MC/QMC Methods for Option Pricing under Stochastic Volatility Models Chuan-Hsiang Han Department of Quantitative Finance, National Tsing-Hua University, Hsinchu,Taiwan 30013, ROC. E-mail: [email protected] Yongzeng Lai Department of Mathematics, Wilfrid Laurier University, Waterloo, Ontario, N2L 3C5, Canada E-mail: [email protected] Abstract In the context of multi-factor stochastic volatility models, which contain the widely used Heston model, we present variance reduction techniques to price European options by Monte Carlo (MC) and QuasiMonte Carlo (QMC) methods. We formulate a stochastic integral as a martingale control for the payoffs to be evaluated. That control corresponds to the cost of an approximate delta hedging strategy. Putting together this control variate method and randomized quasi-Monte Carlo methods, including the Sobol’ sequence and L’Ecuyer type good lattice points, we find that the variance reduction ratios can reach up to 300 times for European options under multi-factor stochastic volatility models. Keywords: Option pricing; Multi-factor stochastic volatility models; control variate method; Monte Carlo and quasi-Monte Carlo methods. 1 Introduction Stochastic volatility models have been an important class of diffusions as an extend of the BlackScholes models. Under stochastic volatility models, closed-form solutions for typical option pricing problems such as European and barrier options barely exist. Numerical methods become essential. Along the development of Monte Carlo methods, there have been some efficient and robust numerical methods proposed within last decade in order to reduce variances. In particular, many of these methods bridge out the concept of derivative pricing and imperfect hedging under incomplete market models. Therefore, reduced variances obtained from these Monte Carlo methods are not only measurements of computing efficiency but also represent risks associated to particular hedging strategy adopted in the algorithms. To the best of authors’ knowledge, Clewlow and Carverhill [4] are the first to utilize the value of some delta hedging portfolio and possibly other Greeks hedge as controls for discounted payoffs to evaluate option prices. They found that the variance obtained by this control variate method under an one-factor stochastic volatility model can be significantly reduced. By means of the least-squares method [22], Potters, Bouchaud, and Sestovic [25] proposed an optimal hedged Monte Carlo method to price options under the ”historical” probability measure. Later, Pochart and Bouchaud [27] extended the same idea to problems when shortfall risk and transaction cost constraints are considered. Heath and Platen [13] and Fouque and Han [6] used different approximate European option prices to approximate delta portfolio under stochastic volatility models and showed significant variance reduction performance. Moreover, in [7], Fouque and Han obtained an asymptotic result to characterize 1 the variance of a martingale control variate under stochastic volatility models when the time scales of driving volatility processes are well-separated. Applications of Barrier options and American options are also presented. Some other variance reduction techniques with less or no financial implication include conditional Monte Carlo [31], importance sampling [11, 5], direct sampling [3], etc. All Monte Carlo methods mentioned so far are fundamentally related to psudo random sequences. However integration methods using the alternative quasi-random sequences (or called low-discrepancy sequences) have drawn lots of attentions in recent years because its theoretical rate of convergence is O(1/n1−ε ) for all ε > 0 subjected to the dimensionality and the regularity of the integrand. A brief introduction about quasiMonte Carlo (QMC in short) methods, or low-discrepancy methods in Section 5. Also, we refer to [12], [17] and [18] for further discussions on Monte Carlo and Quasi-Monte Carlo methods in applications of computational finance. In this paper, we study several possible combinations of control variate methods and quasi-Monte Carlo methods for the pricing problems of European options under multi-factor