative arbitrage if
P(V
p,
(T) > V
p,
(T)) = 1. If () 0, which corresponds to holding the risk-free
money market, then we sometimes substitute the word relative by classical.
Obviously, the existence of strong relative arbitrage necessarily implies that
of relative arbitrage. However, the converse is less obvious. Using the insights
developed in the previous sections, we shall discuss conditions under which the
existence of relative arbitrage implies that of strong relative arbitrage in Theorem 8.
We start by giving an example showing that this implication does not always hold:
Example 8 (Relative arbitrage without strong relative arbitrage). Let K = 1, d = 1,
S(0) = 2, () 0 and (t) = 0 for t [0, 1] [2, ). Set
(t) =
1
S(t)
1
W(1)0>t
1
2 t
for t (1, 2), where
:= inf
_
t 1 :
_
t
1
1
2 s
dW(s) = 1
_
. (3.17)
Then we have < 2, which yields S(2) = 2 on the event W(1) < 0, S(2) = 1 on
the event W(1) 0, and S() being a strictly positive, local martingale.
We consider the buy-and-hold trading strategy () 1, such that D :=
V
2,
(2) = S(2). Since NFLVR is satised here, it is sucient to take the supremum
Chapter 3. Completeness and Relative Arbitrage 96
in (3.7) over
:= ()
t
: E[Z
(2)] = 1 ,
to wit, the subset of
t
that generates the ELMMs. For any ()
we have
Q
is dened by dQ
/dP = Z
E[Z
(2)S(2)] = 2 inf
()
(W(1) 0) = 2.
That is, the cheapest trading strategy to superreplicate one share S(2) at time
T = 2 costs p = 2. Fix any trading strategy (). Then, on the event W(1) < 0
we always have V
2,
(2) = 2 = S(2). This shows that no strong relative arbitrage
exists with respect to () over the time horizon [0, 2].
However, relative arbitrage exists. The trading strategy () 0 yields
V
2,
(2) = 2. Thus, P(V
2,
(2) > S(2)) = P(W(1) 0) = 1/2 > 0. To conclude,
although the cheapest superreplicating price of a given terminal wealth V
p,
(T)
might be p, the trading strategy () might nevertheless be dominated in the sense
of Merton (1973).
Delbaen and Schachermayer (1998) discuss a model in which the stock price
process is a strict local martingale under one measure, but actually a true martingale
under an equivalent measure. The previous example exhibits a stock price process
such that Z
S(t) =
S(t)(t)
_
(t)dt dW
t
_
with () as in Example 8. Corollary 15 of Delbaen and Schachermayer (1995c) di-
rectly yields that this market does not allow for an ELMM. Any stochastic discount
factor
Z
() in the new model can be written as
Z
() = Z
()S()/S(0), where Z
()
denotes a stochastic discount factor in the original model of Example 8.
We now set T = 2 and consider the claim D = 1, which corresponds to
holding exactly $1 at time 2. and obtain sup
()
E[
Z
(2)D] = 1. Thus, despite
the existence of arbitrage opportunities, the cheapest price to hold $1 is again $1 and
no strong classical arbitrage exists, due to reasoning similar to that in Example 8.
However, starting with $1, one can achieve a terminal wealth that is larger than $1
with positive probability by following the trading strategy () 1.
We can now state precise conditions for the existence of relative arbitrage
and strong relative arbitrage opportunities:
Theorem 8 (Conditions for relative arbitrage and strong relative arbitrage). Fix
T > 0 and a trading strategy () admissible for some initial capital p > 0.
(a) There exists a strong relative arbitrage opportunity with respect to () over
the time horizon [0, T] if
p := sup
()
E[Z
(T)V
p,
(T)] < p. (3.18)
The converse holds if a trading strategy () and a constant > 1 exist such
that V
p,
(T) V
p,
(T) ,= 0.
(b) There exists a relative arbitrage opportunity with respect to () over the time
horizon [0, T], if and only if
E[Z
(T)V
p,
(T)] < p (3.19)
Chapter 3. Completeness and Relative Arbitrage 98
for all ()
t
.
(c) In particular, the existence of relative arbitrage implies that of strong relative
arbitrage over the time horizon [0, T] if the market is quasi-complete and
V
p,
(T) is T
S
(T)-measurable.
Proof. We prove (a), (b), and (c) separately:
(a) Assume (3.18) holds. According to Theorem 5, a trading strategy () exists
such that
V
p,
(T) V
p,
(T) + p p > V
p,
(T),
which shows the existence of strong relative arbitrage.
We observe that for any ()
t
and for any trading strategy (), ad-
missible with respect to the initial capital p, the process Z
()V
p,
() is a
supermartingale. Thus, if a strong relative arbitrage () and some > 1 as
in the statement of the theorem exist, then
sup
()
E[Z
(T)V
p,
(T)] sup
()
E[Z
(T)V
p,
(T)]
1
< p,
which implies (3.18).
