Parameter | value | Parameter | value |
r | 0.05 | k | 1.15 |
θ | 0.25 | σ | 0.76 |
ρ | -0.81 | β | 0.15 |
l | 0.5 | S0 | 10 |
v0 | 0.332 |
In this paper, we investigate the pricing of European options under a liquidity-adjusted non-affine stochastic volatility model. An analytical European option pricing formula is successfully derived with the COS method, based on an approximation for the characteristic function of the underlying log-asset price. Numerical analysis reveals that our results are very efficient and of reasonable accuracy, and we also present some sensitivity analysis to demonstrate the effects of different parameters on option prices.
Citation: Shoude Huang, Xin-Jiang He. An analytical approximation formula for European option prices under a liquidity-adjusted non-affine stochastic volatility model[J]. AIMS Mathematics, 2022, 7(6): 10364-10377. doi: 10.3934/math.2022577
[1] | 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 |
[2] | 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 |
[3] | Min-Ku Lee, Jeong-Hoon Kim . Pricing vanilla, barrier, and lookback options under two-scale stochastic volatility driven by two approximate fractional Brownian motions. AIMS Mathematics, 2024, 9(9): 25545-25576. doi: 10.3934/math.20241248 |
[4] | Min-Ku Lee, Jeong-Hoon Kim . Closed-form approximate solutions for stop-loss and Russian options with multiscale stochastic volatility. AIMS Mathematics, 2023, 8(10): 25164-25194. doi: 10.3934/math.20231284 |
[5] | Shoude Huang, Xinjiang He, Shuqu Qian . An analytical approximation of European option prices under a hybrid GARCH-Vasicek model with double exponential jump in the bid-ask price economy. AIMS Mathematics, 2024, 9(5): 11833-11850. doi: 10.3934/math.2024579 |
[6] | 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 |
[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] | 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 |
[9] | 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 |
[10] | Guiwen Lv, Ping Xu, Yanxue Zhang . Pricing of vulnerable options based on an uncertain CIR interest rate model. AIMS Mathematics, 2023, 8(5): 11113-11130. doi: 10.3934/math.2023563 |
In this paper, we investigate the pricing of European options under a liquidity-adjusted non-affine stochastic volatility model. An analytical European option pricing formula is successfully derived with the COS method, based on an approximation for the characteristic function of the underlying log-asset price. Numerical analysis reveals that our results are very efficient and of reasonable accuracy, and we also present some sensitivity analysis to demonstrate the effects of different parameters on option prices.
The Black-Scholes (BS) model [1] has been widely adopted in the pricing of financial derivatives because of its analytical tractability. However, it is generally known that the assumptions of BS model are not appropriate with empirical results. Therefore, many researchers devote much effort to remedying these shortcomings. A great deal of scholars proposed to use the generalized Black-Scholes model to describe the dynamic process of the underlying asset and solved the generalised Black-Scholes model numerically. For example, Mohammadi [1] developed a Quintic B-spline collocation approach for solving generalized Black-Scholes partial differential equation for European option prices. Roul [3] presented a high-order compact finite difference method based on a uniform mesh to obtain a highly accurate result for the generalized Black-Scholes equation. Roul [4] constructed a new computational approach for solving the generalized Black-Scholes equation numerically by combining Crank-Nicolson scheme and sextic B-spline collocation method. Another important improvement of the BS model is to make it a fractional one since it can describe the long-range dependence displayed by financial data. It should be noted that the analytical solution to option prices is generally not available under the framework of the fractional BS model, and numerical algorithms have to be used. Roul [5] designed a high-order numerical approach for solving the fractional BS equation. Roul and Goura [6] presented a numerical technique for solving the fractional BS equation describing European options.
Another main category that is very popular in modifying the BS model is to include stochastic volatility. A breakthrough was made by Heston [7], who proposed a model that not only possesses a range of basic properties including mean reversion and nonnegativity, but also relieve the volatility smile or smirk. Due to the popularity of the Heston model, a number of authors have worked under this framework. For example, Lee et al. [8] obtained an analytic formula for vulnerable options, while Mollapourasl et al. [9] used a radial basis function method to price American put options. He and Zhu [10] went further to introduce a stochastic interest rate into the Heston model and presented a closed-form formula for European options. In particular, the non-affine stochastic volatility model has attracted the extensive attention as an improvement of the Heston model. Christoffersen et al. [11] and Chourdakis [12] further confirmed that the non-affine volatility model represents the time series features of real option price data, which is much better than Heston model. This has prompted a number of authors to consider the non-affine volatility model. For example, Huang and Guo [13] investigated the valuation of discrete barrier options with non-affine volatility and double exponential jump, and Ma et al. [14] studied the pricing problem of vulnerable options with non-affine volatility and stochastic interest rate.
