Startseite Valuation of American Continuous-Installment Options Under the Constant Elasticity of Variance Model
Artikel Öffentlich zugänglich

Valuation of American Continuous-Installment Options Under the Constant Elasticity of Variance Model

  • Guohe Deng EMAIL logo und Guangming Xue
Veröffentlicht/Copyright: 25. April 2016
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

This article prices American-style continuous-installment options in the constant elasticity of variance (CEV) diffusion model where the volatility is a function of the stock price. We derive the semi-closed form formulas for the American continuous-installment options using Kim’s integral representation method and then obtain the closed-form solutions by approximating the optimal exercise and stopping boundaries as step functions. We demonstrate the speed-accuracy of our approach for different parameters of the CEV model. Furthermore, the effects on both option price and the optimal boundaries are discussed and the causes of underestimating or overestimating the option prices are analyzed under the classical Black-Scholes-Merton model, in particular, for the case of elasticity coefficient with numerical examples.

1 Introduction

Installment options are exotic options in which a smaller amount of up-front premium is paid at the time of purchase, and then a sequence of installments are paid up to a fixed maturity. To keep the option alive, the holder must continue to pay the premiums at any payment date, although the holder has the right to choose at any time to terminate installment payments by stopping the option contract. This reduces considerably the cost of entering into a hedging strategy and liquidity risk associated with other OTC derivatives. Therefore, the installment options have been traded actively on exchanges and the OTC markets. An installment option with payments at pre-specified dates is referred to as a discrete-installment (DI) option, whereas its continuous-time limit in which a constant stream of installments are paid at a certain rate per unit time is referred to as continuous-installment (CI) option. In this paper, we consider the American CI options in which the holder has the right, but not the obligation, to stop making installment payments by either exercising the option or stopping the option contract at or prior to expiry.

There are relatively few and quite recent studies about the installment options. Davis, et al.[1, 2] obtained no-arbitrage bounds and static hedging strategies for the European DI options, then by Ben-Ameur, et al.[3] for value the American DI option. Girebsch, et al.[4] derived a closed-form solution to the value of the DI option and examined the limiting case with continuous payment plan. For CI options, Alobaidi, et al.[5] used a partial Laplace transform to the inhomogeneous partial different equation (PDE) for the option value function and analyzed asymptotic properties of the optimal stopping boundary close to maturity. Ciurlia and Roko[6] derived an integral representation of the initial premium for the American-style and applied the multi-piece exponential function (MEF) method to the valuation formulas. Alobaidi and Mallier[7] considered the behavior of the CI options close to expiry by applying an asymptotic expansion. More recently, in [8, 9], Kimura applied the Laplace-Carlson transforms (LCT) to the inhomogeneous PDE for the initial premium and obtained explicit Laplace transforms of both the initial premium and Greeks for the European and American-style CI options. Ciurlia and Caperdoni[10] extended the analysis to the perpetual CI case. Also, Ciurlia [1113] and Mezentsev, et al.[14] provided alternative numerical methods to price the European and American-style CI options. Huang, et al.[15] and Deng[16] have considered the pricing of the American CI option on bond under the interest rate model and a class of American CI option of barrier type, respectively. Finally, Yang, et al.[17], and Yang and Yi[18] considered respectively a parabolic variational inequality problem arising from the European and American-style CI options and proved the existence and uniqueness of the solution to this problem and some properties of the free boundaries.

However, all these works above on the CI options generally assume to be the risky asset price dynamics driven by the celebrated BSM model in [19] and [20]. This means that the volatility of the risky asset price is only a constant, which is too restrictive and contradicted by the empirical evidence in the literature: The existence of a negative correlation between stock returns and realized stock volatility and the inverse relation between the implied volatility and the strike price of an option contract. In the financial literature, this problem has been extended to time-varying volatility models. One popular alternative is the constant elasticity of variance (CEV) model developed by Cox[21]. The CEV model features an evolution process of asset prices that can produce an inverse relationship between asset price and asset price volatility. The log-normally process used by Black and Scholes[19] and Merton[20] is a member of this class. The CEV model has an ability to allow for time-varying volatility of stock returns without having to abandon the assumption that the current stock price, at any point in time, is a sufficient statistic for describing the evolution of stock returns over the next instant, and can capture the implied volatility skew. The literature on the pricing of European or American-style options under the CEV model is quite rich, see, for example, [2331]. However, to my knowledge, there are relatively few works about the valuation of American-style CI options under the time-varying volatility model (see Deng[32]).

The objective of this paper is to value the American-style CI options under the CEV model by using the Kim integral equation approach in [33, 34], which generates efficient and accurate numerical procedures for the boundaries and provides a simple approximation for the initial premium and Greeks of this options. The first contribution of this paper is to derive integral representations for initial premium of the American CI call or put options when the underlying asset follows the CEV model. We then apply these results to obtain an nonlinear integral equation system for the optimal stopping and exercise boundaries. The second contribution is to implement numerical method for valuing the American CI options and provide representative numerical results under the CEV model for a large set of parameter values.

This paper is organized as follows. In Section 2, we review the CEV model and formulate the valuation problem for the American CI option both as parabolic variational inequality problem and a free-boundary problem for call (or put) case, obtaining integral representations for both the option value and the optimal boundaries, we also give the boundary conditions and hedging. In Section 3, applying trapezoidal rule to approximate the quadrature formulas in the Kim integral equations for the optimal boundaries. Numerical examples and discussions are stated in Section 4. Conclusions are in Section 5.

2 The CEV Model and American CI Options

Under the risk-neutral probability measure Q, the CEV model of Cox in [21] assumes that the asset price S = (St)t ≥ 0 evolves the following stochastic differential equation:

dSt=(rδ)Stdt+σ(St)StdWt(1)

with a local volatility function given by

σ(St)=σStθ21(2)

for σ, θ ∈ ℝ, and where r ≥ 0 denotes the instantaneous riskless interest rate, which is assumed to be constant, δ ≥ 0 represents the continuous dividend yield for the asset price, Wt is a one-dimensional standard Brownian motion defined on a filtered probability space (Ω, 𝓕, (𝓕t)0 ≤ tT, Q) where 𝓕t is supposed to be generated by {Wu; 0 ≤ ut} and to be satisfied the usual conditions.

The elasticity of return variance with respect to the asset price is θ − 2 through dυ(St)/dSυ(St)/S=θ2, where υ(St)=σ2(St)=σ2Stθ2 is the return variance. If θ = 2, then the elasticity is zero and the CEV model (1) becomes the BSM model developed by Black and Scholes[19], and Merton [20]. If θ = 1, then the CEV model (1) is the square-root diffusion model of Cox and Ross[22]. We have the Ornstein-Uhlenbeck process by setting θ to zero. For θ < 2 (or θ > 2 ) the local volatility given by (2) is a decreasing (or increasing) function of the asset price. This produces a probability distribution similar to that observed for equities with a heavy left (right) tail and less heavy right (left) tail. The model parameter σ is the scale parameter fixing the initial instantaneous volatility at time t=0,σ0=σ(S0)=σS0θ21.

