Processing math: 100%
Editorial Special Issues

Variational models in elasticity

  • This contribution is part of the Special Issue: Variational Models in Elasticity
    Guest Editors: Lucia De Luca; Marcello Ponsiglione
    Link: www.aimspress.com/mine/article/5510/special-articles
  • Received: 17 June 2020 Accepted: 17 June 2020 Published: 29 June 2020
  • Citation: L. De Luca, M. Ponsiglione. Variational models in elasticity[J]. Mathematics in Engineering, 2021, 3(2): 1-4. doi: 10.3934/mine.2021015

    Related Papers:

    [1] Zhe Li, Xiao-Tian Wang . Valuation of bid and ask prices for European options under mixed fractional Brownian motion. AIMS Mathematics, 2021, 6(7): 7199-7214. doi: 10.3934/math.2021422
    [2] Shou-de Huang, Xin-Jiang He . Analytical approximation of European option prices under a new two-factor non-affine stochastic volatility model. AIMS Mathematics, 2023, 8(2): 4875-4891. doi: 10.3934/math.2023243
    [3] Kung-Chi Chen, Kuo-Shing Chen . Pricing green financial options under the mixed fractal Brownian motions with jump diffusion environment. AIMS Mathematics, 2024, 9(8): 21496-21523. doi: 10.3934/math.20241044
    [4] Shoude Huang, Xin-Jiang He . An analytical approximation formula for European option prices under a liquidity-adjusted non-affine stochastic volatility model. AIMS Mathematics, 2022, 7(6): 10364-10377. doi: 10.3934/math.2022577
    [5] Yao Fu, Sisi Zhou, Xin Li, Feng Rao . Multi-assets Asian rainbow options pricing with stochastic interest rates obeying the Vasicek model. AIMS Mathematics, 2023, 8(5): 10685-10710. doi: 10.3934/math.2023542
    [6] Shaoling Zhou, Huixin Chai, Xiaosheng Wang . Barrier option pricing with floating interest rate based on uncertain exponential Ornstein–Uhlenbeck model. AIMS Mathematics, 2024, 9(9): 25809-25833. doi: 10.3934/math.20241261
    [7] Jiajia Zhao, Zuoliang Xu . Calibration of time-dependent volatility for European options under the fractional Vasicek model. AIMS Mathematics, 2022, 7(6): 11053-11069. doi: 10.3934/math.2022617
    [8] Hijaz Ahmad, Muhammad Nawaz Khan, Imtiaz Ahmad, Mohamed Omri, Maged F. Alotaibi . A meshless method for numerical solutions of linear and nonlinear time-fractional Black-Scholes models. AIMS Mathematics, 2023, 8(8): 19677-19698. doi: 10.3934/math.20231003
    [9] Hanjie Liu, Yuanguo Zhu, Yiyu Liu . European option pricing problem based on a class of Caputo-Hadamard uncertain fractional differential equation. AIMS Mathematics, 2023, 8(7): 15633-15650. doi: 10.3934/math.2023798
    [10] Ming Yang, Yin Gao . Pricing formulas of binary options in uncertain financial markets. AIMS Mathematics, 2023, 8(10): 23336-23351. doi: 10.3934/math.20231186


  • Most of the existing option pricing models assume that the market is frictionless and completely liquid, which leads to the situation that the bid and ask prices are treated as the same when pricing various options. However, there are two kinds of prices listed in real markets, i.e., bid and ask prices. The former refers to the price at which traders sell options, while the latter represents the one at which traders buy options. This prompted the development of conic finance theory [1,2,3], which acknowledges the non-uniqueness of market prices. It basically assumes that the market acts as a central counterparty; it sells and buys assets at ask price and bid price, respectively. The difference between ask and bid prices is usually called the bid-ask spread, which is an indication of the market liquidity.

    In fact, the issue of liquidity has already become a hot topic in the area of risk management and has attracted a lot of research attention on the bid-ask economy. Madan and Cherny [1] solved bid and ask prices of European options analytically under a market liquidity model after introducing a single market stress level. However, a number of authors used market data to further study the model and found that the implied market liquidity is not constant, which is far from the assumption in Madan and Cherny's conic option pricing model [4,5]. It should be pointed out that the underlying asset price in the above literature is modeled by the simplest geometric Brownian motion (GBM) in order to obtain the explicit forms of the distribution function so that bid and ask prices can be further derived by numerical methods.

