Finance Stochast. 4, 443–463 (2000)
c Springer-Verlag 2000
Game options
Yuri Kifer
Institute of Mathematics, The Hebrew University, Givat Ram 91904 Jerusalem, Israel
(e-mail:
[email protected])
Abstract. I introduce and study new derivative securities which I call game
options (or Israeli options to put them in line with American, European, Asian,
Russian etc. ones). These are contracts which enable both their buyer and seller
to stop them at any time and then the buyer can exercise the right to buy (call
option) or to sell (put option) a specified security for certain agreed price. If the
contract is terminated by the seller he must pay certain penalty to the buyer.
A more general case of game contingent claims is considered. The analysis is
based on the theory of optimal stopping games (Dynkin’s games). Game options
can be sold cheaper than usual American options and their introduction could
diversify financial markets.
Key words: American option pricing, optimal stopping game
JEL Classification: G13, C73
Mathematics Subject Classification (1991): 90A09, 60J40, 90D15
1 Introduction
A standard (B , S )-securities market consists of a nonrandom (riskless) component
Bt , which is described as a savings account (or price of a bond) at time t with
an interest r, and of a random (risky) component St , which can be described
as the price of a stock at time t. Both discrete time t ∈ Z+ = {0, 1, 2, . . . }
Dedicated to E.B.Dynkin on his 75th birthday
Partially supported by the Edmund Landau Center for Research in Mathematical Analysis and Related
Areas, sponsored by the Minerva Foundation (Germany).
Manuscript received: June 1999; final version received: November 1999
444
Y. Kifer
and continuous time t ∈ R+ = {t ≥ 0} models can be considered. A standard
American option is a contract which enables its buyer to exercise it, i.e. to sell
(put option) or to buy (call option) the stock for a specific price K , at any
time t which amounts to the gain (K − St )+ in the put and (St − K )+ in the
call option cases. The problem of fair pricing of American options leads to the
optimal stopping of certain stochastic processes (see [Ka1,2], [My], [SKKM1,2]
and references there). In this paper I introduce game options in which the seller
of an option can cancel the contract at any time t. In this case the buyer’s gain is
the sum (K − St )+ + δt in the put and (St − K )+ + δt in the call option case where
δt ≥ 0 is certain penalty paid by the seller. The pricing of these options leads to
a game version of the optimal stopping problem introduced in the discrete time
case by Dynkin [Dy] (for a continuous time version see [Ki]) but for financial
applications it is more appropriate to employ another more general set up studied
in [Ne] and [Oh] in the discrete time case and considered in the continuous time
case in [BF1] (Markov case) and in [LM] (general case).
The formal set up consists of a probability space (Ω, F , P ) together with a
stochastic process St ≥ 0, t ∈ Z+ , or t ∈ R+ describing the price of a unit of
stock, of a family of complete σ-algebras Ft ⊂ F such that Ft is generated by
all Su , 0 ≤ u ≤ t, and of two right continuous with left limits stochastic payoff
processes Xt ≥ Yt ≥ 0 adapted to the filtration {Ft , t ∈ Z+ or t ∈ R+ }.
A game contingent claim (GCC) is a contract between a seller A and a buyer
B which enable A to cancel (terminate) it and B to exercise it at any time t
up to a maturity date (horizon) T when the contract is terminated anyway. If
B exercises the contract at time t then he gets from A the payment Yt but if A
cancels before B exercises then A should pay to B the sum Xt . If A cancels and
B exercises at the same time t then A pays to B the sum Yt . It turns out (see
Remarks 2.2 and 3.4) that if, instead, in the latter case A pays to B in the amount
Xt all results remain the same provided there is no penalty at maturity date.
Assuming that clairvoyance is not possible A and B have to use only stopping
times with respect to the filtration {Ft } as their cancellation and exercise times.
The difference δt = Xt − Yt ≥ 0 is interpreted as the penalty which A pays to B
for the contract cancellation.
What is the fair price V ∗ that B should pay to A for such contract? In
accordance with the modern ideology of option pricing based on hedging it is
natural to require that V ∗ should be the minimal capital which enables A to
invest it into a skillfully managed self-financing portfolio which will cover his
liability to pay to B up to a cancellation stopping time σ no matter what exercise
time B chooses. Namely, for any initial capital Z0 > V ∗ the seller A should be
able to choose a stopping time σ and to manage a self-financing portfolio having
a wealth Zt at time t and being redistributed in discrete times or continuously
between the savings account and the stock shares so that Zt is sufficient for
payment to B provided the latter exercises GCC at the time t ∈ [0, σ]. Thus
hedging in GCC consists in a choice of both a hedging investment policy and of
a cancellation time of the contract. I shall show that this leads to the zero sum
optimal stopping game of two players with the payoffs e −rt Xt and e −rt Yt .
Game options
445
If A is not allowed to terminate the contract before the maturity time T then
we arrive at an American Contingent Claim (ACC). The same can be achieved in
the framework of my model if the penalty is chosen large enough, for instance,
if Xt = Yt + δ and δ > sup EYτ . More precise conditions which ensure that
0≤τ ≤T
GCC becomes ACC can be given, as well. On the other hand, I could modify the
above model so that B is not allowed to terminate the contract until the maturity
date T in the spirit of European (game) options. This also can be considered in
the framework of my model if I take Yt = 0 for t < T and Yt = YT > 0 for
t = T . Observe, that if the penalty δ0 is zero then either A or B should terminate
the contract at once and the price V ∗ equals Y0 . It follows from Theorems 2.1
and 3.1 that the price V ∗ is a continuous increasing function of penalty which
varies, thus, from Y0 to sup0≤τ ≤T E (e −rτ Yτ ).
In the next section I consider the discrete time case where the stock evolution
is described by the popular binomial CRR-model introduced in [CRR]. In Sect. 3
I deal with the continuous time situation where the stock evolution is described
by the geometric Brownian motion. The payoff functions Xt and Yt are supposed
to be right continuous and having left limits. In particular, one can take Yt =
(K − St )+ or Yt = (St − K )+ and Xt = Yt + δt . These cases are naturally to be
denominated put and call game options, respectively, with a penalty process δt ,
t ≥ 0. Other payoff functions leading to exotic game options can be considered,
as well. In Sect. 4 I discuss the case when Yt and Xt have the form β t Y (St ) and
β t X (St ), β ≤ 1.
In this paper I consider only basic problems concerning extension of the option pricing theory to game options and many problems still remain to deal with.
First one can consider a multidimensional case of several stocks which can be
treated in the same way. Incomplete markets also can be treated in my framework employing optional decompositions of supermartingales and superhedging
similarly to [Kr]. Next, the model may include transaction costs, portfolio constraints and uncertainty (random environments) which are important in real stock
exchange trading but, of course, complicate the study. Furthermore, it is important for applications to find convenient formulas and algorithms for computation
of prices of game options. For the binomial CRR-model computations are not
difficult, especially in the Markov case considered in Sect. 4 using recursive formulas there. On the other hand, the geometric Brownian motion model leads to
quite nontrivial free boundary problems and variational inequalities but I show
in Sects. 3 and 4 that discrete time approximations of continuous time models
yield methods for their computation.
Essentially, any contract in modern life presumes explicitly or implicitly
a cancellation option by each side which then has to pay some penalty, and
so it is natural to introduce a buyback option to contingent claims, as well.