Under the CEV model (1), the closed-form pricing solutions for European vanilla call and put options have been derived by Cox[21] for θ < 2 and by Emanuel and MacBeth[25] for θ > 2 as follows.

Lemma 1

Given the European vanilla call option with payoff φ(ST) = max{STK,0} = (STK)+at maturity date T and strike price K. Under the model(1)and(2), then its price Ce(t,T,S,K) at time t ∈[0,T] is

Ce(t,T,S,K)=Steδτ(T)ϕ1(St,K,T)Kerτ(T)ϕ2(St,K,T),(3)
where ϕi(⋅), i = 1,2 are defined respectively as
ϕ1(D(t),D(v),v)=Q(2yD(v)(v);2+22θ,2xD(t)(v)),ifθ<2,N(lnD(t)D(v)+(rδ+12σ2)τ(v)στ(v)),ifθ=2,Q(2xD(t)(v);2θ2,2yD(v)(v)),ifθ>2,
and
ϕ2(D(t),D(v),v)=1Q(2xD(t)(v);22θ,2yD(v)(v)),ifθ<2,N(lnD(t)D(v)+(rδ12σ2)τ(v)στ(v)),ifθ=2,1Q(2yD(v)(v);2+2θ2,2xD(t)(v)),ifθ>2,
with τ(v) = vt and
k(v)=2(rδ)σ2(2θ)[e(rδ)(2θ)τ(v)1],xD(v)=k(v)D2θe(rδ)(2θ)τ(v),yD(v)=k(v)D2θ.

Here Q(x;v,y) is the complementary noncentral Chi-square distribution function taking value on x with v degrees of freedom and the non-centrality parameter y, i.e. Q(x;v,y)=P(χv2(y)x)=1χ2(x;v,y),and N(⋅) is the standard normal distribution function.

According to the Call-Put parity relationship for European options, the pricing formula at time t ∈[0,T] for the European vanilla put option with payoff φ(ST) = (KST)+ is given by

Pe(t,T,S,K)=Kerτ(T)[1ϕ2(St,K,T)]Steδτ(T)[1ϕ1(St,K,T)].(4)

Now we consider a CI option with payoff φ(ST) at maturity date T and strike price K. In addition, let q > 0 be the continuous-installment rate, which means the option holder continuously pays an amount qdt in a time dt. Denote by Vt = V(t,S;q) the initial premium at time t of this option, and 𝓓 = {(t,S):0 ≤ tT, 0 ≤ S < ∞}. Applying Itô’s lemma to Vt, we get

dVt=[Vtt+12σ2Sθ2VtS2+(rδ)SVtSq]dt+σSθ/2VtSdWt.(5)

Constructing the Δ-hedging portfolio consisting of one CI option and an amount −Δ of the asset. Then the value of this portfolio Π at time t is

Πt=VtΔSt.

It is clear that the dynamics of Π in the time interval [t,t + dt] is

dΠt=dVtΔdStΔδStdt.

Substituting (2) and (5) into the above equation , one has

dΠt=[Vtt+12σ2Sθ2VtS2+(rδ)SVtSqΔrSt]dt+σSθ/2(VtSΔ)dWt.

Setting Δ=VtS the coefficient of dWt vanishes. The portfolio is instantaneously risk-free, that is, dΠt = r Πt dt. Thus,

Vtt+12σ2Sθ2VtS2+(rδ)SVtSrVt=q,(6)

which is an inhomogeneous Black-Scholes PDE.

For an American CI options, once the whole option premiums are paid over the time interval [0,T], the holder has the right to exercise the option at any time up to the maturity date T. When the holder exercises the option early, the holder can stop making installment payments. Therefore, the initial premium at time t ∈ [0,T] of the American CI option is a solution of an optimal stopping problem[8]:

Va(t,S;q)=supτe,τsF[t,T]Et{er(Tt)φ(ST)1(τeτsT)+er(τeτst)×φ(Sτeτs)1(τeτs<T)qtτeτsTer(st)ds},(7)

where Et[⋅] = E[⋅|𝓕t] under the probability measure Q, 𝓕[t,T] denotes the set of all 𝓕t-stopping times with values in [t,T], and τe and τs are stopping times of the filtration 𝓕t. The last ingredient of Equation (7) is the net present value of the future payment stream at time t.

It is easy to see from (7) that Va(t,S;q) ≥ 0 holds as it is always possible to stop payments immediately. Moreover, Va(t,S;q) ∈ C(𝓓), Va(t,S;q) is nondecreasing in S for calls (nonincreasing for puts), and Va(t,S;q) is nonincreasing in t [18]. Clearly, the optimal stopping problem (7) corresponds to a parabolic variational inequality problem as follows:

Vat+12σ2Sθ2VaS2+(rδ)SVaSrVa=qifVa>φ(S),Vat+12σ2Sθ2VaS2+(rδ)SVaSrVa<qifVa=φ(S),(Vat+12σ2Sθ2VaS2+(rδ)SVaSrVaq)(Vaφ(S))=0,Va(T,S;q)=φ(ST),S[0,+).(8)

Solving the optimal stopping problem (7) or the parabolic variational inequality problem (8) is equivalent to finding the point (t,St) for which early exit and exercise before maturity are optimal. So there are two free boundaries, one is between {(t,S)|Va(t,S;q) = 0} and {(t,S)|Va(t,S;q) > 0}, which lies in {S|φ(S) = 0}, the other is between {(t,S)|Va(t,S;q) = φ(S)} and {(t,S)|Va(t,S;q) > φ(S)}, which lies in {S|φ(S) ≥ 0}. Denote 𝓢 = {(t,S)| Va(t,S;q) = 0}, 𝓔 = {(t,S)|Va(t,S;q) = ψ(St) > 0} and 𝓒 = 𝓓 ∖ (𝓢∪𝓔), then the sets 𝓢, 𝓔 and 𝓒 are called a exit region, exercise region and continuous region, respectively.

2.1 Call Case

Let Ca(t,S;q) denote the initial premium at time t of the American CI call option. For this call option buyer at each time t, there exists an upper critical asset price Bt above which it is optimal to stop the installment payments by exercising the option early, as well as a lower critical asset price At below which it is advantageous to terminate payments by stopping the option contract. According to these upper and lower critical asset prices the initial premium Ca(t,S;q) satisfies

Ca(t,St;q)=0,if0StAt,>(StK)+,ifAt<St<Bt,=StK,ifBtSt.(9)

The stopping and exercise boundaries are the time paths of lower and upper critical asset price At and Bt, for t ∈[0,T], respectively. Moreover, AtK and BtK for t ∈[0,T]. Therefore, these three regions above become 𝓢1 = {(t,S)| (t,S) ∈[0,T] × [0,At]}, 𝓔1 = {(t,S)|(t,S) ∈[0,T] × [Bt, + ∞)} and 𝓒1 = {(t,S)|(t,S) ∈[0,T] × (At,Bt)}, respectively. And we assume that A0 < S0 < B0.