On the other hand, it needs to be pointed out that the market is assumed to be completely liquid in the above literature, whereas a lot of empirical evidence has already suggested that the market is actually not completely liquid [15,16]. Therefore, the concept of the market liquidity starts to appear in the option pricing theory. In particular, Feng et al. [17] derived an analytical solution to European option prices by incorporating a liquidity discounting factor into the option pricing model, based on which Li and Zhang [18] verified the influence of the market liquidity factor using Shanghai 50ETF options. Moreover, Li and Zhang [19] studied the pricing of European quanto options when the underlying foreign asset is in an imperfectly liquid market, while Xu et al. [20] focused on the variance and volatility swap pricing in the framework of a liquidity-adjusted underlying assets model. The pricing of vulnerable options is also considered by Wang [21] when the liquidity risk is assumed to be captured by a stochastic process. These clearly demonstrate that the effect of liquidity risk on option pricing is not negligible.
Motivated by the above research, we consider the pricing of European options under a liquidity-adjusted non-affine volatility model. The advantage of this model is that it can not only describe the nonlinear characteristics of financial data, but also take the liquidity risk into account. Although it is difficult to derive a closed-form solution for the characteristic function of the underlying log-asset price, we present an analytical approximation to the characteristic function using the first-order Taylor expansion method. Based on this, an approximate price for a European option can be obtained by the COS method [22], which is recently very popular in the area of option pricing [23,24,25,26].
The rest of this article is arranged as follows. A liquidity-adjusted non-affine volatility model is briefly introduced in the Section 2. In Section 3, an approximate formula for the characteristic function of the underlying log-price is derived, based on which an analytical formula for European option prices is presented with the COS method. Some numerical examples and sensitivity analysis are provided to interpret our results in Section 4, after which we conclude in the last section.
Let {Ω,Ft,Q} denote a given complete probability space, and Q is a risk-neutral probability measure. Considering that the non-affine model can explain the nonlinear characteristics of option data in real market, and liquidity-adjusted models perform better than those without considering liquidity risk, we construct a liquidity-adjusted non-affine volatility model for the valuation of European options. Specifically, the price process St and the process of the volatility vt under the risk-neutral world are modeled as follows:
dStSt=rdt+√β2l2+(1−ρ2)vtdW1t+ρ√vtdW2t, | (2.1) |
dvt=k(θ−vt)dt+σvtdW2t. | (2.2) |
where dW1tdW2t=ρdt, r is the risk-free interest rate, l is the level of market liquidity, β is the sensitivity of the stock to the level of market liquidity, k donates the mean-reversion speed, θ is the long-term mean of the volatility, and σ is the instantaneous volatility of the volatility.
With xt=ln(St/K), Eqs (2.1) and (2.2) can be respectively transformed into
dxt=(r−12β2l2−12vt)dt+√β2l2+(1−ρ2)vtdW1t+ρ√vtdW2t, | (2.3) |
dvt=k(θ−vt)dt+σvtdW2t. | (2.4) |
In this section, we approximate the characteristic function of the underlying log-price with the first-order Taylor expansion and based on this, European options can be analytically valuated using the COS method.
The characteristic function of xT is defined as
Φ(x,v,τ;u)=EQ[eiuxT|xt=x,vt=v], |
where T is the maturity time of the option, i is an imaginary unit, and τ=T−t. The solution to this characteristic function is presented in the following theorem.