Moreover, already existing corporate bonds which are both putable by the buyer
and callable by the issuer give an example of GCC but, usually, they are more
difficult to evaluate than GCC’s considered in this paper since such bonds include
various restrictions on possible exercise times and they may depend on other
446
Y. Kifer
value processes of underlying securities. Game options are safer for an investment
company which issues them, and so it can sell them cheaper than usual American
options. Recent problems of hedge funds in some emerging markets may justify
introduction of such derivative securities with a buyback option as this may
protect the issuing company from collapsing and, on the other hand, to ensure
some compensation for the buyer. Moreover, this provides the whole range of
GCC depending on the agreed penalty: from high penalty and high price GCC
which, actually, coincide with usual American contingent claims to low penalty
and low price GCC which provide only little protection for the buyer. In addition,
GCCs contain some elements of games of chance and require more advanced
trading techniques which may be attractive for some investors who could exploit
their more sophisticated skills. Introduction of GCCs could help also to diversify
financial markets. As a market name for such contracts I suggest to call them
Israeli contingent claims (Israeli options) to put them in line with American,
Asian, European, Russian etc. ones.
After this paper was submitted my attention was drawn to [MS] where a
particular case of a GCC was discussed on a heuristic level without indicating
any connection to Dynkin’s games and with a different motivation of introducing
a model which can be considered as a simplified version of a Liquid Yield Option
Note (LYON) which was really traded on markets. Karatzas informed me recently
that a pre-publication version of [CK] contained a remark indicating on the
possibility of continuous time GCCs, suggesting their connection with backward
stochastic differential equations with two reflecting barriers, and asserting that
their price is given by the value of a Dynkin game. This remark does not appear in
the printed version of [CK] and neither version contains any precise definitions
and proofs concerning financial applications of results from [CK]. Moreover,
they seem to had in mind the dual approach to the price process which should
dominate possible payoffs at any time, and so if the price of a GCC is defined
in a natural way via hedging, as I do in this paper, then the results of [CK] do
not imply directly the price formula obtained here. In addition, the approach of
[CK] relying on results about backward stochastic differential equations is more
complex than the arguments considered here, it requires also stronger assumptions
on payoff processes Xt , Yt and, of course, it does not work in the discrete time
case. A preprint version of the present paper has appeared at the end of 1997
though I discussed with several people the idea of game options a couple of
years earlier. All rights on commercial use of game (Israeli) contingent claims
considered in this paper are reserved with the author.
2 Discrete time
Let Ω = {1, −1}N be the space of finite sequences ω = (ω1 , ω2 , . . . , ωN ); ωi = 1
or = −1 with the product
probability
P = {p, q}N , q = 1 − p so that p(ω) =
N
p k q N −k where k = 12 N + ωi . In this section I consider the CRR-model of
i =1
Game options
447
financial market which functions at times n = 0, 1, . . . , N < ∞ and consists of
a savings account Bn with an interest rate r, so that
Bn = (1 + r)n B0 ,
B0 > 0,
r > 0,
(2.1)
and of a stock whose price at time n equals
Sn = S0
n
(1 + ρk ),
S0 > 0,
(2.2)
k =1
where ρk (ω) = 21 (a + b + ωk (b − a)), ω = (ω1 , ω2 , . . . , ωN ).
Thus the “random growth rates” ρk , k = 1, . . . , N form a sequence of independent identically distributed random variables on the probability space (Ω, P )
taking values a and b with probabilities q and p, respectively. As usual, I assume
−1 < a < r < b,
0 < p < 1.
(2.3)
Introduce also the (finite) σ-algebras Fn , n = 0, 1, . . . , N where F0 = {∅, Ω} and
Fn , n = 1, 2, . . . , N is generated by the random variables {ρk , k = 1, . . . , n}.
Recall, (see, for instance, [SKKM1]) that a portfolio strategy π with an initial
capital Z0π = z > 0 and a horizon N is a sequence π = (π1 , . . . , πN ) of pairs
πn = (βn , γn ) where βn , γn are Fn−1 -measurable random variables representing
the number of units on the savings account and of the stock, respectively, at time
n so that the price of the portfolio at time n is given by the formula
Znπ = βn Bn + γn Sn .
(2.4)
A portfolio strategy π is called self-financing if all changes in the portfolio value
are due to capital gains or losses but not to withdrawal or infusion of funds. This
means that (see [SKKM1]), β1 B0 + γ1 S0 = z and for all n > 1,
Bn−1 (βn − βn−1 ) + Sn−1 (γn − γn−1 ) = 0.
(2.5)
Denote by JnN the finite set of stopping times ξ with respect to the filtration
{Fn }0≤n≤N (i.e. {ω : ξ(ω) ≤ k } ∈ Fk , k = n, . . . , N ) with values in {n, n +
1, n + 2, . . . , N }.
A Game Contingent Claim (GCC) is a contract between investors A and
B consisting of a maturity date N < ∞, of selection of a cancellation time
σ ∈ J0N by A, of selection of an exercise time τ ∈ J0N by B and of Fn adapted payoff processes ∞ > Xn ≥ Yn ≥ 0, so that A pledges to pay to B at
time σ ∧ τ = min(σ, τ ) the sum
def
R(σ, τ ) = Xσ Iσ<τ + Yτ Iτ ≤σ
(2.6)
where IQ = 1 if an event Q occurs and = 0 if not. It turns out (see Remark 2.2
below) that if I replace in (2.6) Iσ<τ by Iσ≤τ and Iτ ≤σ by Iτ <σ then the results
below remain the same provided XN = YN .
448
Y. Kifer
A hedge against a GCC with a maturity date N is a pair (σ, π) of a stopping
π
time σ ∈ J0N and a self-financing portfolio strategy π such that Zσ∧n
≥ R(σ, n)
for all n = 0, 1, . . . , N .
The fair price V ∗ of a GCC is the infimum of V ≥ 0 such that there exists
a hedge (σ, π) against this GCC with Z0π = V .
Theorem 2.1 Let P ∗ = {p ∗ , 1 − p ∗ }N be the probability on the space Ω with
r−a
, N < ∞ and E ∗ denotes the corresponding expectation. Then the fair
p ∗ = b−a
∗
price V ∗ of the above GCC equals V0N
which can be obtained from the recursive
−N
∗
relations VNN = (1 + r) YN and for n = 0, 1, . . . , N − 1
∗
∗
= min((1 + r)−n Xn , max((1 + r)−n Yn , E ∗ (Vn+1N
|Fn ))).
VnN
(2.7)
Moreover, for n = 0, 1, . . . , N ,
∗
VnN
= min max E (1 + r)
R(σ, τ )Fn
σ∈JnN τ ∈JnN
∗
−σ∧τ
= max min E (1 + r)
R(σ, τ )Fn .
∗
−σ∧τ
(2.8)
τ ∈JnN σ∈JnN
Furthermore, for each n = 0, 1, . . . , N the stopping times
∗
∗
σnN
or k = N } and
= min{k ≥ n : (1 + r)−k Xk = VkN
(2.9)
∗
∗
τnN
= min{k ≥ n : (1 + r)−k Yk = VkN
}
∗
belong to JnN (since VNN
= (1 + r)−N YN ) and they satisfy
∗
∗
∗
−σnN
∧τ
∗
∗
∗
−σ∧τnN
∗
E (1 + r)
R(σnN , τ )Fn ≤ VnN ≤ E (1 + r)
R(σ, τnN )Fn
(2.10)
for any σ, τ ∈ Jn,N . Finally, there exists a self-financing portfolio strategy π ∗
∗
∗
∗
such that (σ0N
, π ∗ ) is a hedge against this GCC with the initial capital Z0π = V0N
∗
∗
and such strategy is unique up to the time σ0N ∧ τ0N .
Proof. Let π = (π1 , ..., πN ), πn = (βn , γn ) be a self-financing portfolio strategy
with Z0π = z > 0 then Mnπ = (1 + r)−n Znπ (see [SKKM1]),
Mnπ
=z+
n
(1 + r)−k γk Sk −1 (ρk − r),
(2.11)
k =1
which is a martingale with respect to the filtration {Fn }0≤n≤N and the probability
P ∗ . Suppose that (σ, π) is a hedge then by the Optional Sampling Theorem (see
π
[Ne], Theorem II-2-13) for any τ ∈ J0N I have Z0π = E ∗ ((1 + r)−σ∧τ Zσ∧τ
)≥
∗
−σ∧τ
∗
E ((1+r)
R(σ, τ )). Since, by the definition, V is the infimum of such initial
capitals Z0π then V ∗ is not less than the right hand side of (2.8).