    It is well known that the GBM is not suitable for describing the price process of the underlying asset since the so-called implied volatility smile or skew has been widely observed. Many studies have extended the GBM model. Mehrdoust and Najafi [6] studied European option pricing under a fractional Black-Scholes model with a weak payoff function. Lin and He [7] proposed a regime switching fractional Black-Scholes model and obtained the European option pricing formula. Hassanzadeh and Mehrdoust [8] investigated option pricing under a multifactor uncertain volatility model. Nevertheless, one of the most famous modifications is the Heston stochastic volatility model [9], and there are also various modifications to the Heston model [10]. It should be noted that the Heston model with the square root specification cannot describe the nonlinear characteristics of financial time series well in practical applications, although it brings great convenience due to the analytical pricing formula of options. Thus, alternative models have been established, among which the GARCH diffusion model has received increasing attention. This is because the GARCH diffusion model is shown to be able to better describe financial time series [11,12]. We refer interested readers to Kaeck and Alexander [13] and the references therein for more details on the results associated with the GARCH diffusion model.

    On the other hand, the jump-diffusion model represents a refinement of the GBM process, as it effectively captures the discontinuous changes in the underlying stock returns [14]. Moreover, the double exponential jump-diffusion model proposed by Kou [15] is able to reflect high levels of skewness and leptokurtosis exhibited by financial data. Due to its analytical tractability, this model has received great attention from academia and industry ever since it was put forward. For example, Mehrdoust et al. [16] presented the valuation for European options by adding jumps into the Bi-Heston model. Huang and Guo [17] got the semi-analytic solution of vulnerable options by assuming that the price process of the underlying asset follows non-affine stochastic volatility with double exponential jump. Considering the stochastic behaviour and jump risks, Hu et al. [18] investigated the pricing of European crude oil options. Moreover, we should be aware that the spot interest rate plays a decisive role in modern financial industry and it changes stochastically in the market. Grzelak et al. [19], Recchioni and Sun [20], and Chen et al. [21] proved that it performs much better if the option pricing model replaces the constant interest rate with a stochastic one. Although He and Zhu [22] presented a closed-form series solution to European option prices when the volatility and interest rate are both stochastic, which is appealing, their assumption of the CIR interest rate model prevents the interest rate from going negative. This is inconsistent with actual situations, as the short-term government bond markets of the USA and Europe have already witnessed negative interest rates [23]. It was even claimed by Recchioni et al. [24] that models allowing interest rates to take negative values are able to improve the performance of option pricing and implied volatility forecasting.

    Considering all the features discussed above that are able to help improve model performance, we incorporate the Vasicek stochastic interest rate and double exponential jump into the GARCH diffusion model when pricing options in the bid-ask price economy. However, the probability distribution of the log-price process cannot be analytically obtained due to the complexity of the adopted dynamic processes. Fortunately, with the help of the Fourier-cosine method, or the COS method* [25,26], we are able to derive the probability density function by making use of the corresponding characteristic function. Once the density function is obtained, bid and ask prices can be straightforwardly computed with some numerical schemes, including Gaussian quadrature. The accuracy of our proposed approach is verified via numerical comparison with the Monte Carlo simulation, and the sensitivity analysis is also performed so that the effect of the market liquidity parameter on bid and ask prices is clear.

    *The COS method has been widely used in European and American options pricing due to its simplicity, high pricing accuracy, and high computational efficiency. The central idea of the COS method is to reconstruct the density function by the characteristic function of log-price.

    The remainder of the paper is as follows. Section 2 presents a hybrid option pricing model combining the Vasicek stochastic interest rate, GARCH diffusion volatility model, and double exponential jump together. An approximation to the characteristic function is derived in Section 3. In Section 4, using the Fourier cosine method, we obtain the density function of the underlying log-price, and, further, we derive bid-ask prices of European options. Results of numerical experiments are provided in Section 5, with the last section concluding the article.