Theorem 1. If the price process and its volatility process follow the dynamics specified in Eqs (2.1) and (2.2), then the characteristic function of xT can be expressed by
Φ(x,v,τ;u)=exp(iux+B(u,τ)v+A(u,τ)), |
where
B(u,τ)=α01−e−ατ−β2+β1e−ατ,A(u,τ)=[(r−12β2l2)iu+12β2l2(iu)2](T−t)−α3α2[β1τ+ln(−β2+β1e−ατα)]−12θB+12θ(−12(iu+u2))τ, |
and
α2=θσ2,α0=−12(iu+u2),α1=32θ12σρiu−k,β1=α1+α2,β2=α1−α2,α=√α21−4α0α2,α3=12kθ+14ρσiuθ32. |
Proof. The Feynman-Kac theorem indicates that Φ(x,v,τ;u) satisfies the following PDE (partial differential equation)
∂Φ∂t+∂Φ∂x(r−12β2l2−12v)+k(θ−v)∂Φ∂v+12(β2l2+v)∂2Φ∂x2+ρσv32∂2Φ∂x∂v+12σ2v2∂2Φ∂v2=0, | (3.1) |
with the boundary condition given by
Φ(x,v,0;u)=eiuxT. |
Given that the volatility in the proposed model is of mean reversion, it is reasonable to consider the expansion of v2 and v32 at the long-term mean θ using a first-order Taylor expansion. This gives
v2≈2θv−θ2, | (3.2) |
v32≈32θ12v−12θ32. | (3.3) |
Substituting Eqs (3.2) and (3.3) into the PDE (3.1) yields
∂Φ∂t+∂Φ∂x(r−12β2l2−12v)+k(θ−v)∂Φ∂v+12(β2l2+v)∂2Φ∂x2+ρσ(32θ12v−12θ32)∂2Φ∂x∂v+12σ2(2θv−θ2)∂2Φ∂v2=0. | (3.4) |
If we assume that the solution to this PDE can be written in the form of
Φ(x,v,τ;u)=exp(iux+A(u,τ)+B(u,τ)v), | (3.5) |
with the boundary conditions A(u,0)=B(u,0)=0, and substituting Eq (3.5) into Eq (3.4), we obtain
−(∂B∂τv+∂A∂τ)+(r−12β2l2−12v)iu+k(θ−v)B+12(β2l2+v)(iu)2+ρσ(32θ12v−12θ32)iuB+12σ2(2θv−θ2)B2=0. |
As a result, the following two ordinary differential equations can be derived
∂B∂τ=θσ2B2+(32θ12σρiu−k)B−12(iu+u2), | (3.6) |
∂A∂τ=(r−12β2l2)iu+12β2l2(iu)2+kθB−12ρσθ32iuB−12σ2θ2B2. | (3.7) |
Since Eq (3.6) is a Riccati equation, it can be solved with some algebraic computation so that B can be formulated as
B(u,τ)=α01−e−ατ−β2+β1e−ατ, |
where
α2=θσ2,τ=T−t,α0=−12(iu+u2),β1=α1+α2,β2=α1−α2,α=√α21−4α0α2,α1=32θ12σρiu−k. |
Multiplying both sides of Eq (3.6) by θ/2, and substituting the result into Eq (3.7), the following expression can be derived:
∂A∂τ=(r−12β2l2)iu+12β2l2(iu)2+(12kθ+14ρσiuθ32)B−12θ∂B∂τ+12θ(−12(iu+u2)), |
which further leads to
A=[(r−12β2l2)iu+12β2l2(iu)2]τ−α3α2[β1τ+ln(−β2+β1e−ατα)]−12θB+12θ(−12(iu+u2))τ, |
where
α3=12θk+14σρiuθ32. |
This completes the proof.
In this subsection, an analytical approximate price formula for the European options are obtained by the COS method, the results of which are summarized in the following theorem.
Theorem 2. Under the model dynamics (2.1) and (2.2), the pricing formula of European call option is expressed by:
P(x,v,τ)=e−rτN−1∑∑′k=0Re{Φ(kπb−a)eikπx−ab−a}Vcallk, |
where
Vcallk=2b−aK(χk(0,b)−ψk(0,b)), |
and the χ(x1,x2),ψ(x1,x2) is defined in Eqs (3.11) and (3.12).
Proof. It is well known that the price of a European option can be expressed as follows
P(x,v,τ)=e−rτ∫∞−∞v(y,T)f(y|x)dy, | (3.8) |
where x=ln(S0/K), y=ln(ST/K), and f(y|x) is the density function of y given x. The payoff function of European options at expiry P(y,T) can be expressed as
P(y,T)=g(y)=[αK(ey−1)]+,α={1,calloption,−1,putoption. |
Working out the price of a European option in (3.8) typically requires the information of the density function f(y|x), which is unknown and difficult to derive. Fortunately, there exists an approximate formula if the infinite domain of y is truncated into [a,b]
f(y|x)≈2b−aN−1∑∑′k=0Re{Φ(kπb−a)eikπx−ab−a}cos(kπy−ab−a), | (3.9) |
where ∑′ implies the summation whose first term is divided by two, Re{⋅} denotes taking the real part of a complex number, and Φ(u) is the conditional characteristic function of f(y|x).