In order to prove the inequality in the other direction, for any σ ∈ J0N
set Vnσ = max E ∗ (Uτσ |Fn ) where Ukσ = (1 + r)−σ∧k R(σ, k ), k = 0, 1, . . . , N .
τ ∈JnN
Game options
449
Observe that Ukσ is Fσ∧k -measurable (and so, Fk −measurable) since both σ ∧ k
and R(σ, k ) = Xσ∧k Iσ∧k <k + Yσ∧k Iσ∧k =k are Fσ∧k -measurable. It is easy to
check directly and follows from general theorems (see [Ne], Proposition VI-12) that {Vnσ }0≤n≤N is a minimal supermartingale with respect to the filtration
{Fn }0≤n≤N such that Vnσ ≥ Unσ , n = 0, 1, . . . , N .
Now, proceeding in the standard way via the Doob supermartingale decomposition and the martingale representation (on the simple probability space above)
I obtain similarly to Sect. 2 and Sect. 5 in [SKKM1] (see also similar continuous
time arguments in the next section) that there exists a self-financing portfolio strategy π σ = (π1σ , ..., πNσ ), πnσ = (βnσ , γnσ ) with the portfolio value process
σ
Znπ = βnσ Bn + γnσ Sn such that (σ, π) is a hedge.
∗
∗
by (2.9). Then it is easy to see by the backward in, τnN
Next, define σnN
duction in n that (2.8) and (2.10) hold true. This follows also considering the
Dynkin stopping game with the payoff (1 + r)−σ∧τ R(σ, τ ) and applying results
from [Oh] though in the finite horizon case the inductive proof is much simpler.
∗
Now take σ ∗ = σ0N
∈ J0N and construct the corresponding self-financing
∗
σ∗
portfolio strategy π = π , as above, which yields the hedge (σ ∗ , π ∗ ) with the
∗
∗
∗
where the last equality
initial capital V0σ = max E ∗ ((1 + r)−σ ∧τ R(σ, τ )) = V0N
τ ∈J0N
∗
follows from (2.10). This together with the first part of the proof gives V ∗ = V0N
.
∗
∗
∗
σ∗
It remains to obtain the uniqueness. Set τ = τ0N . Since (σ , π ) is a hedge
∗
∗
∗
∗
σ∗
σ∗
∗
then M0π = V0σ = E ∗ ((1 + r)−σ ∧τ R(σ ∗ , τ ∗ )) ≤ E ∗ ((1 + r)−σ ∧τ Zσπ∗ ∧τ ∗ ) =
∗
∗
∗
σ∗
σ
σ
σ
E ∗ Mσπ∗ ∧τ ∗ = M0π
is a martingale. It follows that Zσπ∗ ∧τ ∗ =
since Mnπ
R(σ ∗ , τ ∗ ). Let now π = (π1 , ..., πN ), πn = (βn , γn ) be another self-financing
∗
portfolio strategy with Z0π = V ∗ = V0σ . Then according to the first part of the
proof Mnπ = (1 + r)−n Znπ is a martingale and in the same way as above I again
σ∗
σ∗
have Zσπ∗ ∧τ ∗ = R(σ ∗ , τ ∗ ) = Zσπ∗ ∧τ ∗ , and so Mσπ∗ ∧τ ∗ = Mσπ∗ ∧τ ∗ . Since both Mnπ
σ∗
σ∗
σ∗
and Mnπ are martingales it follows that Mnπ = Mnπ and Znπ = Znπ for all n ≤
σ ∗ ∧ τ ∗ . Since the representation (2.11) is unique, Sn > 0 and ρn =
/ r for all
∗
∗
n then γn = γnσ and βn = βnσ for all n ≤ σ ∗ ∧ τ ∗ completing the proof of
Theorem 2.1.
⊓
⊔
Remark 2.2 Consider a bit more general set up where the payoff function R(σ, τ )
given by (2.6) is replaced by R̂(σ, τ ) = Xσ Iσ<τ + Yτ Iτ <σ + Wσ Iσ=τ where Wn is
Fn -measurable, Yn ≤ Wn ≤ Xn , n = 0, 1, . . . , N and WN = YN . It follows from
∗
[Oh] that if V̂nN
, n = 0, 1, . . . , N are given by (2.8) with R̂ in place of R then
∗
they will still satisfy the recursive relations (2.7) and since V̂NN
= (1 + r)−N YN
∗
∗
I conclude that V̂nN
= VnN
for all n = 0, 1, . . . , N . By [Oh] the stopping times
∗
−k
∗
∗
σ̂nN = min{k ≥ n : (1 + r) Xk = V̂k∗ or k = N } = σnN
and τ̂nN
= min{k ≥ n :
−k
∗
∗
(1+r) Yk = V̂k } = τnN satisfy (2.10) with R replaced by R̂. Next, I define again
π
(σ, π) to be a hedge if Zσ∧n
≥ R̂(σ, n). Now the same proof as in Theorem 2.1
∗
shows that the fair price V̂ ∗ of the GCC with the payoff function R̂ equals V̂0N
,
∗
∗
and so by above, V̂ = V . This is rather interesting since R̂(σ, n) ≥ R(σ, n)
and the strict inequality is also possible when σ = n. So, sometimes (take, for
instance, N = 1, σ = 0, X0 > Y0 ) a hedging portfolio requires less initial capital
450
Y. Kifer
when the payoff function is R than when it is R̂. Still, the fair prices of the
corresponding GCC’s are the same in both cases.
Remark 2.3 Theorem 2.1 can be extended to the infinite horizon case N = ∞ with
the same proof relying on results from [Oh] provided that with P ∗ -probability
one
lim e −rn Xn = 0.
(2.12)
n→∞
In this case Ω becomes the space of sequences and all statements above will be
true now with probability one with respect to the probability P ∗ = {p ∗ , 1−p ∗ }∞ .
Since P ∗ is singular with respect to any other probability {p, 1−p}∞ with p = p ∗
and there is no reason why the stock fluctuations should be connected with this
particular probability p ∗ , it seems that this extension of Theorem 2.1 to the case
N = ∞ may have financial applications only as an approximation of a very large
N case. Still, observe that (2.12) always holds true in the game put option case
Yn = (K − Sn )+ provided the penalty δn does not grow too fast so that with
probability one lim e −nr δn = 0. Moreover, then (2.12) holds also true for the
n→∞
game call option case Yn = (Sn − K )+ since by Jensen’s inequality log(1 + r) >
p ∗ log(1 + b) + (1 − p ∗ ) log(1 + a), and so by the law of large numbers
with
n
P ∗ −probability one limn→∞ n −1 log((1 + r)−n Sn ) = limn→∞ n −1 k =1 log(1 +
ρk ) − log(1 + r) < 0.
Remark 2.4 Similarly to [SKKM1] Theorem 2.1 can be generalized to the case
when consumption or infusion of capital is also possible. In this case the price of
a portfolio π = (π1 , . . . , πN ), πn = (βn , γn ) after new stock prices at time n were
π
announced is Znπ = βn Bn + γn Sn but immediately before that Zn−1
= βn Bn−1 +
γn Sn−1 + gn (see [SKKM1]) where gn is Fn−1 -measurable. An easy modification
of the above proof gives the following formula for the fair price V ∗ of the
σ∧τ
corresponding GCC V ∗ = min max E ∗ ((1+r)−σ∧τ R(σ, τ )+
(1+r)−(k −1) gk )
σ∈J0N τ ∈J0N
k =1
and this minmax equals maxmin of the same expression. Other, correspondingly
modified, assertions of Theorem 2.1 remain true in this case, as well.