It has been known that the parabolic variational inequality problem (8) can be deduced to a free boundary problem[6], that is, Ca(t,S;q) satisfies

{Cat+12σ2Sθ2CaS2+(rδ)SCaSrCa=qinC1,Ca(t,St;q)=0inS1,Ca(t,St;q)=(StK)+inE1,Ca(T,ST;q)=(STK)+,SR+,limStAtCa(t,St;q)=0,limStBtCa(t,St;q)=BtK,limStAtCa(t,St;q)S=0,limStBtCa(t,St;q)S=1.(10)(11)(12)(13)(14)(15)

Theorem 1

Let the asset price S satisfies(1)(2), then the value function Ca(t,S;q) at time t of the American CI call option has the following integral representation

Ca(t,S;q)=Ce(t,T,St,K)+tT{δSteδτ(u)ϕ1(St,Bu,u)(rKq)×erτ(u)ϕ2(St,Bu,u)}duqtTerτ(u)ϕ2(St,Au,u)du.(16)

Moreover, the optimal stopping and exercise boundaries, At and Bt, are solved by the following nonlinear integral equation systems:

BtK=Ce(t,T,Bt,K)+tT{δBteδτ(u)ϕ1(Bt,Bu,u)(rKq)erτ(u)×ϕ2(Bt,Bu,u)}duqtTerτ(u)ϕ2(Bt,Au,u)du,0=Ce(t,T,At,K)+tT{δAteδτ(u)ϕ1(At,Bu,u)(rKq)erτ(u)(17)
×ϕ2(At,Bu,u)}duqtTerτ(u)ϕ2(At,Au,u)du.(18)

Proof

For u ∈ [t,T], let Z(u,S) = er(ut)Ca(u,S;q) be the discounted option value function defined on 𝓓. Then, the function Z(u,S) is convex in S for all u and belongs to the class C1,2(𝓓o). Applying the Itô’s Lemma to Z(u,S), we have

Z(T,S)=Z(t,S)+tTZSdS+tT(Zu+12σ2Sθ2ZS2)du,(19)

which yields

Ca(t,St;q)=er(Tt)Ca(T,ST;q)tTer(ut)σSθ/2CaSdWutTer(ut)[Cau+12σ2Sθ2CaS2+(rδ)SCaSrCa]du.(20)

Since Et(tT|er(ut)σSθ/2CaS|2du)<, then tTer(ut)σSθ/2CuaSdWu is a 𝓕t-martingale and Brownian motion. Substituting Ca(T,ST;q) = (STK)+ and Ca = 1(Au < Su < Bu)Ca + 1(0 ≤ SuAu) × 0 + 1(SuBu) × (SuK) into (20), and taking the conditional expectation with respect to 𝓕t on both sides of (20) under the risk-neutral measure Q, we obtain

Ca(t,St;q)=Et[er(Tt)(STK)+]tTer(ut)Et[Cau+12σ2Sθ2CaS2+(rδ)SCaSrCa]du=Ce(t,T,S,K)qtTer(ut)Et[1(Au<Su<Bu)]du+tTer(ut)Et[(δSurK)1(SuBu)]du.(21)

By Appendix A, one has

Et[1(Au<Su<Bu)]=Q(Su>Au|St)Q(SuBu|St)=ϕ2(St,Au,u)ϕ2(St,Bu,u),(22)

and

Et[(δSurK)1(SuBu)]=δSte(rδ)τ(u)ϕ1(St,Bu,u)rKϕ2(St,Bu,u).(23)

Substituting the above equations (22) and (23) into (21), this leads to (16). Applying the boundary condition (14), and inserting instead of St the value At and Bt of the optimal stopping and exercise boundary at time t in (16), respectively. we obtain the nonlinear integral equation system (17) and (18) for the optimal stopping and exercise boundaries At and Bt.

The integral representation (16) express the initial premium of the American CI call option as the sum of the corresponding European call value, the early exercise premium, and the expected present value of installment payments along the optimal stopping boundary.

Propositoin 1

The optimal stopping and exercise boundaries, At and Bt, have the following properties:

  1. At ∈ ℂ[0,T] and is nondecreasing and with Bt ∈ ℂ[0,T] and is non-increasing;

  2. AT = K andBT=max{K,rKqδ}.

Proof

1) Here only gives the proof of non-decreasing for the function At, for the non-increasing of the function Bt can beproved in the same way. Note that, Since t| → Ca(t,St;q) is non-increasing, then for any ts, s, t ∈[0,T], we have Ca(t,St;q) ≥ Ca(s,Ss;q). Suppose thatAt is strictly decreasing function for t ∈[0,T], that is, there exists some ts, s, t ∈[0,T] such that At > As. But if (t,S) ∈𝓢1 and there is a StAs, we then have (s,St)𝓢1. Since (t,St) ∈𝓢1, we deduce that Ca(t,St;q) = 0 and (s,St)𝓢1, thus Ca(s,St;q) ≥ (StK)+ > 0 = Ca(t,St;q). This is in contradiction with the non-increasing function of time t ∈[0,T].

Next we prove the continuity of the function At on [0,T]. We first show the left-continuous. Recall that At is non-decreasing, so that AtAt. Therefore, we only prove that AtAt, which implies At is left-continuous. Suppose that v < t and let (v,Av) ∈𝓢1. Therefore, on the closed set 𝓢1, we have limvt(v,Av) = (t,At) ∈𝓢1. Furthermore, using the definition of the optimal stopping boundary, we get At = sup{S: S ∈𝓢1(t)}, that is, AtAt. Hence, At = At which leads to At being left-continuous.

We now prove that At is right-continuous. Since the function At is non-decreasing, the inequality AvAt holds for any v > t. Suppose that (v,Av) ∈ 𝓢1, we then have limvt(v,Av) = (t,At +) ∈𝓢1, so that Ca(t,At +;q) = 0. On the other hand, we have Ca(t,At +;q) ≥ Ca(t,At;q) = 0. Therefore, Ca(t,At +;q) > 0 in 𝓓, which is obviously false. Hence, At +At, that is At + = At, and right-continuous is proved.

In the following, we prove 2). Firstly, we prove AT = K. It is easy to see that Ca(t,St;q) = 0 for (t, St) ∈𝓢1, which implies that ATK. If we suppose that ATK, then there exists a domain 𝓓ε = {TεtT, At < S < K} ⊂ 𝓒1(ε is small enough) such that

Cat+12σ2Sθ2CaS2+(rδ)SCaSrCta=q,(t,S)Dε.

So, at t = T in the domain {AT < ST < K}, Ca(t, S; q) satisfies

Cat|t=T=[12σ2Sθ2CaS2+(rδ)SCaSrCta]t=T+q=q>0.