Replacing f(y|x) in (3.8) with its approximation (3.9), and interchanging summation and integration, we can derive the the price of a European option:
P(x,v,τ)=e−rτN−1∑∑′k=0Re{Φ(kπb−a)eikπx−ab−a}Vk, | (3.10) |
where
Vk=2b−a∫bav(y,T)cos(kπy−ab−a)dy. |
Clearly, the only unknown term in the pricing formula is the coefficient Vk, deriving which yields a completely analytical solution. In particular, for a European call option, we have
χk(x1,x2)=∫x2x1excos(kπx−ab−a)dx=11+(kπb−a)2[cos(kπx2−ab−a)ex2−cos(kπx1−ab−a)ex1+kπb−asin(kπx2−ab−a)ex2+kπb−asin(kπx1−ab−a)ex1], | (3.11) |
and
ψk(x1,x2)=∫x2x1cos(kπx−ab−a)dx={b−akπ[sin(kπx2−ab−a)−sin(kπx1−ab−a)],k≠0,x2−x1,k=0. | (3.12) |
After some algebraic calculation, we can arrive at
Vcallk=2b−a∫b0K(ey−1)cos(kπy−ab−a)dy=2b−aK(χk(0,b)−ψk(0,b)), |
where ψk and χk are given by Eqs (3.11) and (3.12), respectively. This completes the proof.
By comparing the properties of European call and put options under the same model setup, we can find that the derivation of European put options is almost similar to the proof of Theorem 2, except for the calculation of the coefficient Vput(a1,b1). Therefore, without going to details, the corresponding European put option pricing formula is presented in the following corollary.
Corollary 1. Under the model dynamics (2.1) and (2.2), the pricing formula of European put option can be presented as:
P(x,v,τ)=e−rτN−1∑∑′k=0Re{Φ(kπb−a)eikπx−ab−a}Vputk, |
where
Vputk=2b−aK(−χk(a,0)+ψk(a,0)). |
Once an approximation formula is successfully derived, it is important to check its accuracy. Moreover, it is also interesting to investigate the influence of liquidity on option prices. Both of these will be addressed in this section.
In this subsection, numerical experiments are used to verify the pricing performance of the proposed method against the Monte Carlo simulation and the FFT method [14,27,28]. Without loss of the generality, we will only take European call options as an example.
Following Fang and Oosterlee (2008), [a,b] can be calculated as
[a,b]=[c1+a0−L√c2+√c4,c1+a0+L√c2+√c4] |
where a0=lnS0, L=10, and cn is the n-cumulant of lnST, whose computing formula is:
cn=1in∂n(lnΦ(u))∂un|u=0. |
We set the number of simulation paths 100,000, the number of time steps is chosen to be 252, N=210, and the damping factor is 2.55. The values of the remaining parameters* for numerical examples are displayed in Table 1. All of the numerical examples were implemented in Matlab R2020a on a PC with an Intel Core i5 CPU.
Parameter | value | Parameter | value |
r | 0.05 | k | 1.15 |
θ | 0.25 | σ | 0.76 |
ρ | -0.81 | β | 0.15 |
l | 0.5 | S0 | 10 |
v0 | 0.332 |
*These parameter values are selected for test purposes, the magnitude of which is consistent with a number of different literatures [10,17,20].
Table 2 presents our approximate prices, FFT and Monte Carlo simulation results of European call options with different strike prices and expiry times. The relative error (R.E.COS) between the COS method and Monte Carlo simulation is calculated as:
R.E.COS=|Pcos−Pmc|Pmc×100%, |
T | K | COS | MC | FFT | R.E.COS | R.E.FFT |
1/4 | 9 | 1.3902 | 1.3992 | 1.3848 | 0.64% | 1.03% |
9.5 | 1.0597 | 1.0656 | 1.0591 | 0.55% | 0.61% | |
10 | 0.7821 | 0.7839 | 0.7783 | 0.23% | 0.71% | |
10.5 | 0.5584 | 0.5555 | 0.5452 | 0.52% | 1.85% | |
11 | 0.3854 | 0.3785 | 0.3860 | 1.82% | 1.98% | |
1/2 | 9 | 1.7514 | 1.7614 | 1.7489 | 0.57% | 0.71% |
9.5 | 1.4507 | 1.4573 | 1.4504 | 0.46% | 0.47% | |
10 | 1.1862 | 1.1885 | 1.1838 | 0.19% | 0.40% | |
10.5 | 0.9576 | 0.9552 | 0.9479 | 0.25% | 0.76% | |
11 | 0.7633 | 0.7563 | 0.7649 | 0.93% | 1.14% | |
1 | 9 | 2.3443 | 2.3557 | 2.3419 | 0.48% | 0.56% |
9.5 | 2.0727 | 2.0811 | 2.0728 | 0.40% | 0.40% | |
10 | 1.8250 | 1.8296 | 1.8249 | 0.25% | 0.26% | |
10.5 | 1.6003 | 1.6009 | 1.5939 | 0.04% | 0.44% | |
11 | 1.3979 | 1.3944 | 1.3987 | 0.25% | 0.31% | |
5 | 9 | 5.0447 | 5.0690 | 5.0430 | 0.48% | 0.51% |
9.5 | 4.8697 | 4.8925 | 4.8614 | 0.47% | 0.64% | |
10 | 4.7021 | 4.7232 | 4.7013 | 0.45% | 0.46% | |
10.5 | 4.5416 | 4.5607 | 4.5817 | 0.42% | 0.46% | |
11 | 4.3879 | 4.4050 | 4.3863 | 0.39% | 0.42% | |
10 | 9 | 6.7740 | 6.8029 | 6.7727 | 0.42% | 0.44% |
9.5 | 6.6587 | 6.6864 | 6.6515 | 0.41% | 0.52% | |
10 | 6.5473 | 6.5736 | 6.5454 | 0.40% | 0.43% | |
10.5 | 6.4394 | 6.4645 | 6.5081 | 0.39% | 0.67% | |
11 | 6.3349 | 6.3588 | 6.3335 | 0.38% | 0.40% | |
Time(sec.) | 0.511 | 1.680 | 1.135 |
where Pcos and Pmc denote the prices of European options computed from Theorem 2 and Monte Carlo simulation. Similarly, the relative error (R.E.FFT) between the FFT and Monte Carlo simulation is defined as:
R.E.FFT=|PFFT−Pmc|Pmc×100%, |
where PFFT denotes the price of European options computed from the FFT method. It is not difficult to find that R.E.COS and R.E.FFT are both less than 2%, while R.E.COS is no greater than (R.E.FFT in all cases. In addition, Table 2 also indicates that the CPU time to calculate the prices of European options using the different methods. We can observe that the Monte Carlo simulation generally takes about three times as long as our approach, and the FFT method is also slower than our approach. All of these clearly reveal that our approach is quite accurate and efficient.
The relative error between the COS method and Monte Carlo simulation with respect to the initial price of the underlying asset S0 and the maturity time T−t is also shown in Figure 1 using a surface plot. One can see clearly that the relative error associated with in-the-money and at-the-money is larger than that of out-of-the money options. Moreover, the shorter the time, the larger the error. This encourages us to find more effective method to price short-tenor options in future research.
In this subsection, we will analyze the impact of market liquidity and stochastic volatility on European call option prices; such an impact will be studied through the influence of the following parameters: (1) the time varying level of market liquidity lt and the maturity time T−t, (2) the sensitivity of market liquidity β and the maturity time T−t, (3) long-term mean θ and risk-free interest rate r, (4) the initial price of the underlying asset S0 and the maturity time T−t.
Figure 2 shows the price of European call options with respect to the market liquidity level lt at different expiry times. Clearly, European call option price is an increasing function of the the maturity time T−t. In addition, a higher market liquidity level lt leads to higher European call option price. This is mainly because a large value of lt imply a greater level of illiquidity, which contributes to a higher underlying price and thus a larger option premium.
What are shown in Figure 3 are European call option prices with respect to the liquidity sensitivity factor β and the maturity time T−t, and the sensitivity parameter is shown to have a positive impact on option prices. This can be understood by the fact that for the same level of market liquidity, the underlying price will be affected more heavily by the market liquidity if it is more sensitive to the market liquidity, which in turn contributes to higher option prices.
Depicted in Figure 4 is how European call option price changes with respect to the long-term average of the volatility θ and risk-free interest rate r. Clearly, European call option prices are an increasing function of the risk-free interest rate, since raising the interest rate is equivalent to increasing the level of the underlying asset price. One can clearly observe that the price of European call options increases with the long-term average of volatility θ, which verifies that the price of the European call option is an increasing function of volatility. This is actually reasonable since a higher level of the long-term average of volatility tends to increase the long-term volatility level, which means that the uncertainty of the underlying asset prices increases, leading to a larger intrinsic value of the option and a higher option price.
Figure 5 plots European option prices against the underlying asset S0 and the maturity time T−t, and as we can observe, a higher level of the underlying asset price tends to increase European call option prices. The reason behind this is that a greater underlying asset price will raise the intrinsic value of the option and results in higher option prices.
This paper studies the pricing problem of European options under the non-affine volatility model in an illiquid market, so that the effect of market liquidity on option prices can be captured. Upon incorporating the market liquidity into the underlying dynamics, we present an approximation to the characteristic function of the underlying log-price, based on which an analytical pricing formula for European options is successfully derived using the COS method. Finally, the sensitivity analysis are used to demonstrate the validity of the model setup.
This research was financially supported by the Youth Growth Project of Guizhou Provincial Department of Education (Qianjiaohe No. KY [2022]035), 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 anonymous referees' constructive comments and suggestions, which greatly help to improve the readability of the manuscript.