Remark 2.5 It is easy also to generalize the above set up allowing dependence
of r, a and b on time, i.e. assuming that ρk (ω) = 12 (ak + bk + ωk (bk − ak )) and
n
Bn = B0 (1 + rk ) where rk , ak , bk ; k = 1, . . . , N are nonrandom sequences
k =1
satisfying −1 < ak < rk < bk . Setting pk∗ =
rk −ak
bk −ak
and P ∗ =
N
k =1
{pk∗ , 1 − pk∗ }, an
easy modification of the proof leads to the corresponding statements of Theorem
n
2.1 with (1+r)−n replaced by (1+rk )−1 . In fact, one can consider also random
k =1
sequences rk , ak , bk satisfying certain conditions.
Game options
451
3 Continuous time
I adopt here a popular model of a financial market consisting of a savings account
with the time evolution
Bt = B0 e rt ,
r ≥0
B0 > 0,
and of a stock whose price St is the geometric Brownian motion
κ2
St = S0 exp (µ − )t + κWt
2
(3.1)
(3.2)
where {Wt }t≥0 is the standard one dimensional Wiener process starting at zero
and κ > 0, µ are some numbers. In the differential form
dBt = rBt dt and dSt = St (µdt + κdWt ),
(3.3)
where the second equation is the Ito stochastic differential equation. Again, r is
interpreted as the interest rate on the savings account and the term µdt + κdWt
is responsible for random “risky” fluctuations of the stock price where σ and
µ are called volatility and appreciation rate, respectively. Let (Ω, F , P ) be the
probability space corresponding to the Wiener process, i.e. Ω is the space of
continuous functions ω = (ωt )t≥0 , ω0 = 0, F is the Borel σ-field generated
by cylinder sets, and P is the Wiener measure on (Ω, F ). Then Wt (ω) = ωt ,
t ≥ 0. Denote by Ft W the σ-algebra generated by {Wu , u ≤ t} and by Ft the
minimal σ-algebra containing Ft W and P −null subsets of F . Then {Ft }t≥0 is
a right-continuous filtration and
St = S0 e µt Qt ,
(3.4)
2
where Qt = e κWt −(κ /2)t , t ≥ 0, is a martingale with respect to the filtration
{Ft }t≥0 .
Recall, (see, for instance, [SKKM2]) that a self-financing portfolio strategy
π with an initial capital Z0π = z > 0 and a horizon T < ∞ is a process
π = (πt )0≤t≤T of pairs πt = (βt , γt ) with progressively measurable with respect
to the filtration Ft , t ≥ 0 processes βt and γt , t ≥ 0 such that
T
rt
e |βt |dt < ∞ and
(γt St )2 dt < ∞
(3.5)
0
0
and the portfolio price
T
Ztπ
at time t ∈ [0, T ] is given by
Ztπ
=z+
t
0
βu dBu +
t
γu dSu
(3.6)
0
where Bu and Su are the same as in (3.1)–(3.3). I call π {Ft }0≤t≤T -progressively
measurable if the processes βt and γt are progressively measurable with respect
to this filtration (see [KS], Sect. 1.1).
452
Y. Kifer
Denote by JtT the set of stopping times with respect to the filtration
{Fu }0≤u≤T with values in [t, T ]. A game contingent claim (GCC) is a contract
between investors A and B consisting of a maturity date T < ∞, of selection of
a cancellation time σ ∈ J0T by A, of selection of an exercise time τ ∈ J0T by
B and of Ft -adapted right continuous with left limits (RCLL) payoff processes
∞ > Xt ≥ Yt ≥ 0, so that A pledges to pay to B at time σ ∧ τ = min(σ, τ ) the
sum
R(σ, τ ) = Xσ Iσ<τ + Yτ Iτ ≤σ .
(3.7)
As in the discrete time case, the exchange between ≤ and < in (3.7) does not
influence the final result (see Remark 3.4) provided XT = YT .
A hedge against such GCC with a maturity date T is a pair (σ, π) of a stopping
time σ ∈ J0T and a {Ft }0≤t≤T -progressively measurable self-financing portfolio
π
strategy π such that Zσ∧t
≥ R(σ, t) with probability one for each t ∈ [0, T ]. Again
∗
the fair price V of a GCC is the infimum of V ≥ 0 such that there exists a
hedge (σ, π) against this GCC with Z0π = V .
Set (see Sect. 2 in [SKKM2]),
Wtµ−r = Wt +
µ−r
t
κ
(3.8)
then the process {Wtµ−r }t≥0 is the standard Wiener process with respect to the
probability P µ−r whose restrictions Ptµ−r to Ft are equivalent to restrictions Pt
of P to Ft and
dPtµ−r
µ−r
1
Wt (ω) −
(ω) = exp −
dPt
κ
2
µ−r
κ
2
t
(3.9)
(see, for instance, [KS], Sect. 3.5).
By (3.3) and (3.8),
dSt = St (rdt + κdWtµ−r )
(3.10)
which together with (3.6) gives
dZtπ = rZtπ dt + κγt St dWtµ−r
(3.11)
for any self-financing portfolio strategy. It is important to observe that both
stochastic differential equations (3.10) and (3.11) do not depend explicitly on µ
which is usually not known.
Assume that
E µ−r sup (e −rt Xt ) < ∞,
(3.12)
0≤t≤T
where E µ−r is the expectation corresponding to the probability P µ−r .
∗
where {VtT∗ }0≤t≤T
Theorem 3.1 The fair price V ∗ of the above GCC equals V0T
µ−r
is the right continuous process such that with P
-probability one,
Game options
453
VtT∗
= essinf esssup E
σ∈JtT τ ∈JtT
µ−r
e
−rσ∧τ
R(σ, τ )Ft
= esssup essinf E µ−r e −rσ∧τ R(σ, τ )Ft .
τ ∈JtT σ∈JtT
(3.13)
(3.13)
Moreover, for each t ∈ [0, T ] and ε > 0 the stopping times
ε
∗
σtT
= inf{u ≥ t : e −ru Xu ≤ VuT
+ ε or u = T } and
(3.14)
τtTε = inf{u ≥ t : e −ru Yu ≥ VuT − ε}
∗
belong to JtT (since VTT
= e −rT YT ) and with P µ−r -probability one they satisfy
ε
ε
ε
E µ−r e −rσtT ∧τ R(σtT
, τ )Ft − ε ≤ VtT∗ ≤ E µ−r e −rσ∧τtT R(σ, τtTε )Ft + ε
(3.15)
for any σ, τ ∈ Jt,T . Furthermore, for each ε > 0 there exists a self-financing
ε
portfolio strategy π ε such that (σ0T
, π ε ) is a hedge against this GCC with the
πε
∗
initial capital Z0 ≤ V0T + ε. Suppose, in addition, that the processes Yt and
−Xt are upper semicontinuous from the left, i.e. in our circumstances they may
∗
have only positive jumps at points of discontinuity. Then the stopping times σtT
=
ε
∗
ε
ε
ε
lim σtT and τtT = lim τtT , which are well defined since σtT and τtT are monotone
ε↓0
ε↓0
in ε, with P µ−r -probability one satisfy
∗
∗
µ−r
−rσ∧τtT∗
∗
µ−r
−rσtT
∧τ
∗
(3.16)
e
R(σ, τtT )Ft
e
R(σtT , τ )Ft ≤ VtT ≤ E
E
∗
0
0
for any σ, τ ∈ Jt,T . Moreover, σtT
∧ τtT∗ = σtT
∧ τtT0 , where σtT
and τtT0 are defined
µ−r
by (3.14) with ε = 0, and so with P
-probability one,
0
∗
µ−r
0
0
−rσtT
∧τtT0
(3.17)
e
VtT = E
R(σtT , τtT )Ft .