stochastic volatility models. We conduct numerical experiments to compare the variance-reduction performance for the martingale control variate method mentioned above with or without randomized quasi-Monte Carlo methods. The organization of this paper is the following. In Section 2, we review the martingale control variate method for European options under multi-factor stochastic volatility models. These models are adequate to capture long-time and short-time volatility effects as documented in [1, 10, 23] and references therein. Additionally we take Heston’s model as a simplified example and obtain a dramatic variance reduction performance. In Section 3. Numerical experiments by Monte Carlo method and quasi-Monte Carlo methods are presented. We test several combinations of martingale control variate methods with or without quasi-Monte Carlo methods, including the Sobol’ sequence and L’Ecuyer type good lattice points together with the Brownian bridge sampling technique. 2 Monte Carlo Pricing under Multi-factor Stochastic Volatility Models Under a risk-neutral probability measure IP ? parametrized by the combined market price volatility premium (Λ1 , Λ2 ) , we consider the following multi-factor stochastic volatility models defined by (0)∗ dSt = rSt dt + σt St dWt , σt = f (Yt , Zt ), q g1 (Yt ) g1 (Yt ) 1 (0)∗ (1)∗ ρ1 dWt + 1 − ρ21 dWt , dYt = c1 (Yt ) + √ Λ1 (Yt , Zt ) dt + √ ε ε ε h i √ dZt = δc2 (Zt ) + δg2 (Zt )Λ2 (Yt , Zt ) dt q √ (0)∗ (1)∗ (2)∗ 2 2 + δg2 (Zt ) ρ2 dWt + ρ12 dWt + 1 − ρ2 − ρ12 dWt , (1) where St is the underlying asset price process with a constant risk-free interest rate r. Its random volatility σt is drivenby two stochastic processes Yt and Zt varying on the time scales ε and 1/δ, respec (0)∗ (1)∗ (2)∗ tively. The vector Wt , Wt , Wt consists of three independent standard Brownian motions. 2 Instant correlation coefficients ρ1 , ρ2 , and ρ12 satisfy |ρ1 | < 1 and |ρ22 + ρ212 | < 1. The volatility function f is assumed to be smooth bounded. Coefficient functions of processes Yt and Zt , namely (c1 , g1 , Λ1 ) and (c2 , g2 , Λ2 ) are assumed to be smooth such that they satisfy the existence and uniqueness conditions for stochastic differential equations. Mean-reverting processes such as Ornstein-Uhlenbeck (OU) processes or square-root processes are typical examples to model driving volatility processes [1, 10, 14]. Under this setup, the joint process (St , Yt , Zt ) is Markovian. Given the multi-factor stochastic volatility model (1), the price of a plain European option with the integrable payoff function H and expiry T is defined by P ε,δ (t, x, y, z) = IE ? t,x,y,z e−r(T −t) H(ST ) , where IE ? t,x,y,z is a short notation for the expectation with respect to IP ? conditioning on the current states St = x, Yt = y, Zt = z. A basic Monte Carlo simulation estimates the option price P ε,δ (0, S0 , Y0 , Z0 ) at time 0 by the sample mean N 1 X −rT (i) e H(ST ) N i=1 (2) (i) where N is the total number of sample paths and ST denotes the i-th simulated stock price at time T . Variance reduction techniques are particularly important to accelerate the computing efficiency of the basic Monte Carlo pricing estimator (2). Next we briefly review the construction of a generic algorithm, i.e. martingale control variate method, recently proposed by Fouque and Han [6, 7]. 2.1 Construction of Martingale Control Variates Assuming that the European option price P ε,δ (t, St , Yt , Zt ) is twice differentiable in state space and once differentiable in time, we apply Ito’s lemma to its discounted price e−rt P ε,δ , then integrate from time 0 to the maturity T . After canceling out the pricing partial differential equation in some nonmartingale terms, the following martingale representation can be obtained √ 1 P ε,δ (0, S0 , Y0 , Z0 ) = e−rT H(ST ) − M0 (P ε,δ ) − √ M1 (P ε,δ ) − δM2 (P ε,δ ), ε where centered martingales are given by Z T ∂P ε,δ ε,δ M0 (P ) = e−rs (s, Ss , Ys , Zs )f (Ys , Zs )Ss dWs(0)∗ , ∂x 0 Z T ∂P ε,δ (s, Ss , Ys , Zs )g1 (Ys )dW̃s(1)∗ , M1 (P ε,δ ) = e−rs ∂y 0 Z T ∂P ε,δ (s, Ss , Ys , Zs )g2 (Zs )dW̃s(2)∗ , M2 (P ε,δ ) = e−rs ∂z 0 where the Brownian motions are q 1 − ρ21 Ws(1)∗ , q (0)∗ (1)∗ = ρ2 Ws + ρ12 Ws + 