Noting that Ca(T,ST;q) = 0, Hence, we have Ca(t,St;q) < Ca(T,ST;q) = 0 in the domain 𝓓ε, which contradicts with the definition of Ca(t,S;q), and we obtain that AT = K.

Now, we prove BT=max{K,rKqδ}. we consider in the following two cases, respectively.

(i) If (rδ)K < q, we then deduce from (9) that BTK. If BTK, we then get BT > K, and moreover, there exists a domain D~ε = {TεtT, K < S < Bt} (ε is small enough), lying in 𝓒1 such that

Cat+12σ2Sθ2CaS2+(rδ)SCaSrCta=q,(t,S)D~ε

So, at t = T in the domain {K < ST < BT}, Ca(t,S;q) satisfies

Cat|t=T=[12σ2Sθ2CaS2+(rδ)SCaSrCa]t=T+q=rK+δS+q>rK+δS+(rδ)K=δ(SK)>0.

Noting that Ca(T,ST;q) = STK. Hence, we have Ca(t,St;q) < Ca(T,ST;q) = STK in the domain D~ε, which contradicts with Ca(t,St;q) > (SK)+, and we conclude that BT = K.

(ii) If (rδ)Kq, we prove that BT=rKqδ. If it does not hold, we assume that BT<rKqδ. Since BTK, then there exists a domain Eε={TεtT,BT<S<rKqδ} involving in 𝓔1 such that Ca(t,St;q) = StK, that is, if (t,S) ∈ 𝓔ε, then

Cat+12σ2Sθ2CaS2+(rδ)SCaSrCaq=rKδSq>0.(24)

However, Cat+12σ2Sθ2CaS2+(rδ)SCaSrCaq<0 in 𝓔1, which contradicts (24). On the other hand, if we assume that BT>rKqδ. Since rKqδ>K, then there exists a domain Cε={TεtT,rKqδ<S<BT} involving in the continuous region 𝓒1 such that

Cat+12σ2Sθ2CaS2+(rδ)SCaSrCaq=0.

In particular, at t = T we have

Cat|t=T=[12σ2Sθ2CaS2+(rδ)SCaSrCta]t=T+q=rK+δS+q>0.

That means, Ca(t,St;q) < Ca(T,ST;q) = STK, which contradicts with the result of Ca(t,St;q) ≥ (SK)+. Combining (i) and (ii), thus BT=max{K,rKqδ}.

2.2 Put Case

In much the same way as for the call case above, we can also consider the valuation of an American CI put option. For each time t ∈[0,T], there must exist a lower critical asset price Ft below which it is optimal to terminate payments by excising the option, as well as an upper critical asset price Gt above which it is advantageous to terminate payments by stopping the option. Let Pa = Pa(t,St;q) be the initial premium function of the put option. According to these lower and upper critical asset prices Ft and Gt, the initial premium function Pa(t,St;q) satisfies

Pa(t,St;q)=KSt,if0StFt,>(KSt)+,ifFt<St<Gt,=0,ifGtSt.(25)

The exercise and stopping boundaries, which are the time paths of lower and upper critical asset prices Ft and Gt, divide the domain 𝓓into three regions: A stopping region 𝓢2 = {(t,S):0 ≤ tT, GtSt < ∞}, a continuous region 𝓒2 = {(t,S):0 ≤ tT, Ft < St < Gt} and an exercise region 𝓔2 = {(t,S):0 ≤ tT, 0 ≤ StFt}. Similar to (9), in view of (25), then the initial premium function Pa(t,St;q) also satisfies the following free-boundary problem:

{Pat+12σ2Sθ2PaS2+(rδ)SPaSrPa=qinC2,Pa(t,St;q)=0inS2,Pa(t,St;q)=(KSt)+inE2,Pa(T,ST;q)=(KST)+,SR+,limStGtPa(t,St;q)=0,limStFtCa(t,St;q)=KFt,limStGtPa(t,St;q)S=0,limStFtPa(t,St;q)S=1.(26)(27)(28)(29)(30)(31)

and Pa(t,St;q) has the integral representation as follows.

Theorem 2

Let the asset price S satisfies(1)(2), then the value function Pa(t,S;q) at time t of the American CI put option with maturity date T and the strike price K, has the integral representation

Pa(t,S;q)=Pe(t,T,St,K)+K(1erτ(T))St(1eδτ(T))+tT{δSteδτ(u)ϕ1(St,Fu,u)(rK+q)erτ(u)×ϕ2(St,Gu,u)}duqtTerτ(u)ϕ2(St,Fu,u)du.(32)

Moreover, the early exercise and the early stopping boundaries Ft and Gt are determined by the following nonlinear integral equation system

KFt=Pe(t,T,Ft,K)+K(1erτ(T))Ft(1eδτ(T))+tT{δFteδτ(u)ϕ1(Ft,Fu,u)(rK+q)erτ(u)×ϕ2(Ft,Gu,u)}duqtTerτ(u)ϕ2(Ft,Fu,u)du,(33)
0=Pe(t,T,Gt,K)+K(1erτ(T))Gt(1eδτ(T))+tT{δGteδτ(u)ϕ1(Gt,Fu,u)(rK+q)erτ(u)×ϕ2(Gt,Gu,u)}duqtTerτ(u)ϕ2(Gt,Fu,u)du(34)
with subject to the terminal conditionsFT=min{K,rK+qδ}and GT = K.

Remark 2

Similar to the call case above, it is proved that the two optimal boundaries, Ft and Gt, are continuous functions in [0,T], and Ft is nondecreasing and Gt is non-increasing.

2.3 Hedging

In order to prevent the risks of price movements of the underlying stock, the option holders usually use the delta hedging to reduce or remove the risks of movements in the stock. The delta of an option, Δ, is the derivative of the option price with respect to the price of the underlying asset. The underlying asset is bought and sold according to the Δ value, which leads to create a risk-less portfolio between the underlying asset and the option. So the Δ is an important parameter in the pricing and hedging of options. As corollaries of Theorems 1 and 2, we obtain Δ value of Ca (or Pa).