∗
Furthermore, there exists a self-financing portfolio strategy π ∗ such that (σ0T
, π∗ )
π∗
∗
µ−r
is a hedge against this GCC with the initial capital Z0 = V0T and with P
−
0
probability one such strategy is unique in this case up to the time σtT
∧ τtT0 .
Proof. Let π be a self-financing portfolio strategy with Z0π = z > 0 then in view
of (3.11), Mtπ = e −rt Ztπ satisfies
Mtπ = M0π + κ
t
e −ru γu Su dWuµ−r ,
t ≤ T < ∞,
(3.18)
0
and so {Mtπ }0≤t≤T is a martingale with respect to the filtration {Ft }0≤t≤T .
Suppose that (σ, π) is a hedge then by the Optimal Sampling Theorem (see [KS],
Sect. 1.3) for any τ ∈ J0T ,
454
Y. Kifer
π
Z0π = E µ−r (e −rσ∧τ Zσ∧τ
) ≥ E µ−r (e −rσ∧τ R(σ, τ )).
(3.19)
If follows that
V ∗ ≥ inf
sup E µ−r (e −rσ∧τ R(σ, τ )).
σ∈J0T τ ∈J0T
(3.20)
In order to prove the inequality in the other direction for any σ ∈ J0T set
Vtσ = esssup E µ−r (Uτσ |Ft )
(3.21)
τ ∈JtT
where Utσ = e −rσ∧t R(σ, t) and I observe that Utσ is Fσ∧t −measurable. In the
same way as in Lemma from Sect. 6 of [SKKM2] I conclude that {Vtσ }0≤t≤T
is a supermartingale. Still, since Utσ is not, in general, right continuous I cannot
use this lemma directly to conclude that {Vtσ }0≤t≤T has a RCLL modification.
It is known that, in order to establish the latter assertion it suffices to show that
the function
ϕt = E µ−r Vtσ = sup E µ−r Uτσ ,
t ∈ [0, T ]
τ ∈JtT
is right continuous (see [KS], Sect. 1.3). Since Vtσ is a supermartingale it follows
that lim ϕs ≤ ϕt . For the opposite inequality, I can still employ the argument
s↓t
from the lemma cited above relying just on the right lower semicontinuity of
Utσ ,
lim Usσ = e −rσ∧t (Xσ Iσ≤t + Yt It<σ ) ≥ Utσ
s↓t
which follows since Xu ≥ Yu , u ∈ [0, T ] and both Xu and Yu are right continuous.
Thus I can and do assume that {Vtσ }0≤t≤T is a RCLL supermartingale. Moreover, in view of (3.12) the family {Vτσ }τ ∈J0T is uniformly integrable with respect
to P µ−r (see Lemma 5.5 in [Ka1]). Hence by the Doob-Meyer decomposition
theorem (see [KS], Sect. 1.4) I can write
Vtσ = M̃tσ − Aσt ,
t ∈ [0, T ],
Aσ0 = 0
(3.22)
where {M̃tσ }0≤t≤T is a RCLL martingale with respect to the filtration {Ft }0≤t≤T
and {At }0≤t≤T is a RCLL nondecreasing process such that At is Ft -measurable.
In view of the martingale representation theorem (see [KS], Sect. 3.4) there exists
a {Ft }0≤t≤T progressively measurable process {γtσ }0≤t≤T such that
t
e −ru γuσ Su dWuµ−r ,
(3.23)
e −ru γuσ Su dWuµ−r − Aσt .
(3.24)
M̃tσ = M̃0σ + κ
0
and so for all t ∈ [0, T ],
σ
Vt =
V0σ
+κ
t
0
Game options
455
Set βtσ = (M̃tσ − e −rt γtσ St )B0−1 , t ∈ [0, T ] and
σ
Ztπ = e rt M̃tσ = βtσ Bt + γtσ St ,
(3.25)
where π σ = (πtσ )0≤t≤T and πtσ = (βtσ , γtσ ) is the {Ft }0≤t≤T -progressively measurable portfolio strategy with the initial capital Z0π = V0σ , then by (3.10), (3.23),
and (3.25),
dZtπ = Ztπ dt + κγtσ St dWtµ−r = βtσ dBt + γtσ dSt ,
(3.26)
i.e. (3.6) holds true, and so π σ is self-financing. Choose a sequence of stopping
times ηn ↓ σ as n ↑ ∞ taking on only finitely many values then it is easy to
check as in Lemma VI-1-5 from [Ne] that (3.21) implies also
Vησn ∧t = esssup E µ−r (Uτσ |Fηn ∧t ),
τ ∈Jηn ∧tT
where, for a stopping time η I denote by JηT the set of stopping times with
values between η and T . This together with (3.21) and (3.22) yield that for any
t ∈ [0, T ] with P µ−r -probability one,
σ
Zηπn ∧t = e rηn ∧t M̃ησn ∧t = e rηn ∧t (Vησn ∧t + Aσηn ∧t ) ≥ e rηn ∧t Vησn ∧t
=e
rηn ∧t
esssup E
µ−r
τ ∈Jηn ∧tT
(Uτσ |Fηn ∧t )
≥e
rηn ∧t
E
µ−r
(3.27)
(Utσ |Fηn ∧t )
= e rηn ∧t Utσ = e r(ηn ∧t−σ∧t) R(σ, t) ≥ R(σ, t)
(3.27)
since Utσ is Fσ∧ t −measurable and Fσ∧ t ⊂ Fηn ∧ t . The right continuity of
σ
Zsπ = e rs M̃sσ in s enables me to pass to the limit in (3.27) as n ↑ ∞ yielding
σ
π
Zσ∧t
≥ R(σ, t), and so (σ, π σ ) is a hedge.
Next, extend the payoff processes Xt and Yt beyond T by Xt = XT and
Yt = YT for all t > T so that the right continuity, existence of left limits, and
upper simicontinuity from the left are preserved. Denote by J0∞ the set of
stopping times with values in [0, ∞] with respect to the filtration {Ft }t≥0 which
is defined for all t ≥ 0 anyway. Consider a game between two players I and II
with the payoff processes e −rt Xt and e −rt Yt so that if I chooses a stopping time
σ and II chooses a stopping time τ then I pays to II the sum
e −rσ Xσ Iσ<τ + e −rτ Yτ Iτ ≤σ = e −rσ∧τ R(σ, τ ).
(3.28)
Next, I intend to apply to this game the results from [LM] which were stated there
for bounded payoff processes but they remain true for Xt ≥ Yt ≥ 0 satisfying
(3.12). It follows from [LM] that
def
Ṽt∗ = essinf esssup E µ−r e −rσ∧τ R(σ, τ )Ft
(3.29)
σ∈Jt∞ τ ∈Jt∞
= esssup essinf E
τ ∈Jt∞ σ∈Jt∞
µ−r
e
−rσ∧τ
R(σ, τ )Ft
(3.29)
456
Y. Kifer
and for each ε > 0 the stopping times
σ̃tε = inf{u ≥ t : e −ru Xu ≤ Vu∗ + ε} and τ̃tε = inf{u ≥ t : e −ru Yu ≥ Ṽu∗ − ε}
(3.30)
satisfy
ε
ε
E µ−r e −r σ̃t ∧τ R(σ̃tε , τ )Ft − ε ≤ Ṽt∗ ≤ E µ−r e −rσ∧τ̃t R(σ, τ̃tε )Ft + ε
(3.31)
for any σ, τ ∈ Jt∞ .
From the definition of Xt and Yt beyond T it follows easily that Ṽt∗ = VtT∗
for all t ∈ [0, T ] since the player II may only decrease his gain if he stops the
game later than T . Then by (3.14) and (3.30), τtTε = τ̃tε ∈ JtT for all t ∈ [0, T ].