1 − ρ21 − ρ212 Ws(2)∗ . W̃s(1)∗ = ρ1 Ws(0)∗ + W̃s(2)∗ 3 (3) (4) (5) (6) These martingales play the role of “perfect” controls for Monte Carlo simulations. Namely, if the martingales (4), (5), and(6) can be exactly computed, then one can just generate one sample ε,δpathε,δto evalu∂P ∂P ε,δ ∂P ate the option price through Equation (3). Unfortunately the gradient components ∂x , ∂y , ∂z of the option price appearing in the martingales is not possibly known in advance when the option price P ε,δ itself is exactly what we want to estimate. However, one can choose an approximate option price to substitute P ε,δ used in the martingales (4, 5, 6) and still retain martingale properties. When time scales ε and 1/δ are well separate; namely, 0 < ε, δ 1, an approximation of the BlackScholes type is derived in [10] P ε,δ (t, x, y, z) ≈ PBS (t, x; σ̄(z)) (7) √ √ ε, δ for continuous piecewise payoffs. We denote by with its pointwise accuracy of order O PBS (t, x; σ̄(z)) the solution of the Black-Scholes partial differential equation with the constant volatility σ̄(z) and the terminal condition PBS (T, x) = H(x). The z-dependent effective volatility σ̄(z) is defined as the square root of an averaging of the variance function f 2 with respect to a limiting distribution of Yt : Z 2 (8) σ̄ (z) = f 2 (y, z)dΦ(y) = f 2 (y, z) , where Φ(y) denotes the invariant distribution of the fast varying process Yt while setting the volatility premium Λ1 as zero. We use the √ bracket to represent such average. In the OU case, we choose that c1 (y) = m1 − y and g1 (y) = ν1 2 with Λ1 = 0 such that 1/ε is the rate of mean reversion, m1 is the long run mean, and ν1 is the long run standard deviation. Its invariant distribution Φ is normal with mean m1 and variance ν12 . We refer to [10] for detailed discussions of such models. Note that the approximate option price PBS (t, x; σ̄(z)) is independent of the variable y such that √ the (i) term M1 (PBS ) diminishes. Since the approximate martingale M2 (PBS ) for (6) is small of order δ. Intuitively we can neglect this term as well. We then select the stochastic integral M0 (PBS ) as a control for the estimator (2) and formulate the following martingale control variate estimator: N i 1 X h −rT (i) (i) e H(ST ) − M0 (PBS ) . N i=1 (9) This is the approach taken by Fouque and Han [6], in which the proposed martingale control variate method is numerically superior to an importance sampling method in [5] for pricing European options. Under OU-type processes to model (Yt , Zt ) in (1) with 0 < ε, δ 1, the variance of the martingale control variate for European options is small of order ε and δ. Recall the asymptotic result shown in [7] Theorem 1 Given the payoff function H being a call or put, the driving volatility processes Yt and Zt are of OU type, and the volatility function f (y, z) are smooth bounded, for any fixed initial state (0, x, y, z), there exist ε > 0 and δ > 0 small enough and a positive constant C such that V ar e−rT H (ST ) − M0 (PBS ) ≤ C max{ε, δ}. Moreover the financial interpretation of the martingale control term Z T ∂PBS M0 (PBS ) = e−rs (s, Ss ; σ̄(Zs ))f (Ys , Zs )Ss dWs(0)∗ ∂x 0 corresponds to the cumulative cost of a delta hedging strategy. This martingale control variate method can be easily applied to hitting time problems like barrier options and optimal stopping time problems like American options as discussed in [7]. 4 2.2 An Example: Martingale Control Variate Method for Heston’s Model Another widely applied mean-reverting process is Heston model [14] which incorporates a square-root diffusion to model the random volatility process. Heston model is an one-factor stochastic volatility model so that it can be viewed as a special case of our multi-factor SV model (1) after the square-root process is rescaled in the following way: q 2κ p 1 (1)∗ (0)∗ 2 (10) Yt ρ1 dWt + 1 − ρ1 dWt dYt = (θ − Yt ) dt + √ ε ε √ and Zt = 0 for all t ≥ 0. The volatility function is defined by f (y, z) = y. The driving volatility (10) admits a limiting distribution while ε approaches zero. It is known as the central Chi-square random variable κ2 χν (0) (11) with the degree of freedom ν = θ/κ2 . Its density function is denoted