The authors declare that there are no conflicts of interest regarding the publication of this paper.
[1] |
F. Black, M. Scholes, The pricing of option and corporate liabilities, J. Polit. Econ., 81 (1973), 637–654. https://doi.org/10.1086/260062 doi: 10.1086/260062
![]() |
[2] |
R. Mohammadi, Quintic B-spline collocation approach for solving generalized Black-Scholes equation governing option pricing, Comput. Math. Appl., 69 (2015), 777–797. https://doi.org/10.1016/j.camwa.2015.02.018 doi: 10.1016/j.camwa.2015.02.018
![]() |
[3] |
P. Roul, V. Goura, A new higher order compact finite difference method for generalised Black-Scholes partial differential equation: European call option, J. Comput. Appl. Math., 363 (2020), 464–484. https://doi.org/10.1016/j.cam.2019.06.015 doi: 10.1016/j.cam.2019.06.015
![]() |
[4] |
P. Roul, V. Goura, A sixth order numerical method and its convergence for generalized Black-Scholes PDE, J. Comput. Appl. Math., 377 (2020), 112881. https://doi.org/10.1016/j.cam.2020.112881 doi: 10.1016/j.cam.2020.112881
![]() |
[5] |
P. Roul, A high accuracy numerical method and its convergence for time-fractional Black-Scholes equation governing European options, Appl. Numer. Math., 151 (2019), 472–493. https://doi.org/10.1016/j.apnum.2019.11.004 doi: 10.1016/j.apnum.2019.11.004
![]() |
[6] |
P. Roul, V. Goura, A compact finite difference scheme for fractional Black-Scholes option pricing model, Appl. Numer. Math., 166 (2021), 40–60. https://doi.org/10.1016/j.apnum.2021.03.017 doi: 10.1016/j.apnum.2021.03.017
![]() |
[7] |
S. L. Heston, A closed-form solution for options with stochastic volatility with applications to bond and currency options, Rev. Financ. Stud., 6 (1993), 327–343. https://doi.org/10.1093/rfs/6.2.327 doi: 10.1093/rfs/6.2.327
![]() |
[8] |
M. K. Lee, S. J. Yang, J. H. Kim, A closed form solution for vulnerable options with Heston's stochastic volatility, Chaos Soliton. Fract., 86 (2016), 23–27. https://doi.org/10.1016/j.chaos.2016.01.026 doi: 10.1016/j.chaos.2016.01.026
![]() |
[9] |
R. Mollapourasl, A. Fereshtian, M. Vanmaele, Radial basis functions with partition of unity method for American options with stochastic volatility, Comput. Econ., 53 (2019), 259–287. https://doi.org/10.1007/s10614-017-9739-8 doi: 10.1007/s10614-017-9739-8
![]() |
[10] |
X. J. He, S. P. Zhu, A closed-form pricing formula for European options under the heston model with stochastic interest rate, J. Comput. Appl. Math., 335 (2018), 323–333. https://doi.org/10.1016/j.cam.2017.12.011 doi: 10.1016/j.cam.2017.12.011
![]() |
[11] |
P. Christoffersen, K. Jacobs, K. Mimouni, Models for S & P 500 dynamics: Evidence from realized volatility, daily returns, and option prices, Rev. Financ. Stud., 23 (2010), 3141–3189. https://doi.org/10.1093/rfs/hhq032 doi: 10.1093/rfs/hhq032
![]() |
[12] |
K. Chourdakis, G. Dotsis, Maximum likelihood estimation of non-affine volatility processes, J. Empir. Financ., 18 (2011), 533–545. https://doi.org/10.1016/j.jempfin.2010.10.006 doi: 10.1016/j.jempfin.2010.10.006
![]() |
[13] |