ε
But then σtT
= σ̃tε ∧ T also satisfies (3.31), and so (3.13) and (3.15) follow from
(3.29) and (3.31).
ε
Now take σ ε = σ0T
∈ J0T and construct the corresponding self-financing
ε
ε
portfolio strategy π = π σ , as above, which yields the hedge (σ ε , π ε ) with the
initial capital
ε
V0σ = sup E µ−r (e −rσ
τ ∈J0T
ε
∧τ
∗
R(σ, τ )) ≤ V0T
+ε
(3.32)
where the last inequality in (3.32) follows from (3.15). Since the fair price V ∗
of the GCC is the infimum of initial capitals for which hedging is possible it
∗
∗
follows that V ∗ ≤ V0T
.
+ ε. This being true for any positive ε yields V ∗ ≤ V0T
∗
∗
∗
On the other hand, by (3.13) and (3.20), V ≥ V0T , i.e. in fact, V = V0T , as
required.
Next, suppose that −Xt and Yt are left upper semicontinuous. Since σtε and
τtε may only grow when ε decreases then
def
ε
∗ def
σtT
∈ JtT and τtT∗ = lim τtTε ∈ JtT .
= lim σtT
ε↓0
ε↓0
0
∗
∧ τtT0 = σtT
∧ τtT∗ follows
Letting ε ↓ 0 in (3.15) one arrives at (3.16) and σtT
easily too (see Theorem 15 in [LM]).
∗
∗
Let now σ ∗ = σ0T
and π ∗ = π σ so that (σ ∗ , π ∗ ) is the corresponding hedge.
∗
∗
Then by (3.16) it follows similarly to (3.31) that V0σ ≤ V0T
. Since I already
∗
∗
∗
∗
σ
∗
proved that V = V0T and by the definition V0 ≥ V , it follows that V0σ = V ∗ .
Finally, I obtain the uniqueness assertion in the same way as in the discrete
time case. Namely, I derive using (3.27) that if π = (πt )0≤t≤T , πt = (βt , γt ) is
σ∗
∗
another self-financing portfolio strategy with Z0π = V ∗ = V0σ then Ztπ = Ztπ
0
for all t ≤ σtT
∧ τtT0 . Now the pair βt and γt is uniquely defined by (3.6) and
(3.11), completing the proof of Theorem 3.1.
⊓
⊔
In the general continuous time case direct effective computations of the fair
price V ∗ are hardly possible. One possibility is to discretize time and to obtain
V ∗ as a limit of values Vj∗ of a sequence of discrete time games for which Vj∗
can be obtained by the backward induction as in (2.7). Namely, let Jk(n)
,T be the
Game options
457
set of stopping times from J0T taking on only values jn −1 T for j = k , k + 1, ..., n
and denote by Vk∗(n)
,T the value of the corresponding Dynkin game when stopping
is allowed only at times jn −1 T , j = k , k + 1, ..., n, i.e.
Vk∗(n)
,T = inf
sup E µ−r e −rσ∧τ R(σ, τ )|Fkn −1 T
(n)
σ∈Jk(n)
,T τ ∈Jk ,T
(3.33)
inf E µ−r e −rσ∧τ R(σ, τ )|Fkn −1 T .
sup
σ∈Jk(n)
,T
τ ∈Jk(n)
,T
As in Theorem 2.1 these values satisfy the recursive relations
−rkn
Vk∗(n)
,T = min e
−1
Xkn −1 T , max e −rkn
T
−1
T
Ykn −1 T , E µ−r Vk∗(n)
+1,T |Fkn −1 T
,
(3.34)
∗(n)
∗(n)
Vn,T
= YT , and so, in principle, one can compute V0,T
which is the price of
the corresponding GCC when A and B can cancel and exercise only at times
jn −1 T , j = 0, 1, ..., n.
Proposition 3.2 Assume that with P µ−r −probability one Xt and Yt , t ∈ [0, T ]
are continuous processes. Then
∗(n)
∗
.
V ∗ = V0T
= lim V0,T
n→∞
(3.35)
(n)
Proof. For any τ ∈ J0T denote by dn (τ ) the stopping time from J0,T
defined
−1
−1
−1
by dn (τ )(ω) = jn T if (j − 1)n T < τ (ω) ≤ jn T . In view of the recursive
relations (3.34), as in any discrete bounded time Dynkin’s game, there exist
(n)
(n)
(n)
optimal stopping times σ0,T
, τ0,T
∈ J0,T
such that
(n)
(n)
∗(n)
E µ−r e −rσ0,T ∧dn (τ ) R(σ0,T
∧ dn (τ )) ≤ V0,T
≤ E µ−r e
(n)
−rdn (σ)∧τ0,T
(3.36)
(n)
R(dn (σ) ∧ τ0,T
)
for any σ, τ ∈ J0T . Set
ε(n)
T (ω) =
max(|Xt (ω) − Xs (ω)|, |Yt (ω) − Ys (ω)|).
sup
s,t∈[0,T ],|t−s|≤n −1 T
Then for any σ, τ ∈ J0T ,
−rdn (σ)∧dn (τ )
e −rσ∧dn (τ ) R(σ ∧ dn (τ )) + ε(n)
R(dn (σ) ∧ dn (τ ))
T ≥e
(3.37)
≥ e −rn
−1
T −rdn (σ)∧τ
e
R(dn (σ) ∧ τ ) − ε(n)
T .
(n)
(n)
Now, applying (3.16) with t = 0, σ = σ0,T
, τ = τ0,T
, applying (3.36) with σ =
(n)
∗
∗
∗
σ0T
, τ = τ0,T
, and applying the first inequality in (3.37) with σ = σ0T
, τ = τ0T
(n)
∗
and the second inequality in (3.37) with σ = σ0,T
, τ = τ0T
I derive easily that
e −rn
−1
T
∗(n)
∗
∗
µ−r (n)
V0T
− E µ−r ε(n)
εT .
T ≤ V0,T ≤ V0T + E
458
Y. Kifer
µ−r
By continuity of Xt and Yt , ε(n)
−almost surely as n → ∞ and since
T → 0 P
µ−r (n)
I assume (3.12) it follows that E
εT → 0 as n → ∞ yielding (3.35).
⊓
⊔
Next, I discuss the infinite horizon case.
Proposition 3.3 Suppose that the processes Xt ≥ Yt ≥ 0 are defined for all
t ∈ [0, ∞) and satify conditions of Theorem 3.1, in particular (3.12), with T = ∞.
Assume that P µ−r −almost surely
lim e −rt Xt = 0.
t→∞
(3.38)
Then the conclusion of Theorem 3.1 remains true for T = ∞, i.e. the fair prices
∗
Vt∞
of the GCC with infinite horizon starting at t satisfy (3.13)–(3.17) with T =
∞.
∗
Set V ∗ = V0∞
and let V ∗(n) be the value of the Dynkin game corresponding
to the payoff processes e −rt Xt and e −rt Yt but where stopping is allowed only at
times jn −1 , j = 0, 1, 2, .... Then
V ∗ = lim V ∗(n) .
n→∞
(3.39)
Proof. The first part of Proposition 3.3 follows in the same way as Theorem 3.1
(cf. [Ka1], Sect. 6) taking into account that (3.38) enables me to apply the results
from [LM] again. For the second part, set γN (ω) = supt≥N e −rt Xt (ω). Then, in
the same way as in the proof of Proposition 3.2 considered with nN in place of
n and with N in place of T I obtain easily that
−1
)
)
e −rn V ∗ − E µ−r (ε(nN
+ γN ) ≤ V ∗(n) ≤ V ∗ + E µ−r (ε(nN
+ γN ).
N
N
(3.40)
Since (3.12) is assumed to be true with T = ∞ then (3.38) implies that
E µ−r γN → 0 as N → ∞. Hence, letting in (3.40), first, n → ∞ and then
N → ∞ I arrive at (3.39).