by x ν/2−1 −x 1 1 exp φ(x) = 2 ν/2 ν κ 2 Γ( 2 ) κ2 2κ2 (12) for x > 0, where Γ(·) is the Gamma function. Under the risk-neutral pricing probability measure, the European option price is P ε (t, x, y) = IE ? t,x,y e−r(T −t) H(ST ) . Following the singular perturbation expansion [9], one can derive an asymptotic result P ε (t, x, y) ≈ PBS (t, x; σ̄) (13) when Rε 1. As defined in (8) but without the z-component, the effective volatility σ̄ is defined from σ̄ 2 = f (y)2 φ(y)dy. By change of variable z = κy2 , the effective variance σ̄ 2 κ2 = ν/2 ν 2 Γ( 2 ) Z ∞ z ν/2 e−z/2 dz 0 2 = κν = θ (14) is deduced. We readily observe that the effective variance σ̄ 2 is exactly the long-run mean θ of the square root process (10). Under the one-factor Heston model specified in Table 1 and 2, the effective variance is σ̄ 2 = θ = 0.3. Several variance ratios are illustrated in Table 3 and show significant reduction gained from the martingale control variate estimator (9) versus the basic Monte Carlo estimator (2). Notice that the variance reduction ratios performs better when ε are either small (close to 1/50) or large (close to 1/0.1). In the small ε regime, one can use singular perturbation analysis to argue that the approximate option price PBS (t, x; σ̄) is close to the true option price. Hence the variance reduction is effective. In the large ε regime, the process Yt is slowly moving √ √ around Y0 = 0.3. This initial value is particularly chosen so that the driving volatility process Y t is close to the effective volatility σ̄ = θ when 0 ≤ t ε. Therefore, one can apply regular perturbation argument to explain the accuracy of the option price approximation and the variance reduction effect. Both singular and regular perturbation analysis can be found in [10]. 5 Table 1: Parameters used in the rescaled Heston’s model ( 10). r 5% θ 0.09 κ 0.1 ρ -0.3 f (y) √ y Table 2: Initial conditions and European call option parameters. $S0 50 Y0 0.09 $K 50 T years 1 Table 3: Comparison of simulated European call option prices and their standard errors defined in (23) for different time scale ε. Let “MC price” denote the sample mean obtained from the basic Monte Carlo simulation; “ MC+ CV price” denote the sample mean obtained from the Monte Carlo simulation using martingale control variates. Their standard errors are shown in parenthesis next to the option price estimates. Let V M C denote the sample variance obtained from the basic Monte Carlo method, and V M C+CV (PBS (σ̄)) denote the sample variance computed from the martingale control variates method with the option price approximation PBS (σ̄) in (13). ε 1/50 1/25 1/10 1 5 10 25 50 MC price (std err) 7.0771 (0.1080) 6.9862 (0.1072) 7.1530 (0.1085) 7.1588 (0.1065) 7.0064 (0.1059) 7.0785 (0.1090) 7.1145 (0.1101) 6.9737 (0.1095) MC+CV price (std err) 7.0883 (0.0064) 7.0646 (0.0069) 7.0466 (0.0079) 7.0244 (0.0096) 7.0236 (0.0092) 7.0577 (0.0080) 7.0729 (0.0068) 7.0737 (0.0063) 6 V M C /V M C+CV (PBS (σ̄)) 287 244 187 122 132 185 266 298 3 Numerical Test Results In this section, we will compare efficiencies for pricing European call option using control variates technique developed in the above, combined with Monte Carlo and quasi-Monte Carlo methods. 3.1 Quasi-Monte Carlo Methods We briefly describe the QMC methods here, and refer details to the literature. Consider the problem of estimating an s-dimensional integral Z f (x)dx, (15) µ= Cs where C s = [0, 1]s is the s-dimensional unit hypercube, f (x) is such a function that (at least) the integral (15) exists. The integral (15) can be expressed as the expectation Z f (x)dx =E[f (X)], (16) µ= [0,1]s where X is a random vector uniformly distributed on [0, 1]s . Thus, µ can be approximated by MC simulation. In fact, if X1 , X2 , · · ·, XN are iid and uniformly distributed on [0, 1]s , then the (sample) mean N 1 X µ bN = f (Xj ), (17) N j=1 is an unbiased estimator of µ, i.e., E(b µN ) = µ. The corresponding (sample) variance (assuming N > 1) " N #2 N N X X 1 X 1 2 2 2 σ bN = (18) [f (Xj ) − µ bN ] = [f (Xj )] − f (Xj ) N −1 (N − 1)N j=1 j=1 j=1 is an unbiased estimator of the variance of f (X): 2 2 2 2 Z σ = E [f (X)−µ] = E [f (X)] − µ = 2 Z f (x)dx− Cs 2 f (x)dx . (19) Cs Thus, lim µ bN = µ N →∞ (20) with probability one by the Strong Law of Large Numbers. Notice that the asymptotic convergence rate O( √1N ) of the MC method is independent of the dimension of the problem, the most important advantage of this method. Based on the Central Limit Theorem, we can construct the (1−p) confidence interval (CI) for a given confidence level p ∈ (0, 1): CI(p) = [b µN − zp/2 δN , µ bN + zp/2 δN ] (21) where δN = √σbNN is the (sample) standard error, and zp/2 > 0 is the quantile of the (one-dimensional) standard normal distribution such that Z zp/2 1 2 √ e−y /2 dy = 1 − p. 2π −zp/2 7 The main drawback of the MC method is that it converges slowly. Thus, various variance reduction methods were developed. Furthermore, the so called quasi-Monte Carlo methods were also developed. A QMC method to estimate integral (15) is a quadrature rule having the form µ≈ N 1 X f (xj ) N j=1 (22) where {x1 , · · ·, xN } ⊂ [0, 1]s is a deterministic sequence, instead of iid and uniformly distributed on [0, 1]s . Such a deterministic sequence, called a quasi-random sequence or low discrepancy sequence (LDS), satisfies certain properties. An LDS is usually more uniformly distributed over [0, 1]s than a pseudo-random sequence does. There are two classes of low-discrepancy sequences. One is called the digital net sequences, such as Halton’s sequence, Sobol’s sequence, Faure’s sequence, and Niederreiter’s (t, s)−sequence, etc. s A QMC method using this kind of LDS has convergence rate O( (logNN ) ) when estimating integral (15). The other class is the integration lattice rule points. This type of LDS is especially efficient for estimating multivariate integrals with periodic and smooth integrands, and it has convergence rate (log N )αs O( N α ), where α > 1 is a parameter related to the smoothness of the integrand. The monograph by Niederreiter [24] gives detailed information on digital net sequences and lattice rule points, while the monographs by Hua and Wang [15], and Sloan and Joe [28] describe lattice rules. Since their early successful applications in financial problems (see [18] and [26]), the quasi-Monte Carlo (QMC) Methods have been used more and more popular. For a class of low dimensional (say, less than 10) European style options, the QMC methods using lattice rules are especially efficient when estimating both the option prices and sensitivities, the lattice rules are much more efficient than the digital net sequences, while the digital net sequences are still more efficient than the MC method. For details, see [2] and [19]. L’Ecuyer also made contributions to lattice rules based on linear congruential generator. One of the features of this type of lattice rule points (referred to L’Ecuyer’s type lattice rule points, LTLRP, thereafter) is that it is easy to generate high dimensional LTLRP point sets with convergence rate comparable to digital net sequences. Details of LTLRP can be found in [21] and references therein. We will apply the LTLRP as well since our problem is high dimensional. Besides the above LDS, we also apply the Brownian bridge (BB) sampling technique to our test problems. Detailed information about Brownian bridge sampling can be found in [12]. To compare the efficiencies of different methods, we need a benchmark for fair comparison. If the exact value of the quantity to be estimated can be found, then we use the absolute error or relative error for comparison. Otherwise, we use the unbiased sample variance σ bn2 for comparison. For LDS sequences, we define the unbiased sample variance σ bn2 by introducing random shift as follows. Assume that we estimate µ = E[f (X)], s s where X is an s−dimensional random vector. Let {xi }m i=1 ⊂ I = [0, 1] be a finite LDS sequence and (j) {rj }nj=1 ⊂ I s be a finite sequence of random vectors. For each fixed j, we have a sequence {yi }m i=1 , (j) m where yi = xi + rj . Such a sequence still has the same convergence rate as {xi }i=1 does as proved in [29]. Denote 8 m 1 X (j) f (yi ), µj = m i=1 and n 1X µ bn = µj . n j=1 The unbiased sample variance is (assuming n > 1) Pn σ bn2 = 2 −µ bn ) = n−1 j=1 (µj n Pn 2 j=1 µj − P (n − 1)n n j=1 µj 2 . Then the standard