Corollary 1

The Δ value of an American CI call or put option is respectively given by

Δc=ΔcE+Πc(t,S;q),(35)
Δp=ΔpE+Πp(t,S;q),(36)
where
Πc(t,S;q)=tT{δeδτ(u)ϕ1(Su,Bu,u)+δSteδτ(u)ϕ1S(St,Bu,u)(rKq)erτ(u)ϕ2S(St,Bu,u)}duqtTerτ(u)ϕ2S(St,Au,u)du
and
Πp(t,S;q)=(1eδτ(T))+tT{δeδτ(u)ϕ1(St,Fu,u)+δSteδτ(u)ϕ1S(St,Fu,u)(rK+q)erτ(u)ϕ2S(St,Gu,u)}duqtTerτ(u)ϕ2S(St,Fu,u)du,
with the first-order derivative of ϕi(S, D(v), v) with respect to the asset price S, ϕiS(S, au, u), i = 1, 2, are respectively given by
ϕ1S(St,D(v),v)=(2θ)xSt(v)St[Q(2yD(v)(v);4+22θ,2xSt(v))Q(2yD(v)(v);2+22θ,2xSt(v))],ifθ<2,1Stστ(v)n(lnStD(v)+(rδ+12σ2)τ(v)στ(v)),ifθ=2,(θ2)xSt(v)St[Q(2xSt(v);2θ22,2yD(v)(v))Q(2xSt(v);2θ2,2yD(v)(v))],ifθ>2,
ϕ2S(St,D(v),v)=(2θ)xSt(v)St[Q(2xSt(v);22θ,2yD(v)(v))Q(2xSt(v);22θ2,2yD(v)(v))],ifθ<2,1Stστ(v)n(lnStD(v)+(rδ12σ2)τ(v)στ(v)),ifθ=2,(θ2)xSt(v)St[Q(2yD(v)(v);2+2θ2,2xSt(v))Q(2yD(v)(v);4+2θ2,2xSt(v))],ifθ>2,
heren(x)=12πe12x2.

Remark 3

Under the CEV model (1)(2), the delta values of both the European vanilla call and put option, ΔcE and ΔpE, are respectively given by Larguinho et al.[27].

3 Numerical Method

Here we present our numerical method for pricing the American CI options derived in Section 2 by solving the nonlinear integral equation systems using the quadrature formulas studied by Kim[33] (named Kim’s method), as well as considered by Kallast and Kivinukk[34], and others [15, 16, 32]. This method consists of three steps: the first two steps are needed to find the numerical values of the early exercise and the stopping boundaries from the nonlinear integral equation system. When the values of the early exercise and the stopping boundaries are obtained, the third step, numerical integration of Equation (16) (or (32)), yields the value of this option based on the Richardson extrapolation method. The key idea of this method is to discrete the quadrature formulas. In the following, we take the call case as example, and for the put case is similarly discussed.

Now we divide the interval time [0, T] into n equally spaced time subintervals [ti−1, ti] with length Δt=Tn. Thus ti = i Δ t for i = 0, 1, 2, ⋯, n. Denote Ai = Ati, Bi = Bti for i = 0, 1, 2, ⋯, n. Since tn = T, by the terminal conditions in Proposition 1 2), we get

An=K,Bn=max{K,rKqδ}.(37)

Define

F1(x,y,u)=δxeδτ(u)ϕ1(x,y,u)+(qrK)erτ(u)ϕ2(x,y,u),(38)
F2(x,y,u)=qerτ(u)ϕ2(x,y,u).(39)

Then, the Kim’s equations (16)–(18) can be rewritten as

Ca(t,S;q)=Ce(t,T,S,K)+tTF1(St,Bu,u)du+tTF2(St,Au,u)du,(40)
BtK=Ce(t,T,Bt,K)+tTF1(Bt,Bu,u)du+tTF2(Bt,Au,u)du,(41)
0=Ce(t,T,At,K)+tTF1(At,Bu,u)du+tTF2(At,Au,u)du.(42)

In order to obtain the step-function approximation (Bi, Ai), i = 0, 1, 2, ⋯, n for the two optimal exercise and stopping boundaries functions (Bt, At), the trapezoidal rule is used to discrete the quadrature representations for the right-hand sides of (41) and (42). Putting t = tn−1 in (41) and (42), then

Bn1K=Ce(tn1,tn,Bn1,K)+Δt2[F1(Bn1,Bn,Δt)+F1(Bn1,Bn1,0Δt)]+Δt2[F2(Bn1,An,Δt)+F2(Bn1,An1,0Δt)],0=Ce(tn1,tn,An1,K)+Δt2[F1(An1,Bn,Δt)+F1(An1,Bn1,0Δt)]+Δt2[F2(An1,An,Δt)+F2(An1,An1,0Δt)].

Since the values (Bn, An) are known due to (37), using the backward recursive equations above the values (Bn−1, An−1) are obtained. Similarly, for the values (Bi, Ai), i = n − 2, n − 3, ⋯, 0 are determined by the following backward recursive equations

BiK=Ce(ti,tn,Bj,K)+Δt2[F1(Bj,Bn,(ni)Δt)+F1(Bi,Bi,0Δt)]+Δtk=i+1n1[F1(Bi,Bk,(ki)Δt)+F2(Bi,Ak,(ki)Δt)]+Δt2[F2(Bi,An,(ni)Δt)+F2(Bi,Ai,0Δt)],(43)
0=Ce(ti,tn,Ai,K)+Δt2[F1(Ai,Bn,(ni)Δt)+F1(Ai,Bi,0Δt)]+Δtk=i+1n1[F1(Ai,Bk,(ki)Δt)+F2(Ai,Ak,(ki)Δt)]+Δt2[F2(Ai,An,(ni)Δt)+F2(Ai,Ai,0Δt)].(44)

Noting that the values Fi(x, y, 0 Δ t), i = 1, 2 in Formulas (43) and (44) are calculated with the help of the limit process, that is,

F1(x,y,0)=limu0+F1(x,y,u)={qrK,if x=y,0,if xy,(45)
F2(x,y,0)=limu0+F2(x,y,u)={q, if x=y,0,if xy.(46)

Numerical values of (Bi, Ai), i = n − 1, n − 2, ⋯, 0, can be found from (43) and (44) using the Newton-Raphson iteration method. We denote that

g1(Bi,Ai)=Ce(ti,tn,Bj,K)+Δt2[F1(Bj,Bn,(ni)Δt)+F1(Bi,Bi,0Δt)]+Δtk=i+1n1[F1(Bi,Bk,(ki)Δt)+F2(Bi,Ak,(ki)Δt)]+Δt2[F2(Bi,An,(ni)Δt)+F2(Bi,Ai,0Δt)]+KBi,(47)
g2(Bi,Ai)=Ce(ti,tn,Ai,K)+Δt2[F1(Ai,Bn,(ni)Δt)+F1(Ai,Bi,0Δt)]+Δtk=i+1n1[F1(Ai,Bk,(ki)Δt)+F2(Ai,Ak,(ki)Δt)]+Δt2[F2(Ai,An,(ni)Δt)+F2(Ai,Ai,0Δt)].(48)

By the Newton-Raphson iteration the values (Bi, Ai), i = n − 1, n − 2, ⋯, 0, have approximations (Bi(k),Ai(k)) of order k, where for k = 0, 1, ⋯

Bi(k+1)Ai(k+1)=Bi(k)Ai(k)g1Big1Aig2Big2Ai1|(Bi(k),Ai(k))g1(Bi(k),Ai(k))g2(Bi(k),Ai(k)).(49)

For the first approximation of values (Bi, Ai) we use the values (Bi(0),Ai(0)) = (Bi+1, Ai+1), for i = n − 1, n − 2, ⋯, 0. For example, (Bn1(0),An1(0)) = (Bn, An), which is determined by (37). Using the iteration formula (49), we can get the value (Bn−1, An−1) with the order k = 3 or k = 4, which gives the accuracy of size of 10−10[34]. When the value (Bn−1, An−1) is obtained we can set (Bn2(0),An2(0)) = (Bn−1, An−1) for getting value (Bn−2, An−2) by Formula (49). Similarly, we can get all others (Bi, Ai) for i = n − 3, ⋯, 0.