⊓
⊔
Remark 3.4 Suppose that the payoff function R(σ, τ ) given by (3.7) is replaced
by R̂(σ, τ ) = Xσ Iσ≤τ + Yτ Iτ <σ . Assume also that XT = YT . Then by Lemma
5 from [LM], (3.13) will remain true when R is replaced by R̂ with the same
process VtT∗ . Then (3.15)–(3.17) will hold true, as well, with R̂ in place of R. As
in the discrete time case, it follows that the fair price V̂ ∗ of the GCC with the
∗
∗
payoff function R̂ equals V ∗ = V0T
with V0T
given by (3.13) if hedging pairs
π
(σ, π) are required to satisfy Zσ∧t ≥ R̂(σ, t) with P µ−r -probability one for each
t ∈ [0, T ]. The proof is the same as in Theorem 3.1 (and even a bit easier since
R̂(σ, t) is right continuous).
Remark 3.5 Similarly to [SKKM2] and [Ka1] Theorem 3.1 can be generalized to
the case when consumption is also possible. If {gt }0≤t≤T is a {Ft }0≤t≤T adapted
T
consumption process (see [Ka1]) with E µ−r ( |gt |dt) < ∞ then the fair price V ∗
0
Game options
459
of the GCC in this case will be given by V ∗ = inf
sup E µ−r (e −rσ∧τ R(σ, τ )+
σ∈J0T τ ∈J0T
σ∧τ
e −ru gu du).
0
Remark 3.6 It is not difficult to generalize the set up considering r = rt , µ = µt ,
and κ = κt in (3.3) being {Ft }0≤t≤T −adapted stochastic processes satisfying
certain conditions. In the usual American contingent claim case this was done
in [Ka1]. On the other hand, one can deal with rt , µt , and σt being stochastic
processes independent of the driving Wiener process Wt , i.e. to consider a (B , S )market in a random dynamical environment.
4 Markov case
Taking into account that the stock fluctuation processes {Sn }n≥0 and {St }t≥0
given by (2.2) and (3.2), respectively, are Markov processes one can employ
other methods of computations of the fair price V ∗ of a GCC if Xn and Yn or Xt
and Yt depend only on Sn or St , correspondingly.
I start with the discrete time case. Let Xn = β n X (Sn ) and Yn = β n Y (Sn ),
n = 0, 1, 2, . . . , 0 < β ≤ 1 for some Borel functions X and Y on (0, ∞).
Particular cases of this situation are Yn = β n (K − Sn )+ and Yn = β n (Sn − K )+ ,
which are discounted put and call game options, with the penalty process having
the form δn = β n δ(Bn ) of just δn = β n δ for some constant δ > 0. In this
case it follows from (2.7), (2.8), and the Markov property that there exist Borel
∗
functions vk = vk (x ), k = 0, 1, . . . on (0, ∞) such that VnN
= (αβ)−n vN −n (Sn ),
−1
where α = (1 + r) , and for n = 0, 1, . . .
vn+1 (x ) = U vn (x ),
v0 (x ) = Y (x )
(4.1)
where U g(x ) = min(X (x ), max(Y (x ), αβEx∗ g(S1 ))) and Ex∗ is the expectation for
P ∗ provided S0 = x . This provides recursive formulas for computation of the fair
price V ∗ (x ) = vN (x ) of the GCC with the horizon N < ∞ given S0 .
By (4.1) the sequence vn , n = 0, 1 . . . is monotone nondecreasing, Y ≤ vn ≤
X and the limit
v(x ) = lim vn (x ) = lim U n Y (x )
(4.2)
n→∞
n→∞
satisfies the equation U v = v. Moreover, it follows from (2.8) and the equality
∗
, S0 = x that v equals the value of the infinite game between the
vN (x ) = V0N
players I and II described in Sect. 2 when only finite stopping times are allowed
(cf. [El] and [Oh]), and so the fair price V ∗ = V ∗ (x ) of the GCC with the infinite
horizon N = ∞ given S0 = x equals v(x ). By (2.9) the corresponding optimal (or
rational) stopping times (saddle point) for the GCC with N < ∞ are given by
∗
σnN
= min{0 ≤ n ≤ N : X (Sn ) = vN −n (Sn ) or n = N },
∗
τnN
= min{0 ≤ n ≤ N : Y (Sn ) = vN −n (Sn )}
(4.3)
460
Y. Kifer
and for the N = ∞ case σ ∗ = min{n ≥ 0 : X (Sn ) = v(Sn )}, τ ∗ = min{n ≥ 0 :
Y (Sn ) = v(Sn )} provided that with P ∗ -probability one σ ∗ and τ ∗ are finite.
∗
Let Yn = β n Y (Sn ), Xn = β n (Y (Sn ) + δ) and consider the fair price V0N
= vN
as a function of δ ≥ 0 for each fixed initial stock price S0 = x > 0. It is clear that
vN = vN (x , δ) is continuous, nondecreasing, and piecewise linear in δ for every
x > 0. Moreover, there exist 0 ≤ δ0 (x ) ≤ δ1 (x ) < ∞ such that vN (x , δ) = Y (x )
for all 0 ≤ δ ≤ δ0 (x ) and when δ ≥ δ1 (x ) then vN (x , δ) equals the fair price
of the standard American option (where only B is allowed to exercise) with
the horizon N . For δ0 (x ) < δ < δ1 (x ) the graph of vN (x , δ) in δ may have
different degrees of complexity depending on parameters of the problem. On
Fig. 1 I exhibit the graph of vN (x , δ) in δ for the game put option case where
Y (z ) = (10 − z )+ , N = 20, x = 9.65, β = 1, a = −0.1, b = 0.1, r = 0.05.
0.42
0.41
0.4
0.39
0.38
0.37
0.36
00.35
0.05
0.1
0.15
0.2
0.25
del
0.3
Fig. 1.
Consider, next, the case β = 1 and Y (x ) = (x − K )+ . Since
E ∗ (Sn+1 |Fn ) = Sn (1 + p ∗ b + (1 − p ∗ )a) = (1 + r)Sn = α−1 Sn
(4.4)
then αn Sn is a martingale, and so αn Y (Sn ) = (αn Sn − αn K )+ is a submartingale
(with respect to the probability P ∗ ). Thus, in view of the Optional Sampling
Theorem, for the game call option case with a horizon N < ∞ the fair price is
given by
V ∗ = min E ∗ ((ασ (Sσ − K )+ + δσ )Iσ<N + αN (SN − K )+ Iσ=N ).
σ∈J0N
(4.5)
This corresponds to the well known fact that American call options with an
expiration date N < ∞ coincide with the corresponding European call options.
In the game call option case it follows that the buyer B should exercise as late
as possible, i.e. at the expiration date N if N < ∞. On the other hand, the
Game options
461
seller A should choose an optimal cancellation stopping time which minimizes
in (4.5) and it is easy to give examples when this stopping time is nontrivial (i.e.
nonconstant).
For each m = 0, 1 . . . , set CmA = {x : vm (x ) < X (x )}, CmB = {x : vm (x ) >
Y (x )}, Cm = CmA ∩ CmB and DmA = {x : vm (x ) = X (x )}, DmB = {x : vm (x ) = Y (x )},
Dm = DmA ∩ DmB so that Cm ∪ Dm = R since Y ≤ vm ≤ X .
Since the sequence vn is nondecreasing then assuming that X (x ) > Y (x ) for
A
B
all x I have CnA ⊂ Cn−1
⊂ . . . ⊂ C0A = R, ∅ = C0B ⊂ . . . ⊂ Cn−1
⊂ CnB , and
A
A
A
B
B
B
Dn ⊃ Dn−1 ⊃ . . . ⊃ D0 = ∅, R = D0 ⊃ . . . ⊃ Dn−1 ⊃ Dn , n = 0, 1, . . . . By
(4.3),
∗
σnN
= min{0 ≤ n ≤ N : Sn ∈ DNA −n or n = N } and
∗
τnN
= min{0 ≤ n ≤ N : Sn ∈
(4.6)
DNB −n },
so that A or B should stop when the stock price Sn gets to the domain DNA −n or
DNB −n , respectively.