    Consider a filtered probability space {Ω,F,{Ft}0tT,Q}, with Q being a risk-neutral probability measure. We assume that the market is frictionless, i.e., there are no transaction costs and the assets can be traded continuously. The underlying asset price process St, the volatility process vt, and the stochastic interest rate process rt under Q are

    dStSt=(rtλm)dt+vtdWs(t)+(eJ1)dNt, (2.1)
    dvt=kv(θvvt)dt+σvvtdWv(t), (2.2)
    drt=kr(θrrt)dt+σrdWr(t), (2.3)

    with dWs(t)dWv(t)=ρdt. It should be pointed out that the model can be more sophisticated if one incorporates the correlation between the underlying price and interest rate. However, introducing such correlation could break down the analytical tractability according to a number of different literature [19,27,28,29,30]. Thus, it remains an open question on how to effectively price options with the incorporation of such correlation, and we would like to leave this to future work. The mean reversion speed is denoted by κv, while θv and σv respectively represent the long-run mean and instantaneous volatility of volatility. λ is the constant intensity of the Poisson process Nt. We have that m=EQ(eJ1), where J is the jump size following an asymmetric double exponential distribution whose density function can be presented as

    f(J)=pη1eη1J1{J0}+qη2eη2J1{J<0},η1>1,η2>0,

    where p and q respectively represent the probability of jumping upward and that of jumping downward, with p+q=1. This further implies that m=pη1η11+qη2η2+11.

    Due to the existence of the stochastic interest rate, it is a natural treatment to make a measure transform to convert the price dynamics of the underlying asset under the original measure Q into those under the T-forward measure QT. In order to achieve this, we need to choose the T-discount bond price as the numeraire, which will be provided below. In particular, if we denote P(t,T) as the price of a risk-free zero-coupon bond maturing at time T, when the evolution of rt follows Eq (2.3), the P(t,T) can be formulated as

    P(t,T)=exp{Ar(τ)Br(τ)r(t)},

    where

    τ=Tt,B(τ)=1ekrτkr,

    and

    A(τ)=σ2r2k2rθr2k2rτ+k2rθrσ2rk2rBr(τ)+σ2r4k2rBr(2τ).

    The measure changing from the risk-neutral probability measure Q to the T-forward measure QT can be established by the following Radon-Nikodym derivative:

    dQTdQFT=Θ(T),

    where

    Θ(T)=exp(Ttr(z)dz)E[exp(Ttr(z)dz)Ft]=exp{1krTt(1ekr(Tz))σrdWr(z)12k2rTt(1ekr(Tz))2σ2rdz}.

    Thus, we can express the target model dynamics under QT as

    dStSt=(rtλm)dt+vtdWTs(t)+(eJ1)dNt, (2.4)
    dvt=kv(θvvt)dt+σvvt(ρdWs(t)+(1ρ2)dWv(t)), (2.5)
    drt=(krθrσ2rBr(τ)krrt)dt+σrdWr(t), (2.6)

    where dWv(t)dWv(t)=0.

    Letting xt=lnSt, we can transform Eqs (2.4)–(2.6) into the following form:

    dxt=(rvt2λm)dt+vtdWs(t)+(J1)dNt, (2.7)
    dvt=kv(θvt)dt+σvt(ρdWs(t)+(1ρ2)dWv(t)), (2.8)
    drt=(krθrσ2rBr(τ)krrt)dt+σrdWr(t). (2.9)

    In this subsection, the COS method is briefly introduced for the completeness of the paper. It is well-known that the price P(x,t) of a European option at time t is an expectation under the risk neutral measure according to the classic option pricing theory [31], which does not take bid-ask spread into account, i.e.,

    P(x,t)=er(Tt)E[P(y,T)x]=er(Tt)RP(y,T)f(yx)dy,

    where x=lnSt, y=lnST, r is the risk-free rate, and T is the maturity time. f(y|x) is the probability density function of the underlying process, and P(y,T) is the payoff function of the option at maturity.

    Without significantly losing accuracy, given the special choice of [a,b], we can obtain

    P(x,t)er(Tt)baP(y,T)f(yx)dy.

    The key point of the COS method is that the density f(y|x), which is unknown in most cases, is approximated by a Fourier-cosine series expansion on [a,b], i.e.,

    f(yx)=+k=0Ak(x)cos(kπyaba),

    where Ak(x)=2babaf(yx)cos(kπyaba)dy, for k=0,1,,N1.