error of the randomized LDS is defined as σ bn stderr = √ . n (23) And the variance of the MC method is calculated as usual based on the pseudo random point set {xi + rj } with mn points. Thus, we can easily construct the confidence interval for a given confidence level. The variance reduction ratio of a QMC method over the MC method is defined by variance reduction ratio = The sample variance of the MC method . The sample variance of the randomized QMC method (24) Here we do not include the CPU times for different methods. The main reason is that all the programs were run on a main frame using C++ under the UNIX operating system. There were many other jobs were also running when our programs were run on the main frame machine. Besides, we noticed that the differences of CPU times are not significant. We assume that the underlying asset S is given by (1). In our computations, we use C++ on Unix as our programming language. The pseudo random number generator we used is ran2() in [30]. In our comparisons, the sample sizes for MC method are 10240, 20480, 40960, 81920, 163840, and 327680, respectively; and those for Sobol’ sequence related methods are 1024, 2048, 4096, 8192, 16384, and 32768, respectively, each with 10 random shifts; and the sample sizes for LTLRP related methods are 1021, 2039, 4093, 8191, 16381, and 32749, respectively, and again, each with 10 random shifts. In the following examples, we divide the time interval [0, T ] into m = 128 subintervals. In the following Tables 7, 8 and 9, numbers without parentheses in columns (except column 1, which contains the numbers of points) are option values, and numbers within parentheses in the quasi-Monte Carlo methods are variance reduction ratios. Example 1: The European call option with payoff f (ST ) = max(ST − K, 0) = (ST − K)+ , where K is the strike price. We take input parameters as follows: S0 = $55, K = $50, r = 0.1 = 10%, T = 1 year, m1 = m2 = −0.8, ν1 = 0.5, ν2 = 0.8, ρ1 = ρ2 = −0.2, ρ12 = 0.0, y0 = z0 = −1.0. For ε and δ, we take two groups of values: ε = 1/5, δ = 1.0 and ε = 1/50, δ = 0.5. The results are listed in Tables 7 and 8, where MC+CV stands for Monte Carlo method using control variate technique, Sobol+BB means the quasi-Monte Carlo method using Sobol’ sequence with Brownian bridge sampling technique, LTLRP for QMC method using L’Ecuyer type lattice rule points, etc. Table 7: Comparison of simulated European call option values and variance reduction ratios for ε = 1/5, δ = 1.0 9 N 1024 2048 4096 8192 16384 32768 MC 12.034 12.216 12.166 12.145 12.144 12.206 MC+CV 12.217(16.7) 12.165(22.1) 12.135(20.3) 12.165(19.1) 12.169(26.3) 12.166(21.1) Sobol’ 12.240(1.6) 12.201(4.6) 12.190(1.1) 12.169(1.5) 12.255(0.2) 12.150(3.1) Sobol+CV 12.154(35.1) 12.197(127.2) 12.156(27.7) 12.160(30.2) 12.157(64.6) 12.167(78.5) Sobol+BB 12.030(2.2) 12.235(2.4) 12.141(2.2) 12.183(8.0) 12.179(3.9) 12.133(2.6) Sobol+CV+BB 12.159(55.3) 12.180(112.4) 12.156(31.1) 12.173(85.3) 12.155(161.0) 12.157(44.8) Table 7: Comparison of simulated European call option values and variance reduction ratios for ε = 1/5, δ = 1.0 (continued). N 1024 2048 4096 8192 16384 32768 LTLRP 12.376(1.2) 12.179(2.1) 12.251(3.2) 12.173(1.9) 12.156(1.6) 12.170(2.6) LTLRP+CV 12.201(25.4) 12.175(41.1) 12.155(60.8) 12.163(40.5) 12.170(180.1) 12.167(57.0) LTLRP+BB 12.100(4.7) 12.160(2.9) 12.083(2.5) 12.201(4.0) 12.138(2.6) 12.144(3.2) LTLRP+CV+BB 12.129(112.1) 12.196(5.1) 12.153(50.5) 12.168(135.9) 12.154(77.0) 12.152(137.3) Table 8: Comparison of time (in seconds) used in the simulation of the above European option N 1024 2048 4096 8192 16384 32768 MC 6 11 23 45 91 182 MC+CV 8 17 34 68 136 271 Sobol’ 6 11 22 45 88 177 Sobol+CV 7 15 29 61 120 240 Sobol+BB 6 11 23 47 93 186 Sobol+CV+BB 8 17 34 68 134 273 Table 8: Comparison of time (in seconds) used in simulation of the above European option (continued). N 1024 2048 4096 8192 16384 32768 LGLP 5 10 19 39 77 155 LGLP+CV 8 15 30 62 121 244 LGLP+BB 8 17 35 69 137 273 LGLP+CV+BB 6 12 23 48 95 191 From Tables 7, we observed the following facts. Using the control variate technique, the variance reduction ratios are around 20 for pseudo-random sequences. Without control variate, both Sobol’ sequence and L’Ecuyer type LATTICE RULE POINTS, even combined with Brownian bridge sampling technique, the variance reduction ratios are only a few times better than the MC sampling at most. However, when combined with control variate, the variance reduction ratios for the Sobol’ sequence increase from about 28 to 128; and the variance reduction ratios for the L’Ecuyer type LATTICE RULE POINTS increase from about 25 to 180. This implicitly indicates that the new controlled payoff e−rT (ST − K)+ − M0 (PBS ) is smoother than the original call payoff e−rT (ST − K)+ . It can be easily seen that under the Black-Scholes model with the constant volatility σ, the controlled payoff is exactly equal to the Black-Scholes option price PBS (0, S0 ; σ), which is C ∞ , while the original call payoff function is only continuous but not differentible. 10 Another interesting observation is that the variance reduction ratios do not always increase when the two low-discrepancy sequences are combined with control variate and Brownian bridge sampling, compared with when they are combined with control variate without Brownian bridge sampling. For this option, the Sobol’ sequence and the LTLRP are “comparable”: sometimes the Sobol’ sequence achieves better efficiency, and sometimes the other way around. Regarding time used in simulations, from Table 8 we observed that the time differences among methods without control variates are not significant, but the time differences between methods with and without control variates are not ignorable. Similar conclusions are true regarding time used in simulations for other cases. Table 9: Comparison of simulated plain vanilla European call option values and variance reduction ratios for 1/ε = 50.0, δ = 0.5 N 1024 2048 4096 8192 16384 32768 MC 11.97 11.97 12.03 11.97 11.99 12.04 MC+CV 12.03(30.3) 12.00(32.7) 11.97(34.6) 11.99(34.4) 12.00(33.2) 11.99(34.0) Sobol’ 12.24(2.2) 12.02(1.6) 12.03(1.3) 12.03(2.0) 12.02(1.4) 11.99(1.7) Sobol+CV 11.99(219.2) 12.00(300.4) 11.99(107.8) 11.99(104.9) 11.99(92.8) 12.00(156.6) Sobol+BB 11.75(2.9) 12.00(4.0) 11.97(2.7) 11.98(3.3) 11.99(6.5) 11.99(3.9) Sobol+CV+BB 11.98(66.6) 12.01(69.9) 11.98(133.5) 11.99(99.9) 11.99(119.4) 11.99(133.4) Table 9: Comparison of simulated plain vanilla European call option values and variance reduction ratios for 1/ε = 50.0, δ = 0.5 (continued). N 1024 2048 4096 8192 16384 32768 LTLRP 12.15(1.8) 12.06(2.7) 12.06(2.2) 11.99(3.2) 11.97(2.9) 11.99(6.7) LTLRP+CV 11.99(39.0) 11.99(75.4) 11.99(65.9) 12.00(173.9) 12.00(173.7) 11.99(227.1) LTLRP+BB 11.85(8.1) 11.99(3.9) 11.95(2.4) 12.04(2.0) 11.97(5.2) 12.00(13.2) LTLRP+CV+BB 11.96(172.8) 11.96(159.6) 11.99(37.0) 12.00(25.3) 11.98(45.6) 11.98(87.1) The conclusion is similar to that of Table 8, but variances are further reduced in this case. 4 Conclusion In this paper, we study the option pricing problems for European options under multi-factor stochastic volatility model. Based on the construction of martingale control variates by Fouque and Han [7], we implemented this method with Monte Carlo and randomized quasi-Monte Carlo methods. For complex stochastic volatility models, our tests show that quasi-Monte Carlo methods are still more efficient than Monte Carlo methods. Control variates combined with quasi-Monte Carlo methods are the most efficient ways in dealing with such option pricing problems. A theoretical problem induced from these numerical experiments is an error analysis to quantify the error of quasi-Monte Carlo methods under martingale control variate. This analysis can be used to explain the smooth effect by martingale control variate. We leave this problem in our future research. Acknowledgement: We are grateful to a financial support by the Institute for Mathematics and its Applications (IMA) at the University of Minnesota, USA. The work of C-H Han is supported by NSC grant 95-2115-M-007-017-MY2, Taiwan; and the work of Y. Lai was partially supported by an Natural 11 Sciences and Engineering Research Council (NSERC) of Canada grant. References [1] S. Alizadeh, M. Brandt and F. Diebold, Range-based estimation of stochastic volatility model, Journal of Finance, 57, 1047-1091, 2002. [2] P. Boyle, Y. Lai and K. S. Tan, Pricing Options Using Lattice Rules, North American Actuarial Journal, 9(3), 50-76, 2005. [3] M. Broadie and O. Kaya, Exact Simulation of Stochastic Volatility and other Affine Jump Diffusion Processes, Operations Research Vol 54(2), 217-231, 2005. [4] L. Clewlow and A. Carverhill, On the simulation of contingent claims, Journal of Derivatives 2:66-74, 1994. [5] J.