(b) In a similar vein, assume that a relative arbitrage () with respect to ()
exists. Then,
E[Z
(T)V
p,
(T)] < E[Z
(T)V
p,
(T)] p.
for all ()
t
, yielding (3.19). For the reverse direction, let us introduce,
as in Lemma 9, a ctional market with d + 1 stocks
_
S
1
()
V
p,
()
, . . . ,
S
d
()
V
p,
()
,
1
V
p,
()
_
.
Chapter 3. Completeness and Relative Arbitrage 99
Then, (3.19) yields that no ELMM exists for the new market. Thus, the
ctional market allows for classical arbitrage and Theorems 4 and 11 of Del-
baen and Schachermayer (1995c) yield the existence of a relative arbitrage
opportunity.
(c) If the market is quasi-complete, then the supremum in (3.18) is always a max-
imum, and consequently, relative arbitrage implies strong relative arbitrage
in the case of quasi-completeness.
The next example illustrates the fact that p = p in (3.18) does not necessarily
exclude a strong relative arbitrage opportunity:
Example 10 (Diminishing strong relative arbitrage). We use the same setting as in
Example 8 with () modied as follows:
(t) =
1
S(t)
i=1
1
i
1
[W(1)[[i1,i)>t
1
2 t
for t (1, 2), where the stopping is dened as in (3.17). This yields S(2) = 11/i
on the event [W(1)[ [i 1, i) for all i N. We obtain
p = sup
()
E[Z
(2)S(2)] = 2 inf
()
i=1
1
i
Q
([W(1)[ [i 1, i))
_
= 2,
where
and Q
Ete de Probabilites
de Saint-Flour XII-1982, pages 143303. Springer.
Levental, S. and Skorohod, A. (1995). A necessary and sucient condition for ab-
sence of arbitrage with tame portfolios. Annals of Applied Probability, 5(4):906
925.
Loewenstein, M. and Willard, G. A. (2000a). Local martingales, arbitrage, and
viability. Free snacks and cheap thrills. Economic Theory, 16(1):135161.
Loewenstein, M. and Willard, G. A. (2000b). Rational equilibrium asset-pricing
bubbles in continuous trading models. Journal of Economic Theory, 91(1):17
58.
Lyaso, A. (2010). The FTAP in the special case of Ito process nancial markets.
Madan, D. and Yor, M. (2006). Itos integrated formula for strict local martingales.
In Seminaire de Probabilites, XXXIX, pages 157170. Springer.
Merton, R. C. (1973). Theory of rational option pricing. Bell Journal of Economics,
4(1):141183.
Meyer, P. (1972). La mesure de H. Follmer en theorie de surmartingales. In
Seminaire de Probabilites, VI. Springer.
BIBLIOGRAPHY 106
Mijatovic, A. and Urusov, M. (2009). On the martingale property of certain local
martingales.
Novikov, A. (1972). On an identity for stochastic integrals. Theory of Probability
and its Applications, 17(4):717720.
Pag`es, H. (1987). Optimal consumption and portfolio policies when markets are
incomplete.
Pal, S. and Protter, P. E. (2010). Analysis of continuous strict local martingales
via h-transforms. Stochastic Processes and their Applications, 120(8):14241443.
Parthasarathy, K. (1967). Probability Measures on Metric Spaces. Academic Press.
Penrose, R. (1955). A generalized inverse for matrices. Proceedings of the Cambridge
Philosophical Society, 51:406413.
Platen, E. (2002). Arbitrage in continuous complete markets. Advances in Applied
Probability, 34(3):540558.
Platen, E. (2006). A benchmark approach to nance. Mathematical Finance,
16(1):131151.
Platen, E. (2008). The law of minimum price.
Platen, E. and Heath, D. (2006). A Benchmark Approach to Quantitative Finance.
Springer.
Platen, E. and Hulley, H. (2008). Hedging for the long run.
Protter, P. E. (2003). Stochastic Integration and Dierential Equations. Springer,
2nd edition.
Revuz, D. and Yor, M. (1999). Continuous Martingales and Brownian Motion.
Springer, 3rd edition.
Ruf, J. (2011+). Hedging under arbitrage. Mathematical Finance, forthcoming.
Schweizer, M. (1992). Martingale densities for general asset prices. Journal of
Mathematical Economics, 21:363378.
Shreve, S. E. (2004). Stochastic Calculus for Finance II. Continuous-Time Models.
Springer.
Sin, C. (1998). Complications with stochastic volatility models. Advances in Applied
Probability, 30(1):256268.
BIBLIOGRAPHY 107
Strasser, E. (2003). Necessary and sucient conditions for the supermartingale
property of a stochastic integral with respect to a local martingale. In Seminaire
de Probabilites, XXXVII, pages 385393. Springer.
Stroock, D. W. and Varadhan, S. R. S. (2006). Multidimensional Diusion Pro-
cesses. Springer, Berlin. Reprint of the 1997 edition.
Wong, B. and Heyde, C. C. (2004). On the martingale property of stochastic
exponentials. Journal of Applied Probability, 41(3):654664.
Yan, J.-A. (1998). A new look at the Fundamental Theorem of Asset Pricing.
Journal of the Korean Mathematical Society, 35(3):659673.