    As a result,

    P(x,t)=er(Tt)baP(y,T)+k=0Ak(x)cos(kπyaba)dy=12(ba)er(Tt)+k=0Ak(x)Vk12(ba)er(Tt)N1k=0Ak(x)Vk,

    where Vk=2babaP(y,T)cos(kπyaba)dy, and the prime of is used to indicate that the first term of the summation should be multiplied by a weight of 1/2.

    Meanwhile, Ak(x)Fk(x). Since

    Φ1(ω)=baeiωxf(x)dxReiωxf(x)dx=Φ(ω),

    we have

    Ak(x)=2babaf(yx)cos(kπyaba)dy=2baRe{Φ1(kπba;x)exp(ikaπba)},Fk(x)=2baRe{Φ(kπba)exp(ikaπba)},

    where Re{} is an operator to take the real part.

    In summary,

    P(x,t)er(Tt)N1k=0Re{Φ(kπba;x)eikπaba}Vk,

    where Φ is the characteristic function. Detailed contents about the COS method can be found in Fang and Oosterlee [25].

    This subsection presents the joint characteristic function of the underlying log-price, stochastic volatility, and stochastic interest rate. The following theorem provides the solution to the joint characteristic function under the T-forward measure QT.

    Theorem 3.1. Given that the underlying asset price follows the dynamics in Eqs (2.7)–(2.9), the characteristic function for xT is given by

    Φ(x,v,r,τ;u)=exp{iux+C(τ,u)v+D(τ,u)r+E(τ,u)|x=xt,v=vt,r=rt},

    where

    C(τ,u)=α01eατβ2+β1eατ,D(τ,u)=iukr{1exp(krτ)},E(τ,u)=12θvC(τ,u)α3α2[β1τ+ln(β2+β1eατα)]14(iuθv+u2θv)τλmiuτ+H(τ,u),

    with

    α0=12(iu+u2),α1=32iuσvρθ12vkv,α2=θvσ2v,β1=α1+α2,β2=α1α2,α=α214α0α2,α3=14ρσviuθ32v+12kvθv,H(τ,u)=λΛ(u)τ+(iuθr1k2r(iuσ2r+12u2σ2r))τ+iukrθrk2rekrτ+1k3r(iuσ2r+12u2σ2r)(12e2krτ2ekrτ)+32k3r(iuσ2r+12u2σ2r)iukrθrk2r,

    and Tt,τ=Tt,i=1, Λ(u)=pη1η1iu+qη2η2+iu1.

    Proof. By applying the Feynman-Kac theorem, Φ(x,v,r,τ;u) satisfies the following partial integral-differential equation (PIDE):

    Φτ+(rv2λm)Φx+v22Φx2+kv(θvv)Φv+12σ2vv22Φv2+12σ2r2Φr2+v32σvρ2Φxv+(krθrσ2rBr(τ)krr)Φr+λ+[Φ(x+J)Φ(x)]f(J)dJ=0. (3.1)

    The boundary condition for Eq (3.1) is given by

    Φ(x,v,r,0;u)=eiuxT.

    According to several different studies [32,33,34], this PIDE has an exponential-affine solution of the form

    Φ(x,v,r,τ;u)=exp{iux+C(τ,u)v+D(τ,u)r+E(τ,u)},

    with boundary conditions

    C(0,u)=D(0,u)=E(0,u)=0.

    Moreover,

    +[Φ(x+J)Φ(x)]f(J)dJ=+[EQ[eiu(x+J)]EQ[eiux]]f(J)dJ=+[EQ[eiux(eiuJ1)]]f(J)dJ=+EQ[eiux]EQ[eiuJ1]f(J)dJ=Φ(x,v,r,τ;u)Λ(u), (3.2)

    where Λ(u)=pη1η1iu+qη2η2+iu1.

    Equation (3.1) is very difficult to solve since it is a nonlinear PDE, and thus we first linearize it approximately. The idea is to approximate v32,v2 in the PIDE using Taylor expansions around the long-run mean of variance as follows:

    v2=2θvvθ2v, (3.3)
    v32=32θ12vv12θ32v. (3.4)

    With the substitution of Eqs (3.2)–(3.4) into PDE (3.1), we obtain

    (Cτv+Dτr+Eτ)+(rtλmv2)iu+12v(iu)2+kv(θvv)C+12σ2rD2+(krθrσ2rBr(τ)krr)D+(32θ12vv12θ32v)σvρiuC+λΛ(u)+σ2v2(2θvvθ2v)C2=0.