Once the numerical values (Bi, Ai), i = n, n − 1, ⋯, 0 are all obtained, using the same trapezoidal rule above, then the option price (40) can be determined as follows:

C^a(0,S0;q)=Ce(0,T,S0,K)+Δt2[F1(S0,Bn,T)+F2(S0,An,T)]+Δti=1n1[F1(S0,Bi,iΔt)+F2(S0,Ai,iΔt)].(50)

The Δ value of this call option can be obtained by differentiating (50) with respect to S, that is,

Δ^c=ΔcE+Δt2[dF1(S0,Bn,T)+dF1(S0,An,T)]+Δti=1n1[dF1(S0,Bi,iΔt)+Δti=1n1F2(S0,Ai,iΔt)],(51)

where dF1(x, y, u) and dF2(x, y, u) are given by

dF1(x,y,u)=δeδτ(u)ϕ1(x,y,u)+δxeδτ(u)ϕ1S(x,y,u)+(qrK)erτ(u)ϕ2S(x,y,u),dF2(x,y,u)=qerτ(u)ϕ2S(x,y,u),

respectively. The estimates of Ĉa(0, S0;q) is accelerated through Richardson extrapolation. For instance, given the estimates C^(j)a corresponding to n = j, j = 1, 2, 3, the following three-point Richardson extrapolation can be used to compute Ca(0, S0;q):

Ca(0,S0;q)9C^(3)a8C^(2)a+C^(1)a2.(52)

4 Results and Discussions

This section provides numerical examples to illustrate the American CI option valuation under the CEV model using the Kim’s method. Firstly we examine the accuracy and efficiency of the numerical method being implemented. The method is used to an American CI call option position with maturity date T = 0.5 year, using n = 50, 150, 300 and 500 for the Kim’s integral equation systems. This has been compared against a Crank-Nicolson (CN) scheme with projected successive-over-relaxation (PSOR) technique using 4 time steps per day and 4000 and 8000 space-nodes. The parameters used are: K = 100, q = 1, σ0 = 0.2, r = 0.05 and δ = 0.04. The results are summarized in Table 1 for three spot prices 95, 100 and 105 representing out-the-money (OTM), at-the-money (ATM) and in-the-money (ITM) option, and for seven levels of θ = (−6, −4, −2, 0, 1, 2, 3)[24] showing its effect on option price. In order to determine the computational speed of the Kim’s method, we computed the CPU times for all the alternative algorithms. Since the CPU time for a single evaluation is very small, we computed the CPU time for seven levels of θ computations. All calculations have been implemented using Matlab 7.0.0(R14) running on 2.80GHz Pentium PC and reported in Table 1.

Table 1

American CI call option price found numerically under the CEV model

θ−6−4−20123CPU time(s)
     S0 = 95
CN 40002.55632.66032.77212.89232.96713.02760.8292492.75
CN 80002.55602.65412.76592.88612.95703.02660.8268593.44
Kim 502.49892.60392.71672.83762.95083.02910.7942185.47
Kim 1502.52392.62872.74112.86192.95623.02720.8191210.45
Kim 3002.55582.65432.76762.88822.95683.02680.8267255.43
Kim 5002.55692.66172.77322.89322.95703.02670.8267437.74
     S0 = 100
CN 40005.38375.34795.32685.31645.32105.32152.8203473.42
CN 80005.38355.34685.32775.31525.31565.32072.8187505.83
Kim 505.31715.28395.26425.25495.30775.32202.7848126.79
Kim 1505.34425.31075.29065.28105.31435.32092.8180224.76
Kim 3005.38335.34655.32855.31545.31535.32062.8184327.64
Kim 5005.38385.34695.32885.31555.31565.32062.8183476.47
     S0 = 105
CN 40008.94738.75478.59058.44468.38018.31695.8686542.93
CN 80008.94578.75008.58338.43768.37548.31555.8647745.78
Kim 508.87708.68718.52448.37958.36748.31515.8373190.51
Kim 1508.90418.71378.55088.40578.37418.31455.8582241.28
Kim 3008.94698.75148.58408.43618.37518.31445.8648297.67
Kim 5008.94748.75378.59068.44238.37508.31435.8651441.28

From Table 1, it can be seen that the CN scheme has higher convergence in terms of computational accuracy with a maximum error of less than 1% at each spot values. Thus we take the CN results as being the “true” solution to an accuracy of 8000 space-nodes. For all the values of n used, the call option price found using Kim’s method match those found using the CN scheme with 8000 space-nodes. We conclude from these results that the numerical method in Section 3 has converged to three decimal places as n increasing. The Kim’s method takes less than 500 seconds to compute the seven option prices. The CN approach takes more than 500 seconds for 8000 space-nodes. This numerical experiment shows that the Kim’s method is accurate and efficient. We therefore select n = 300 for the purpose of generating all further results.

Table 2 illustrates the effects of the elasticity factor θ on both prices and delta values of American CI call option under the CEV model. The risk-free interest rate is 5% per annum (r = 0.05), the underlying asset pays dividend yield 4% (δ = 0.04), the volatility parameter σ0 = 0.3, and the option strike price K = 100. We choose the current asset price S has values 95, 100 and 105, the option maturity date T = 0.5 or 1 year, and the installment rate q has values 1, 3 and 8. We employ seven different values of θ to show its effects on option prices (in Panel A) and delta hedging (in Panel B): θ = −6, −4, −2, 0, 1, 2, 3. We point out the following features of the results. First, the value of θ has positive or negative impact on option prices. For θ < 0, the OTM options and the ITM options react to the values of θ just oppositely. Whereas OTM options increase in value with increasing θ, both ITM option and ATM option prices go down. Second, for a given options, the impacts of θ on option prices appear to be always monotonic. In other words, as θ increases from zero to two, the option prices either increases monotonically for both the ATM options and OTM options or decreases monotonically for the ITM options. However, all option prices for the case of θ = 3 are less than those of θ ∈ [0, 2]. From Panel B, the features of the impacts of θ on the delta values are similar to those of option prices.