S. Huang, X. Guo, A Fourier-cosine method for pricing discretely monitored barrier options under stochastic volatility and double exponential jump, Math. Probl. Eng., 2020 (2020). https://doi.org/10.1155/2020/4613536 doi: 10.1155/2020/4613536
![]() |
[14] |
C. Q. Ma, S. J. Yue, H. Wu, Y. Ma, Pricing vulnerable options with stochastic volatility and stochastic interest rate, Comput. Econ., 56 (2020), 391–429. https://doi.org/10.1007/s10614-019-09929-4 doi: 10.1007/s10614-019-09929-4
![]() |
[15] |
R. K. Chou, S. L. Chung, Y. J. Hsiao, Y. H. Wang, The impact of liquidity on option prices, J. Futures Markets, 31 (2011), 1116–1141. https://doi.org/10.1002/fut.20531 doi: 10.1002/fut.20531
![]() |
[16] |
P. Christoffersen, M. Fournier, K. Jacobs, The factor structure in equity options, Rev. Financ. Stud., 31 (2018), 595–637. https://doi.org/10.1093/rfs/hhx089 doi: 10.1093/rfs/hhx089
![]() |
[17] |
S. P. Feng, M. W. Hung, Y. H. Wang, The importance of stock liquidity on option pricing, Int. Rev. Econ. Financ., 43 (2016), 457–467. https://doi.org/10.1016/j.iref.2016.01.008 doi: 10.1016/j.iref.2016.01.008
![]() |
[18] | Z. Li, Research on the option pricing with liquidity risk, Guang Zhou: South China University of Technology, 2018. |
[19] |
Z. Li, W. G. Zhang. Y. J. Liu, European quanto option pricing in presence of liquidity risk, N. Am. J. Econ. Financ., 45 (2018), 230–244. https://doi.org/10.1016/j.najef.2018.03.002 doi: 10.1016/j.najef.2018.03.002
![]() |
[20] |
D. X. Xu, B. Z. Yang, J. H. Kang, N. J. Huang, Variance and volatility swaps valuations with the stochastic liquidity risk, Phys. A, 566 (2021), 125679. https://doi.org/10.1016/j.physa.2020.125679 doi: 10.1016/j.physa.2020.125679
![]() |
[21] |
X. Wang, Pricing vulnerable options with stochastic liquidity risk, N. Am. J. Econ. Financ., 60 (2022), 101637. https://doi.org/10.1016/j.najef.2021.101637 doi: 10.1016/j.najef.2021.101637
![]() |
[22] |
F. Fang, C. W. Oosterlee, A novel pricing method for European options based on Fourier-cosine series expansions, Siam J. Sci. Comput., 31 (2008), 826–848. https://doi.org/10.1137/080718061 doi: 10.1137/080718061
![]() |
[23] |
F. Fang, C. W. Oosterlee, Pricing early-exercise and discrete barrier options by Fourier-cosine series expansions, Numer. Math., 114 (2009), 27–62. https://doi.org/10.1007/s00211-009-0252-4 doi: 10.1007/s00211-009-0252-4
![]() |
[24] |
F. Fang, C. W. Oosterlee, A Fourier-based valuation method for Bermudan and barrier options under Heston's model, Siam J. Financ. Math., 2 (2011), 439–463. https://doi.org/10.1137/100794158 doi: 10.1137/100794158
![]() |
[25] |
S. M. Zhang, J. H. Geng, Fourier-cosine method for pricing forward starting options with stochastic volatility and jumps, Commun. Stat-Theor. M., 46 (2017), 9995–10004. https://doi.org/10.1080/03610926.2016.1228960 doi: 10.1080/03610926.2016.1228960
![]() |
[26] |
W. G. Yu, Y. D. Yong, G. F. Guan, Y. J. Huang, W. Su, C. R. Cui, Valuing guaranteed minimum death benefits by cosine series expansion, Math.-Basel, 7 (2019), 835–850. https://doi.org/10.3390/math7090835 doi: 10.3390/math7090835
![]() |
[27] | P. Carr, D. Madan, Option valuation using the fast Fourier transform, J. Comput. Financ., 2 (1999), 61–73. |
[28] |