    Denote

    G(τ,u)=λΛ(u)+12σ2rD2+(krθrσ2rBr(τ))D.

    Then, by matching coefficients, we can derive the following three ordinary differential equations:

    Cτ=σ2vθvC2+(32θ12vσvρiukv)C12(iu+u2), (3.5)
    Dτ=iukrD, (3.6)
    Eτ=λmiu+kvθvC12σ2vθ2vC212σvρθ32viuC+G(τ,u). (3.7)

    According to the boundary condition D(0,u)=0, we obtain

    D(τ,u)=iukr{1exp(krτ)}.

    Equation (3.5) is a Riccati equation, whose general solution can be presented as

    C(τ,u)=α01eατβ2+β1eατ,

    where

    α0=12(iu+u2),α1=32iuσvρθ12vkv,α2=θvσ2v,β1=α1+α2,β2=α1α2,α=α214α0α2.

    From Eq (3.5), we can rewrite Eq (3.7) as

    Eτ=θv2Cτ+(14ρσviuθ32v+12kvθv)C14(iuθv+u2θv)+G(τ,u)λmiu.

    Integrating both sides of the above-mentioned equation, we have the following result:

    E(τ,u)=12θvC(τ,u)α3α2[β1τ+ln(β2+β1eατα)]14(iuθv+u2θv)τλmiuτ+H(τ,u),

    where

    α3=14ρσviuθ32v+12kvθv,H(τ,u)=(iukrθra1k2r(iuσ2r+12u2σ2r))τ+iukrθrk2rekrτ+1k3r(iuσ2r+12u2σ2r)(12e2krτ2ekrτ)+λΛ(u)τ+32k3r(iuσ2r+12u2σ2r)iukrθrk2r.

    Combining (13) and the above two expressions, Theorem 3.1 follows.

    According to conic finance theory, there are two types of market prices. The minimal acceptable price for selling a claim X is known as the ask price aγ(X), while the maximal acceptable price for purchasing the claim is referred to as the bid price bγ(X). Then, the following lemma should be introduced. For more detailed information, one can refer to [1,2,3].

    Lemma 3.1. If we assume that X represents the cashflow of the claim at its expiry time T, the ask and bid prices of the claim are respectively determined by

    aγ(X)=inf{a:EΨγ[aexp(rT)X]0}=exp(rT)EΨγ[X]=exp(rT)+xdΨγ(FX(x))=exp(rT)[0(1Ψγ(1FX(x)))dx++0Ψγ(1FX(x))dx],

    and

    bγ(X)=sup{b:EΨγ[exp(rT)Xb]0}=exp(rT)EΨγ[X]=exp(rT)+xdΨγ(FX(x))=exp(rT)[0Ψγ(FX(x))dx++0(1Ψγ(FX(x)))dx],

    where r is the risk-free interest rate, Ψ is the probability distortion function, γ denotes the degree of the distortion, and FX(x) is the distribution of random X. In addition, according to Lemma 3.1, the ask and bid prices for European call options will be shown in the following theorem.

    Theorem 3.2. Given the MINMAXVAR distortion function, i.e., Ψγ(w)=1(1w11+γ)1+γ,γ0,w[0,1], it follows from Eqs (2.1)–(2.3) that the ask and bid prices for European call options can be expressed as

    aγ(C)=P(t,T)+K1(1(1FS(x))11+γ)1+γdx,bγ(C)=P(t,T)+K(1(FS(x))11+γ)1+γdx,

    where

    FS(x)=FlnS(lnx)=12˜A0(lnxa)+baπN1k=1˜Akksin(kπlnxaba),

    with ˜A0=2baRe{Φ(0)}. K denotes the strike price of the option, while aγ(C) and bγ(C) represent the ask and bid prices for the option, respectively.