In the continuous time case the stock price fluctuations St form the Markov
diffusion process solving the stochastic differential equation (3.10). Suppose that
Xt = e −βt X (St ), Yt = e −βt Y (St ), β > 0 and T = ∞. In particular one can
take Yt = e −βt (K − St )+ or Yt = e −βt (St − K )+ which are discounted put or
call game options, respectively, with some penalty process δt = e −βt δ(St ) or
even δt = e −βt δ for some constant δ > 0. Though (3.10) is a degenerate at
zero equation but considering instead the stochastic differential equation for Lt =
log St one can deal with nondegenerate diffusions. Then the fair price V ∗ =
V ∗ (x ) of the GCC with infinite horizon given S0 = x being the value of the
infinite time optimal stopping game with the payoff processes e −(r+β)t X (St ) and
e −(r+β)t Y (St ) can be described via certain variational inequalities (see [BF1]).
Numerical schemes for solving this type of variational inequalities were studied
in [Di]. Some computational algorithms for variational inequalities emerging in
American options were justified in [JLL] and it seems quite plausible that they
can be extended to the game options case. A free boundary approach to even
a more general problem for nonzero sum games can be found in [BF2], and
so computational methods for elliptic (or parabolic, if the finite horizon case is
considered) free boundary problems can be employed here, as well.
Observe that discretizing time one obtains recursive relations of the form
(2.7). Then the value of the continuous time game, and so of the corresponding
GCC, can be obtained by letting the discretization step to zero. Namely, let
Ut g(x ) = min(X (x ), max(Y (x ), e −(r+β)t Exµ−r g(St ))).
(4.7)
Suppose that limt→∞ e −(r+β)t X (St ) = 0 P µ−r − almost surely, which holds true
if, for instance, β > 0 and X (x )(|x | + 1)−1 is bounded (since e −rt St is a positive
martingale). Then
V ∗ (x ) = lim lim Unk−1 Y (x ).
(4.8)
n→∞ k →∞
∗(n)
limk →∞ Unk−1 Y
−1
Indeed, V
=
is allowed only at times jn
is the value of the Dynkin game when stopping
, j = 0, 1, 2, ... (see [El]), and so (4.8) follows from
462
Y. Kifer
Proposition 3.3. In the finite horizon case such discrete time approximations can
be carried out also when β = 0 since then one relies on Proposition 3.2 in place
of Proposition 3.3. Another possibility is to approximate continuous time GCC’s
based on the geometric Brownian motion stock evolution by discrete time GCC’s
based on the binomial CRR model stock evolution. For the American options
case this together with error estimates was studied in [La] and after an appropriate
modification the method there seems to be extendable to the game options case.
Observe that Y ≤ V ∗ ≤ X and consider 4 domains D A = {x : X (x ) = V ∗ (x )},
B
D = {x : Y (x ) = V ∗ (x )}, C A = {x : X (x ) > V ∗ (x )}, and C A = {x : Y (x ) <
V ∗ (x )}. On C = C A ∩ C B the function V ∗ (x ) satisfies the equation
1 2 2 d 2 V ∗ (x )
dV ∗ (x )
+ rx
κ x
= (r + β)V ∗ (x )
2
2
dx
dx
(4.9)
with the free boundary conditions V ∗ |∂D A = X and V ∗ |∂D B = Y . A more specific
analysis of this problem for the case Y (x ) = (x − K )+ or Y (x ) = (K − x )+ and
X (x ) = Y (x ) + δ can be carried out, in principle, along the lines of Sect. 8 from
[SKKM2] though it is more complicated here. Still, when β = 0 one conclusion
follows easily in the game call option case. Namely, (3.10) implies that e −rt St
is a martingale, and so in the game call option case e −rt Yt = (e −rt St − e −rt K )+
is a submartingale. Thus, by the Optional Sampling Theorem the fair price V ∗
for the game call option with a finite horizon T < ∞ is given by
V ∗ = inf E µ−r ((e −rσ (Sσ − K )+ + δσ )Iσ<T + e −rT (ST − K )+ Iσ=T )
σ∈J0T
and the buyer should not exercise before the expiration date though the seller
has to find an optimal cancelation time.
References
[BF1]
[BF2]
[CK]
[CRR]
[Di]
[Dy]
[El]
[JLL]
[Ka1]
[Ka2]
Bensoussan, A., Friedman, A.: Non-linear variational inequalities and differential games
with stopping times. J. Funct. Anal. 16, 305–352 (1974)
Bensoussan, A., Friedman, A.: Nonzero-sum stochastic differential games with stopping
times and free boundary problems. Trans. Amer. Math. Soc. 231, 275–327 (1977)
Cvitanić, J., Karatzas, I.: Backward stochastic differential equations with reflection and
Dynkin games. Ann. Probab. 24, 2024–2056 (1996)
Cox, J.C., Ross, R.A., Rubinstein, M.: Option pricing: a simplified approach. J. Financ.
Econom., 7, 229–263 (1976)
Di Marco, S.C.: On the numerical resolution of Isaac’s inequalities. RAIRO Rech. Opér.
31(4), 363–373 (1997)
Dynkin, E.B.: Game variant of a problem on optimal stopping. Soviet Math. Dokl. 10,
270–274 (1969)
Elbakidze, N.V.: Construction of the cost and optimal policies in a game problem of
stopping a Markov process. Th. Probab. Appl. 16, 163–168 (1976)
Jaillet, P., Lamberton, D., Lapeyre, B.: Variational inequalities and the pricing of American options. Acta Appl. Math. 21, 263–289 (1990)
Karatzas, I.: On the pricing of American options. Appl. Math. Optim. 17, 37–60 (1988)
Karatzas, I.: Optimization problems in the theory of continuous trading. SIAM J. Control
Optim. 27, 1221–1259 (1989)
Game options
463
Kifer, Y.: Optimal stopping in games with continuous time. Th. Probab. Appl. 16,
545–550 (1971)
[Kr]
Kramkov, D.O.: Optional decomposition of supermartingales and hedging contingent
claims in incomplete security markets. Probab. Theory and Rel. Fields 105, 459–479
(1996)
[KS]
Karatzas, I., Shreve, S.E.: Brownian Motion and Stochastic Calculus. Berlin Heidelberg
New York: Springer 1988
[La]
Lamberton, D.: Error estimates for the binomial approximation of American put options.
Ann. Appl. Probab. 8, 206–233 (1998)
[LM]
Lepeltier, J.P., Maingueneau, M.A.: Le jeu de Dynkin en theorie generale sans
l’hypothese de Mokobodski. Stochastics 13, 24–44 (1984)
[My]
Myneni, R.: The pricing of the American option. Ann. Appl. Probab. 2, 1–23 (1992)
[MS]
McConnell, J.J., Schwartz, E.S.: LYON taming. J. of Finance 41, 561–577 (1986)
[Ne]
Neveu, J.: Discrete-Parameter Martingales. Oxford, Amsterdam: North-Holland 1975
[Oh]
Ohtsubo, Y.: Optimal stopping in sequential games with or without a constraint of
always terminating. Math. Oper. Res. 11, 591–607 (1986)
[SKKM1,2] Shiryaev, A.N., Kabanov, Y.M., Kramkov, D.O., Melnikov, A.B.: To the theory of
computations of European and American options I (Discrete time), II (Continuous time).
Th. Probab. Appl. 39, 14–60, 61–102 (1994)
[Ki]