Table 2

Option prices and Delta values of American CI calls under the CEV model

S0Tqθ = −6θ = −4θ = −2θ = 0θ = 1θ = 2θ = 3
A. Option prices
951/215.02005.11575.24755.35365.49715.60842.0492
34.17844.29544.44224.54674.69674.80841.9848
82.41142.57252.74372.80542.96023.07131.8239
117.90217.97348.04328.14188.31858.48193.7060
36.32156.37606.49476.58626.77896.94483.5486
83.04263.22573.43943.44433.63873.79733.1552
1001/218.36948.21958.12578.02178.06508.07614.2314
37.53997.39027.31297.20097.24837.26064.1519
85.70585.59705.53805.38785.43965.45573.9533
1111.739911.512211.239011.028311.060011.08336.0024
310.21669.94109.68319.45689.50009.54025.8238
86.97326.64596.49716.19586.24606.42165.3772
1051/2112.232411.817911.484311.166111.100511.00927.1248
311.434411.010610.683310.354110.290510.19537.0582
89.67899.25368.94268.56788.50658.40446.8917
1115.921715.403614.779214.248114.131414.00928.8113
314.459913.883813.238812.688212.578612.44938.6492
811.377110.662310.07579.44199.33859.19958.2438
B. Delta values
951/210.61940.57420.53240.40070.42280.45480.4064
30.61850.56910.52830.39590.41720.44910.4093
80.58440.54050.50040.37420.39370.42570.4166
110.74280.67590.60800.42070.45580.49370.4349
30.73290.67410.60340.41600.44930.50800.4394
80.70580.62300.55980.38900.41760.62300.4507
1001/210.73770.68500.63650.50180.52120.55060.5054
30.74320.68770.63840.50200.52060.54970.5058
80.75300.68920.63830.50230.51830.54670.5066
110.81370.75400.68310.49090.52590.55830.5072
30.82620.76330.68550.49210.52530.55750.5079
80.85770.76330.68040.49430.52140.58780.5095
1051/210.81860.76660.71970.59250.61100.63890.5952
30.82560.77230.72330.59680.61460.64220.5928
80.84580.78670.73850.61390.62970.65680.5870
110.86620.81130.74230.55430.58990.62260.5714
30.87790.82240.74810.56030.59430.62610.5681
80.90840.85000.76290.58440.61300.64200.5598

Figure 1 plots optimal boundaries for American CI call option as functions of the time under the CEV model with different elasticities θ. The left panel in Figure 1 displays some exercise regions and stopping regions for θ = −6, −4, −2, 0. we can see that the exercise region area decreases and the exit region area increases with increasing in value of θ. This shows that the holders face with big risk. However, in the right panel of Figure 1 we can see that both the exercise region and the stopping region area increase with increasing in value of θ > 0(=1, 2, 3). This displays that both the opportunities the holders have had and the chances the holders have lost are symmetry.

Figure 1 Optimal boundaries for calls under the CEV model with different elasticities θ. Parameters used: S0 = 95, σ0 = 0.2, q = 1, T = 0.5, K = 100, r = 0.05 and δ = 0.04
Figure 1

Optimal boundaries for calls under the CEV model with different elasticities θ. Parameters used: S0 = 95, σ0 = 0.2, q = 1, T = 0.5, K = 100, r = 0.05 and δ = 0.04

5 Conclusion

This paper has examined the valuation of the American CI options under the CEV model which has the ability of capturing the implied volatility skew and can correct the pricing biases based on the BSM model. An integral representations for option price and two free boundaries of this option are obtained using the decomposition technique. We also gain the semi-closed form formulas for the Δ value of this option. Based on the numerical integration scheme [33, 34], we provide some numerical results, and analyze the impact of the elasticity coefficient on the option price and the behavior of the optimal boundaries.


Supported by the National Natural Science Foundation of China (11461008), the Humanities and Social Science Research Foundation of the Ministry of Education of China (13YJA910003), Guangxi Natural Science Foundation (2013GXNSFAA019005), and the Key Research Foundation of Science and Technology of the Education Department of Guangxi Province (2013ZD010)


References

[1] Davis M, Schachermayer W, Tompkins R. Pricing, no-arbitrage bounds and robust hedging of installment options. Social Science Electronic Publishing, 2001, 1(6): 597–610.10.2139/ssrn.262854Suche in Google Scholar

[2] Davis M, Schachermayer W, Tompkins R. Installment options and static hedging. Journal of Risk Finance, 2002, 3(2): 46–52.10.1007/978-3-0348-8291-0_12Suche in Google Scholar

[3] Ben-Ameur H, Breton M, François P. A dynamic programming approach to price installment options. European Journal of Operational Research, 2006, 169(2): 667–676.10.1016/j.ejor.2004.05.009Suche in Google Scholar

[4] Griebsch S, Kühn C, Wystup U. Instalment Options: A closed-form solution and the limiting case. Mathematical Control Theory and Finance, Springer, Heidelberg, 211-229, 2008.10.1007/978-3-540-69532-5_12Suche in Google Scholar

[5] Alobaidi G R, Mallier R, Deakin S. Laplace transforms and installment options. Mathematical Models and Methods in Applied Sciences, 2004, 14(6): 1167–1189.10.1142/S0218202504003581Suche in Google Scholar

[6] Ciurlia P, Roko I. Valuation of American continuous-installment options. Computational Economics, 2005, 25(1-2): 143–165.10.1007/s10614-005-6279-4Suche in Google Scholar

[7] Alobaidi G R, Mallier R. Installment options close to expiry. Journal of Applied Mathematics and Stochastic Analysis, 2006, 2006(1): 1–9.10.1155/JAMSA/2006/60824Suche in Google Scholar

[8] Kimura T. American continuous-installment options: valuation and premium decomposition. SIAM Journal on Applied Mathematics, 2007, 71(2): 517–539.10.1137/080740969Suche in Google Scholar

[9] Kimura T. Valuing continuous-installment options. European Journal of Operational Research, 2010, 201(1): 222–230.10.1016/j.ejor.2009.02.010Suche in Google Scholar

[10] Ciurlia P, Caperdoni C. A note on the pricing of perpetual continuous-installment options. Mathematical Methods in Economics and Finance, 2009, 4(1): 11–26.Suche in Google Scholar

[11] Ciurlia P. Valuation of European continuous-installment options. Computers & Mathematics with Applications, 2011, 62(6): 2518–2534.10.1016/j.camwa.2011.04.073Suche in Google Scholar

[12] Ciurlia P. An integral representation approach for valuing American-style installment options with continuous payment plan. Nonlinear Analysis: Theory & Methods and Applications, 2011, 74(16): 5506–5524.10.1016/j.na.2011.03.066Suche in Google Scholar

[13] Ciurlia P. On a general class of free boundary problems for European-style installment options with continuous payment plan. Communications on Pure & Applied Analysis, 2011, 10(4): 1205–1224.10.3934/cpaa.2011.10.1205Suche in Google Scholar

[14] Mezentsev A, Pomelnikov A, Ehrhardt M. Efficient numerical valuation of continuous installment options. Advances in Applied Mathematics & Mechanics, 2011, 3(2): 141–164.10.4208/aamm.10-m1025Suche in Google Scholar