    Proof. We can reconstruct the density function flnS of the random variable, i.e., the log-asset price lnST with a truncated region [a,b] using the results provided by Fang and Oosterlee [25] as

    flnS(x)2baN1k=0Re{Φ(kπba)eikπxaba}cos(kπxaba), (3.8)

    where Φ{} is the characteristic function of the density function f(x), which is obtained in Theorem 3.1, and Re{} is an operator to take the real part. It should be noted that the prime of is used to alert that the first term of the summation should be multiplied by a weight of 1/2.

    Meanwhile, let

    ˜Ak=2baRe{Φ(kπba)eikπxaba},k=1,2,,N1,

    where Φ{} is the characteristic function of the density function f(x) which is obtained in Theorem 3.1. Then, the distribution function FlnS of the log-asset price lnST is given by

    FlnS(y)=yflnS(x)dyyaflnS(x)dy=12˜A0(ya)+baπN1k=1˜Akksin(kπyaba).

    As a result, the distribution function FS of the underlying price ST can be directly obtained through

    FS(x)=FlnS(lnx)=12˜A0(lnxa)+baπN1k=1˜Akksin(kπlnxaba). (3.9)

    By employing Lemma 3.1 and the MINMAXVAR distortion function, European call ask and bid prices can be respectively expressed as

    aγ(C)=P(t,T)+K1(1(1FS(x))11+γ)1+γdx,bγ(C)=P(t,T)+K(1(FS(x))11+γ)1+γdx,

    where the formula of FS(x) is provided in Eq (3.9).

    Following Theorem 3.2, we can obtain the ask and bid prices of European put options as the following corollary.

    Corollary 3.1 If the European put option is also controlled by the stochastic differential equations (2.1)–(2.3), given the same distortion function in Theorem 3.2 and utilizing a similar derivation, the ask and bid prices of European put options can be expressed as

    aγ(P)=P(t,T)K0Ψγ(FS(x))dx=P(t,T)K01(1(FS(x))11+γ)1+γdx,bγ(P)=P(t,T)K0(1Ψγ(1FS(x)))dx=P(t,T)K0(1(1FS(x))11+γ)1+γdx.

    It should be remarked that the integrals involved in ask and bid prices given by Theorem 3.2 and Corollary 3.1 can be numerically computed with the Gaussian quadrature, which is one of the best quadratures with a high degree of accuracy and efficiency. This can be very easily implemented by the MATLAB built-in function quadgk, which ensures the speed for the implementation of the proposed approximation method when calculating ask and bid prices.

    In this subsection, we benchmark our results by making use of a Monte Carlo (hereafter, MC) simulation. All the computation is implemented using MATLAB 2016a on a computer equipped with an Intel Core i3 CPU @ 2.53 GHz.

    First, we follow Fang and Oosterlee [18, 19] to select [a,b] as

    [a,b]=[c1+x0Lc2+c4,c1+x0+Lc2+c4],

    when computing our approximation formula, where x0=lnS0,L=10, and cn is the n-th cumulant of lnST.

    Moreover, each sample path in the MC simulation is generated with the time interval being uniformly divided containing M1=252 points, and we use M2=100,000 as the number of sample paths. If the payoff produced by the i-th sample path is denoted by payoff(i), i=1,2,,M2, we can obtain

     bid =P(t,T)M2i=1[Ψγ(iM2)Ψγ(i1M2)] payoff (i),

    and

    ask=P(t,T)M2i=1[Ψγ(M2i+1M2)Ψγ(M2iM2)] payoff (i),

    In addition, we utilize a discrete scheme of stochastic differential equations (2.1)–(2.3) as follows:

    S(t+Δt)=St+St((rλm)Δt+vtε1Δt+(eJ11)(N(t+Δt)Nt),v(t+Δt)=vt+kv(θvvt)Δt+σvvt(ρε1+1ρ2ε2)Δt,r(t+Δt)=rt+kr(θrrt)Δt+σrε3Δt,

    where Δt=T252, εiN(0,1),i=1,,3, and ε1,ε2,ε3 are independent of each other. N(t) is a Poisson process with parameter λ. Then,

     payoff (i)=max{Si,TK,0},

    where Si,T is the price of the underlying asset which is produced by the i-th sample path at the maturity time T.