C. X. Feng, J. Tan, Z. Y. Jiang, S. Chen, A generalized European option pricing model with risk management, Phys. A, 545 (2020), 123797. https://doi.org/10.1016/j.physa.2019.123797 doi: 10.1016/j.physa.2019.123797
![]() |
Parameter | value | Parameter | value |
r | 0.05 | k | 1.15 |
θ | 0.25 | σ | 0.76 |
ρ | -0.81 | β | 0.15 |
l | 0.5 | S0 | 10 |
v0 | 0.332 |
T | K | COS | MC | FFT | R.E.COS | R.E.FFT |
1/4 | 9 | 1.3902 | 1.3992 | 1.3848 | 0.64% | 1.03% |
9.5 | 1.0597 | 1.0656 | 1.0591 | 0.55% | 0.61% | |
10 | 0.7821 | 0.7839 | 0.7783 | 0.23% | 0.71% | |
10.5 | 0.5584 | 0.5555 | 0.5452 | 0.52% | 1.85% | |
11 | 0.3854 | 0.3785 | 0.3860 | 1.82% | 1.98% | |
1/2 | 9 | 1.7514 | 1.7614 | 1.7489 | 0.57% | 0.71% |
9.5 | 1.4507 | 1.4573 | 1.4504 | 0.46% | 0.47% | |
10 | 1.1862 | 1.1885 | 1.1838 | 0.19% | 0.40% | |
10.5 | 0.9576 | 0.9552 | 0.9479 | 0.25% | 0.76% | |
11 | 0.7633 | 0.7563 | 0.7649 | 0.93% | 1.14% | |
1 | 9 | 2.3443 | 2.3557 | 2.3419 | 0.48% | 0.56% |
9.5 | 2.0727 | 2.0811 | 2.0728 | 0.40% | 0.40% | |
10 | 1.8250 | 1.8296 | 1.8249 | 0.25% | 0.26% | |
10.5 | 1.6003 | 1.6009 | 1.5939 | 0.04% | 0.44% | |
11 | 1.3979 | 1.3944 | 1.3987 | 0.25% | 0.31% | |
5 | 9 | 5.0447 | 5.0690 | 5.0430 | 0.48% | 0.51% |
9.5 | 4.8697 | 4.8925 | 4.8614 | 0.47% | 0.64% | |
10 | 4.7021 | 4.7232 | 4.7013 | 0.45% | 0.46% | |
10.5 | 4.5416 | 4.5607 | 4.5817 | 0.42% | 0.46% | |
11 | 4.3879 | 4.4050 | 4.3863 | 0.39% | 0.42% | |
10 | 9 | 6.7740 | 6.8029 | 6.7727 | 0.42% | 0.44% |
9.5 | 6.6587 | 6.6864 | 6.6515 | 0.41% | 0.52% | |
10 | 6.5473 | 6.5736 | 6.5454 | 0.40% | 0.43% | |
10.5 | 6.4394 | 6.4645 | 6.5081 | 0.39% | 0.67% | |
11 | 6.3349 | 6.3588 | 6.3335 | 0.38% | 0.40% | |
Time(sec.) | 0.511 | 1.680 | 1.135 |
Parameter | value | Parameter | value |
r | 0.05 | k | 1.15 |
θ | 0.25 | σ | 0.76 |
ρ | -0.81 | β | 0.15 |
l | 0.5 | S0 | 10 |
v0 | 0.332 |
T | K | COS | MC | FFT | R.E.COS | R.E.FFT |
1/4 | 9 | 1.3902 | 1.3992 | 1.3848 | 0.64% | 1.03% |
9.5 | 1.0597 | 1.0656 | 1.0591 | 0.55% | 0.61% | |
10 | 0.7821 | 0.7839 | 0.7783 | 0.23% | 0.71% | |
10.5 | 0.5584 | 0.5555 | 0.5452 | 0.52% | 1.85% | |
11 | 0.3854 | 0.3785 | 0.3860 | 1.82% | 1.98% | |
1/2 | 9 | 1.7514 | 1.7614 | 1.7489 | 0.57% | 0.71% |
9.5 | 1.4507 | 1.4573 | 1.4504 | 0.46% | 0.47% | |
10 | 1.1862 | 1.1885 | 1.1838 | 0.19% | 0.40% | |
10.5 | 0.9576 | 0.9552 | 0.9479 | 0.25% | 0.76% | |
11 | 0.7633 | 0.7563 | 0.7649 | 0.93% | 1.14% | |
1 | 9 | 2.3443 | 2.3557 | 2.3419 | 0.48% | 0.56% |
9.5 | 2.0727 | 2.0811 | 2.0728 | 0.40% | 0.40% | |
10 | 1.8250 | 1.8296 | 1.8249 | 0.25% | 0.26% | |
10.5 | 1.6003 | 1.6009 | 1.5939 | 0.04% | 0.44% | |
11 | 1.3979 | 1.3944 | 1.3987 | 0.25% | 0.31% | |
5 | 9 | 5.0447 | 5.0690 | 5.0430 | 0.48% | 0.51% |
9.5 | 4.8697 | 4.8925 | 4.8614 | 0.47% | 0.64% | |
10 | 4.7021 | 4.7232 | 4.7013 | 0.45% | 0.46% | |
10.5 | 4.5416 | 4.5607 | 4.5817 | 0.42% | 0.46% | |
11 | 4.3879 | 4.4050 | 4.3863 | 0.39% | 0.42% | |
10 | 9 | 6.7740 | 6.8029 | 6.7727 | 0.42% | 0.44% |
9.5 | 6.6587 | 6.6864 | 6.6515 | 0.41% | 0.52% | |
10 | 6.5473 | 6.5736 | 6.5454 | 0.40% | 0.43% | |
10.5 | 6.4394 | 6.4645 | 6.5081 | 0.39% | 0.67% | |
11 | 6.3349 | 6.3588 | 6.3335 | 0.38% | 0.40% | |
Time(sec.) | 0.511 | 1.680 | 1.135 |