[15] Huang G A, Deng G, Huang L H. Valuation for an American continuous-installment put option on bond under Vasicek interest rate model. Journal of Applied Mathematics and Decision Sciences, 2009, 2009(1): 1–11.10.1155/2009/215163Suche in Google Scholar

[16] Deng G. American continuous-installment options of barrier type. Journal of Systems Science and Complexity, 2014, 27(4): 928–949.10.1007/s11424-013-0310-ySuche in Google Scholar

[17] Yi F H, Yang Z, Wang X H. A variational inequality arising from European installment call options pricing. SIAM Journal on Mathematical Analysis, 2008, 40(1): 306–326.10.1137/060670353Suche in Google Scholar

[18] Yang Z, Yi F H. A variational inequality arising from American installment call options pricing. Journal of Mathematical Analysis & Applications, 2009, 357(1): 54–68.10.1016/j.jmaa.2009.03.045Suche in Google Scholar

[19] Black F, Scholes M. The pricing of options and corporate liabilities. Journal of Political Economy, 1973, 81(5–6): 637–654.10.1142/9789814759588_0001Suche in Google Scholar

[20] Merton R C. Theory of rational option pricing. The Bell Journal of Econoalca, 1973, 4(1): 141–183.10.2307/3003143Suche in Google Scholar

[21] Cox J. Notes on option pricing I: constant elasticity of variance diffusions. Working Paper, Stanford University, 1975 (reprinted in J. Portf. Manage., 1996, 22: 15–17).10.3905/jpm.1996.015Suche in Google Scholar

[22] Cox J, Ross S A. The valuation of options for alternative stochastic processes. J. of Financial Economics, 1976, 3(1): 145–166.10.1016/0304-405X(76)90023-4Suche in Google Scholar

[23] Boyle P P, Tian Y S. Pricing lookback and barrier options under the CEV process. Journal of Financial and Quantitative Analysis, 1999, 34(2): 241–264.10.2307/2676280Suche in Google Scholar

[24] Davydov D, Linetsky V. Pricing and hedging path-dependent options under the CEV process. Management Science, 2001, 47(7): 949–965.10.1287/mnsc.47.7.949.9804Suche in Google Scholar

[25] Emanuel D, MacBeth J. Further results on the constant elasticity of variance call option pricing model. Journal of Financial & Quantitative Analysis, 1982, 17(4): 533–554.10.2307/2330906Suche in Google Scholar

[26] Hsu Y L, Lin T I, Lee C F. Constant elasticity of variance (CEV) option pricing model: Integration and detailed derivation. Mathematics & Computers in Simulation, 2008, 79(1): 60–71.10.1016/j.matcom.2007.09.012Suche in Google Scholar

[27] Larguinho M, Dias J C, Braumann C A. On the computation of option prices and greeks under CEV model. Quantitative Finance, 2013, 13(6): 907–917.10.1080/14697688.2013.765958Suche in Google Scholar

[28] Detemple J, Tian W. The valuation of American options for a class of diffusion processes. Management Science, 2002, 48(7): 917–937.10.1287/mnsc.48.7.917.2815Suche in Google Scholar

[29] Nunes J P V. Pricing American options under the constant elasticity of variance model and subject to bankruptcy. Journal of Financial & Quantitative Analysis, 2009, 44(5): 1231–1263.10.1017/S0022109009990329Suche in Google Scholar

[30] Wong H Y, Zhao J. Valuing American options under the CEV model by Laplace-Carson transforms. Opeartions Research Letters, 2010, 38(5): 474–481.10.1016/j.orl.2010.07.006Suche in Google Scholar

[31] Ruas J P, Dias J C, Nunes J P. Pricing and static hedging of American options under the jump to default extended CEV model. Journal of Banking and Finance, 2013, 37(11): 4059–4072.10.1016/j.jbankfin.2013.07.019Suche in Google Scholar

[32] Deng G. Pricing American continuous-installment options under stochastic volatility model. Journal of Mathematical Analysis and Applications, 2015, 424(2): 802–823.10.1016/j.jmaa.2014.11.049Suche in Google Scholar

[33] Kim I J. The analytical valuation of American options. Review of Financial Studies, 1990, 3(4): 547–572.10.1093/rfs/3.4.547Suche in Google Scholar

[34] Kallast S, Kivinukk A. Pricing and hedging american options using approximations by Kim integral equations. European Finance Review, 2003, 7(3): 361–383.10.1023/B:EUFI.0000022128.44728.4cSuche in Google Scholar

[35] Sankaran M. Approximations to the non-central Chi-square distribution. Biometrica, 1963, 50(1–2): 199–204.10.1093/biomet/50.1-2.199Suche in Google Scholar

Appendix A

Let Z1, Z2, ⋯, Zv be v-dimensional independent standard normal random variables, and m1, m2, ⋯, mv be constants, then i=1v (Zi + mi)2 is the noncentral distribution with v degrees of freedom and noncentral parameter y = i=1vmi2, and is denoted by χv2 (y). Further, the cumulative distribution function of χv2 (y) is given by

χ2(x;v,y)=P(χv2(y))=ey/2j=0+(y/2)jj!2v/2Γ(v/2+j)0xzv/2+j1ez/2dz,x>0.(A.1)

And the probability density function of χv2 (λ) is

fχv2(y)(x)=e(x+y)/22v/2j=0+yjxv/2+j1j!22jΓ(v/2+j)=12(xy)(v2)/4e12(x+y)I(v2)/2(yx),(A.2)

where Ik(⋅) is the modified Bessel function of the first kind of order k. Obviously, the complementary distribution function of χv2(y) is Q(x;v,y)=P(χv2(y)x)=1χ2(x;v,y).

It is well known that

Q(SuDu|St=Dt)
=1Q(2k(u)Dt2θe(rδ)(2θ)(ut);22θ,2k(u)Du2θ), if θ<2N(lnDtDu+(rδ12σ2)(ut)σut), if θ=21Q(2k(u)Du2θ;2+2θ2,2k(u)Dt2θe(rδ)(2θ)(ut)), if θ>2.(A.3)

See [21] for θ < 2 and [25] for θ > 2.

We can see from Equation (A.2) that numerical computation for χ2(x;v, λ) is involved with the infinite sum. As for an analytic approximation, in [35], Sankaran showed that the complementary distribution function Q(x;v, y) is approximated by the normal cumulative distribution function N(⋅), that is

Q(x;v,y)N[1hp[1h+0.5(2h)mp](xv+y)hh2p(1+mp)],(A.4)

where

h=123(v+y)(v+3y)(v+2y)2,p=v+2y(v+y)2,m=(h1)(13h).
Received: 2015-8-8
Accepted: 2015-9-30
Published Online: 2016-4-25

© 2016 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 20.11.2025 von https://www.degruyterbrill.com/document/doi/10.21078/JSSI-2016-149-20/html
Button zum nach oben scrollen