    With the values of other parameters provided in Table 1, bid and ask prices of European call options are shown in Tables 2 and 3 with different values of strike prices, respectively. The absolute relative error (Abs.R.E.) is defined as

     Abs.R.E. =|PcosPmc|Pmc×100%.
    Table 1.  Parameter values for the numerical experiments.
    Parameter r0 λ kv θv σv kr θr γ S0
    value 0.05 3 1.15 0.3 0.2 0.75 0.1 0.25 1
    Parameter p1 η1 η2 σr ρ K T v0
    value 0.3 10 5 0.25 0.7 1.5 1 0.25

     | Show Table
    DownLoad: CSV
    Table 2.  Comparisons of the CPU time and accuracy for Fourier-cosine method and the Monte Carlo simulation of the bid price of European call options.
    K T-t F.C. method MC Simulation Abs.R.E.
    0.9 1/4 0.1100 0.1050 4.61%
    0.9 1/2 0.1282 0.1332 3.75%
    cpu time(sec) 2.133 8.949
    1 1/4 0.0714 0.0676 5.62%
    1 1/2 0.0998 0.0958 4.17%
    cpu time(sec) 2.128 8.817
    1.1 1/4 0.0447 0.0421 6.18%
    1.1 1/2 0.0742 0.0711 4.36%
    cpu time (sec) 2.123 8.569

     | Show Table
    DownLoad: CSV
    Table 3.  Comparisons of the CPU time and accuracy for Fourier-cosine method and the Monte Carlo simulation of the ask price of European call options.
    K T-t F.C. method MC Simulation Abs.R.E.
    0.9 1/4 0.2922 0.2857 2.28%
    0.9 1/2 0.3931 0.3980 1.23%
    cpu time(sec) 2.813 8.443
    1 1/4 0.2235 0.2179 2.57%
    1 1/2 0.3349 0.3306 1.30%
    cpu time(sec) 2.798 8.689
    1.1 1/4 0.1671 0.1627 2.70%
    1.1 1/2 0.2801 0.2767 1.22%
    cpu time (sec) 2.832 8.784

     | Show Table
    DownLoad: CSV

    One should note that the results derived with our analytical approximation and those from the MC simulation are close to each other, which is a clear indication that the approximation is of high accuracy. Meanwhile, Figure 1 shows how the bid-ask prices of European call options change with different sample paths M2. It shows that the results of MC simulation usually fluctuate within a confidence interval. It also indicates that the results of MC simulation and our approach get close to each other as the number of paths increases.

    Figure 1.  The European call option price under different sample paths M2.

    In this subsection, the impact of parameter changes on bid and ask prices will be investigated, with a focus on: (i) The long-run mean level γ; (ii) The jump intensity λ and time to maturity Tt; (iii) The long-run mean level θv and θr.

    The effect of different γ on the bid-ask prices of European call options are shown in Figure 2. We find that, as the market liquidity indicator γ increases, or equivalently the liquidity of the market becomes lower, the bid-ask spread becomes larger and cannot be ignored. As γ approaches 0, the bid and ask prices will converge to a single price. This shows it is reasonable to consider the market liquidity in option pricing.

    Figure 2.  European call option price with different γ.

    Figure 3 demonstrates how European option prices are affected by different λ and Tt. Both bid and ask prices increase with the time to maturity, which is consistent with financial intuition since longer time implies larger time values of options. Larger jump intensity also contributes to greater option prices, since the underlying asset price becomes more volatile when there are possibly more jumps, which leads to higher risks and larger option premiums.

    Figure 3.  European call option price under different λ and Tt.

    What is plotted in Figure 4 is the influence of θv and θr on European option prices. It is clear that both θv and θr have a positive impact on European call option prices. This is reasonable since bigger θv means that the underlying asset prices are more volatile in the long run. Also, bigger θr indicates that the expected return of the underlying asset under the risk-neutral world is larger. Both will lead to higher option premiums.

    Figure 4.  European call option price under different θv and θr.

    This article aims to develop a pricing framework for European options within a hybrid GARCH-Vasicek model incorporating double exponential jumps in the bid-ask price economy. We derive analytical formulas for calculating bid and ask prices of European options utilizing the COS method, employing an approximation approach to obtain the characteristic function. We have also shown how both bid and ask prices vary with different values of parameters.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work is supported by the Creative Research Groups of the National Natural Science Foundation of Guizhou of China under Grants Qian Jiao HE KY ZI (No. 2019069), the National Natural Science Foundation of China (No. 12101554) and the Fundamental Research Funds for Zhejiang Provincial Universities (No. GB202103001).