-P. Fouque and C.-H. Han, Variance reduction for Monte Carlo methods to evaluate option prices under multi-factor stochastic volatility models, Quantitative Finance, Volume 4(5), 597606, 2004. [6] J.-P. Fouque and C.-H. Han, A control variate method to evaluate option prices under multi-factor stochastic volatility models, submitted, 2004. [7] J.-P. Fouque and C.-H. Han, A Martingale Control Variate Method for Option Pricing with Stochastic Volatility, to appear in ESAIM: Probability and Statistics. [8] J.-P. Fouque, G. Papanicolaou, R. Sircar, Derivatives in Financial Markets with Stochastic Volatility, Cambridge University Press, 2000. [9] J.-P. Fouque, G. Papanicolaou, R. Sircar and K. Solna, Singular perturbation in option pricing, SIAM Journal on Applied Mathematics, Vol. 63, No. 5, 1648-1665, 2003. [10] J.-P. Fouque, G. Papanicolaou, R. Sircar and K. Solna, Multi-scale Stochastic Volatility Asymptotics, SIAM Journal on Multi-scale Modeling and Simulation 2(1), 22-42, 2003. [11] E. Fournie, J. Lebuchoux and N. Touzi, Small noise expansion and important sampling, Asymptotic Analysis, 14, 361-376, 1997. [12] P. Glasserman, Monte Carlo Methods for Financial Engineering, Springer-Verlag, New York, 2003. [13] D. Heath and E. Platen, A variance reduction technique based on integral representations, Quantitative Finance, 2:362-369, 2002. [14] S. Heston, A closed-form solution for options with stochastic volatility with applications to bond and currency options, Review of Financial Studies, 6(2): 327-343, 1993. [15] L. Hua and Y. Wang, Applications of Number Theory in Numerical Analysis, Springer-Verlag, 1980. [16] J. Hull, Option, Futures and Other Derivatives, Fourth Edition, Prentice Hall, 2000. 12 [17] P. Jackel, Monte Carlo Methods in Finance, John Wiley & Sons Ltd. 2002. [18] C. Joy, P. Boyle and K. S. Tan, Quasi-Monte Carlo Methods in Numerical Finance, Management Science, 42(6), 926-938, 1996. [19] Y. Lai, Higher Rank Lattice Rules and Applications in Finance, Proceedings of the 17th IASTED International Conference on Modeling and Simulation, 522-527, 2006. [20] D. Lamberton and B. Lapeyre, Introduction to Stochastic Calculus Applied to Finance, Chapman & Hall/CRC, 2000. [21] P. L’Ecuyer and C. Lemieux, Variance Reduction via Lattice Rules, Management Science, 46(9), 1214-1235, 2000. [22] F. Longstaff and E. Schwartz, “Valuing American Options by Simulation: A Simple Least-Squares Approach,” Review of Financial Studies 14: 113-147, 2001. [23] G. Molina, C.-H. Han, and J.-P. Fouque, MCMC Estimation of Multi-scale Stochastic Volatility Models, Preprint, 2003. [24] H. Niederreiter, Random Number Generation and Quasi-Monte Carlo Methods, SIAM, Philadelphia, 1992. [25] M. Potters, J.-P. Bouchaud, and D. Sestovic, Hedged Monte-Carlo: low variance derivative pricing with objective probabilities, Physica A 289 (2001) 517-525. [26] S. Paskov and J. Traub, Faster Valuation of Financial Derovatives, Journal of Portfolio Management, 22(1), 113-120, 1995. [27] B. Pochart and J.-P. Bouchaud, Option pricing and hedging with minimum local expected shortfall, Science & Finance (CFM) working paper, Science & Finance, Capital Fund Management, 2003. [28] I. Sloan and S. Joe, Lattice Methods for Multiple Integration, Clarendon Press, Oxford, 1994. [29] B. Tuffin, On the use of low discrepancy sequences in Monte Carlo methods, Monte Carlo Methods and Applications Vol. 2 No. 4, 295-320, 1996. [30] W. H. Press, S. A. Teukolsky, W. T. Vetterling and B. P. Flannery, Numerical recipes in C : The Art of Scientific Computing, New York : Cambridge University Press, 1992. [31] A. Willard, Calculating prices and sensitivity for path-independent derivative securities in multifactor models, Journal of Derivatives, Vol. 5, No. 1, 45-61, 1997. 13
© Copyright 2025 Paperzz