    The authors would like to gratefully acknowledge the anonymous referees' constructive comments and suggestions, which greatly help to improve the readability and clarity of the manuscript.

    We confirm that there are no known conflicts of interest associated with this publication, and there has been no significant financial support for this work that could have influenced its outcome.



    [1] Alicandro R, De Luca L, Garroni A, et al. (2014) Metastability and dynamics of discrete topological singularities in two dimensions: A Γ-convergence approach. Arch Ration Mech Anal 214: 269-330.
    [2] Allaire G (2012) Shape Optimization by the Homogenization Method, Springer Science & Business Media.
    [3] Almi S, Lazzaroni G, Lucardesi I (2020) Crack growth by vanishing viscosity in planar elasticity. Mathematics in Engineering 2: 141-173.
    [4] Ambrosio L (1990) Existence theory for a new class of variational problems. Arch Ration Mech Anal 111: 291-322.
    [5] Ball JM (1976) Convexity conditions and existence theorems in nonlinear elasticity. Arch Ration Mech Anal 63: 337-403.
    [6] Ball JM, James RD (1987) Fine phase mixtures as minimizers of energy. Arch Ration Mech Anal 100: 13-52.
    [7] Bellettini G, Coscia A, Dal Maso G (1998) Compactness and lower semicontinuity properties in S BD(?). Math Z 228: 337-351.
    [8] Bethuel F, Brezis H, Hélein F (1994) Ginzburg-Landau Vortices, Boston: Birkhäuser.
    [9] Braides A, Defranceschi A (1998) Homogenization of Multiple Integrals, New York: Oxford University Press.
    [10] Canevari G, Zarnescu A (2020) Polydispersity and surface energy strength in nematic colloids. Mathematics in Engineering 2: 290-312.
    [11] Crismale V, Orlando G (2020) A lower semicontinuity result for linearised elasto-plasticity coupled with damage in W1, γ, γ > 1. Mathematics in Engineering 2: 101-118.
    [12] Dal Maso G (2012) An Introduction to Γ-Convergence, Springer Science & Business Media.
    [13] Dal Maso G (2013) Generalised functions of bounded deformation. J Eur Math Soc 15: 1943-1997.
    [14] De Philippis G, Rindler F (2020) Fine properties of functions of bounded deformation - an approach via linear PDEs, Mathematics in Engineering 2: 386-422.
    [15] Franfcfort GA, Marigo JJ (1998) Revisiting brittle fracture as an energy minimization problem. J Mech Phys Sol 46: 1319-1342.
    [16] Friedrich M (2020) Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering 2: 75-100.
    [17] Friesecke G, James RD, Müller S (2002) A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Commun Pure Appl Math 55: 1461-1506.
    [18] Giacomini A, Ponsiglione M (2006) A Γ-convergence approach to stability of unilateral minimality properties in fracture mechanics and applications. Arch Ration Mech Anal 180: 399-447.
    [19] Griffith AA (1921) The phenomena of rupture and flow in solids. Philos Trans Royal Soc A 221: 163-198.
    [20] Mateu J, Mora MG, Rondi L, et al. (2020) A maximum-principle approach to the minimisation of a nonlocal dislocation energy. Mathematics in Engineering 2: 253-263.
    [21] Novaga M, Pozzetta M (2020) Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering 2: 527-556.
    [22] Suquet PM (1978) Existence et régularité des solutions des équations de la plasticité. C R Acad Sci Paris Sér A 286: 1201-1204.
    [23] Tartar L (2009) The general theory of homogenization: A personalized introduction, Springer Science & Business Media.
    [24] Zeppieri CI (2020) Homogenisation of high-contrast brittle materials. Mathematics in Engineering 2: 174-202.
  • This article has been cited by:

    1. Indu Rani, Chandan Kumar Verma, Analyzing Short-Rate Models for Efficient Bond Option Pricing: A Review, 2024, 5, 2662-2556, 10.1007/s43069-024-00351-7
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3782) PDF downloads(766) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog